Open access

Biomechanics of Cartilage and Osteoarthritis

Written By

Herng-Sheng Lee and Donald M. Salter

Submitted: 29 August 2014 Published: 01 July 2015

DOI: 10.5772/60011

From the Edited Volume

Osteoarthritis - Progress in Basic Research and Treatment

Edited by Qian Chen

Chapter metrics overview

2,366 Chapter Downloads

View Full Metrics

1. Introduction

The etiology of osteoarthritis (OA) is not known with certainty. Undoubtedly, many factors contribute to articular cartilage failure but consistently abnormal biomechanics appear central to the condition. Indeed it is recognized that mechanical loading that is either below or in excess of the physiology range leads cartilage degeneration. There are currently no cures for OA and no effective pharmacological treatments that slow or halt disease progression. Physical activity is one of the most widely prescribed non-pharmacological therapies for OA management, based on its ability to limit pain and improve physical function. The detailed mechanisms underlying these beneficial effects of exercise and physical therapy are largely unknown. Structural integrity is important for joint function and can be lost as a consequence of a range of physical, biomechanical and inflammatory factors. This chapter will overview how joint loading influences cartilage structure and how mechanical loading is perceived by chondrocytes resulting in cellular responses that are either chondroprotective or promoting inflammatory and catabolic responses initiating and progressing OA.

Advertisement

2. Mechanical integrity of joint structure

The synovial or diarthrodial joint allows movement between bones and permits transmission of mechanical loads. The mechanical integrity of the joint elements including articular cartilage, synovium, subchondral bone, joint capsule, ligaments and periarticular connective tissues cooperates to provide optimal function. Loss of mechanical integrity results in a range of pathological changes within the joint recognized as osteoarthritis.

In a synovial joint the articulating bone ends are covered by a thin, highly hydrated specialized connective tissue, articular cartilage. A high interstitial fluid content distinguishes cartilage from most other connective tissues and contributes to the mechanical properties of the tissue [1,2]. The major components of articular cartilage are type II collagen, proteoglycans, noncollagenous proteins and glycoproteins. Type II collagen forms the fibrillar meshwork which provides tensile strength [3-6] and entraps aggregating hydrophilic proteoglycans which help maintain the high tissue hydration [7-9]. Disruption of the fibrillar meshwork or loss of proteoglycan results in an in-ability of cartilage to distribute loads and its contribution to near frictionless joint movement leading, in time, to progressive structural damage and pathological features of OA.

Advertisement

3. Articular cartilage responses to mechanical loading in vivo

Mechanical loading within a physiological range is necessary to maintain joints in a healthy state. During normal daily activity articular cartilage is exposed to a range of mechanical forces during joint movement. Peak forces across the human hip and knee joints have been shown to reach 4 and 7 times body weight, respectively, during normal walking [10,11]. In vivo, mechanical loading is applied cyclically and the cells within cartilage, chondrocytes, are exposed to a composite of radial, tangential and shear stresses [12]. The effects of mechanical load bearing on the development and microscopic structure of the articular cartilage have been studied in some detail [13]. Elevated loading increases cartilage thickness, causes hypertrophy of the superficial zone chondrocytes, and increases the amount of intercellular matrix [14-17]. In normal human joints, load-bearing areas of the cartilage are thicker with a higher proteoglycan concentration and are mechanically stronger than non-load-bearing regions of the same joint [18-20]. Increasing weight-bearing of joints, in a variety of animal models, leads to elevation of proteoglycan content within articular cartilage [15,16,21-23]. In contrast, removal of load bearing leads to a reduction in proteoglycan content [13]. In a dog model, immobilization of a joint by placing a leg in a cast leads to cartilage atrophy, loss of Safranin O staining, and a decrease in its uronic acid content [21]. These changes are reversible on remobilization. Mechanical regulation is also an important factor for chondrogenesis and has been involved in the development of cell-based therapies for cartilage degeneration and disease [24].

3.1. Mechanical stress within articular cartilage

Articular cartilage is exposed to surprisingly large mechanical loads during joint movement. Using an instrumented hip prosthesis mechanical stresses have been measured in a 74-year-old female [25]. Rising from a chair, pressures in the hip joint cartilage can reach nearly 20 MPa and during walking, pressures cycle between atmospheric and 3-4 MPa at a frequency of around 1 Hz. With walking or running forces at the joint surface may vary from near zero to several times the whole body weight within a period of 1 second [10,11]. Loading of articular cartilage generates a combination of tensile, compressive and shear stress in the material. The tensile modulus of healthy human articular cartilage varies from 5-25 MPa, depending on the site of movement on the joint surface (i.e., high or low weight bearing regions), and the depth and orientation of the test specimen relative to the joint surface [4,26]. The compressive modulus varies from 0.4-2.0 MPa [27,28]. Articular cartilage responds to shearing forces by both stretching and deformation of the solid matrix. The dynamic shear modulus is within the range of 0.2-2.0 MPa for healthy bovine or canine cartilage [29-31]. These physiological stresses are important regulators of cartilage metabolism and integrity as mechanical loading serves to maintain fluid flow and ion phase function within the tissue and act to stimulate chondrocyte metabolism [32].

Advertisement

4. Articular cartilage explant and chondrocyte responses to mechanical loading in vitro

Rodan et al. [33] studied the effects of application of compressive forces (80 g/cm2) to chick tibial epiphyses (16-day-old embryos) in culture and found that glucose consumption reduced to half of controls. Twenty four hours after the release of pressure, glucose utilization again increased, approaching control levels. The same pressure also stimulated thymidine incorporation into DNA. Exposing chick tibial epiphyses to continuous compressive forces (60 g/cm2, equal to 5.865 kPa) caused a reduction of both cAMP and cGMP [34]. An equivalent hydrostatic pressure applied directly to cells isolated from chick tibial epiphyses also affects cyclic nucleotide accumulation [34]. Veldhuijzen et al. developed a model system that exposed cultured monolayer chondrocytes on the walls of tissue culture tubes to intermittent compressive forces of 12.8 kPa for 6 hours at a frequency of 0.3 Hz [35]. Contrary to the effect of continuous compressive forces, intermittent compressive forces caused a rise in levels of cAMP and a reduction in DNA synthesis. Palmoski and Brandt [36] studied the effects of both static and intermittent mechanical stress on full-thickness plugs of canine articular cartilage. When the plugs were exposed to compressive force using a regime of 60 sec on/60 sec off, glycosaminoglycan synthesis was reduced to 30-60% of controls. However, when a regime of 4 sec on/11 sec off was employed, the glycosaminoglycan synthesis increased by 34%, although protein synthesis and DNA, uronic acid, and water content remained unaltered indicating that different frequencies of cyclical strain produce differences in metabolic activity within chondrocytes.

Some models designed to test the effects of mechanical force on chondrocytes in vitro have focused on the effects of cell stretching. In these models there is usually deformation of a cell-laden, flexible membrane which can be regulated according to (1) the method of deformation of the membrane - by control of either the displacement or the force, and (2) the shape and mounting of the deformable membrane - either a circular membrane held at its periphery or a rectangular strip held at the two ends [37]. The devices utilized in the production of the displacement include (a) a vacuum driven diaphragm (silicone elastomer membrane, 2.5 mm in thickness) [38,39], (b) pin shaped displacement (silicone elastomer membrane, 0.254 mm in thickness) [40], (c) glass dome displacement (polytetrafluoroethylene membrane, 0.025 mm in thickness) [41,42], (d) air or fluid displacement (polyurethane membrane, 0.094 mm in thickness) [43-45], and (e) a circular groove displacement (silicone elastomer membrane, 0.076 mm in thickness; polyurethane membrane, 0.094 mm in thickness) [46,47]. The FlexercellTM strain unit [38] consists of a computer-controlled vacuum unit and a baseplate on which are held the culture dishes. These dishes have a flexible base. A vacuum is applied to the dishes via the baseplate. When a precise vacuum level is applied to the system, the bases of the culture plates are deformed by known percentage elongation that is maximal at the edge of the culture dish, but decreases towards the center. Using the system, straining the base of a culture dish leads to strain of the attached cultured cells. When the vacuum is released, the bases of the dishes return to their original conformation.

By stretching a supportive flexible membrane on which chondrocytes were cultured, Lee et al. [48] found that a cyclic 10% mechanical stretch for 8 hours increased glycosaminoglycan synthesis and decreased protein and collagen synthesis. DeWitt et al. [49] showed increased radiosulphate and 14C-glucosamine incorporation into glycosaminoglycans by chick epiphyseal chondrocytes in high density cultures subjected to a 5.5% strain at a frequency of 0.2 Hz. Protein synthesis after 24 hours mechanical strain remained unchanged. Using the FlexercellTM strain system, Fujisawa et al. [50] investigated the influence of cyclic tension force on the metabolism of cultured chondrocytes. Two levels of force (5 kPa or 15 kPa) and three frequencies 30 cycles/min (1 sec on/1 sec off), 0.5 cycles/min (1 sec on/119 sec off) and 0.25 cycles/min (1 sec on/239 sec off) were used. Both 5 and 15 kPa of high frequency cyclic mechanical stress for 48 hours significantly inhibited the syntheses of DNA, proteoglycan, collagen, and protein. The expression of interleukin-1, matrix metalloproteinase (MMP)-2 and MMP–9 mRNA were induced by 15 kPa of high frequency force. The production of pro- and active-MMP-9 which would lead to cartilage breakdown in vivo was also increased at this pressure and frequency of stimulation. Reducing the applied frequency decreased the inhibition of proteoglycan synthesis. Mechanical stretch producing 25% maximal elongation at a frequency of 0.05 Hz for 48 hours also induces the expression of high molecular weight heat shock protein (HSP) 105 kDa in the human chondrocytic cell line CS-OKB [51]. These findings suggest that the frequency of cyclic tension force is one of the key determinants of chondrocyte metabolism.

Using confocal microscopy and fluorescent techniques it is possible to monitor and measure intracellular calcium ion concentration in isolated articular chondrocytes subjected to controlled deformation with the edge of a glass micropipette [52]. Intracellular calcium ion concentration reaches a peak within 5 sec following 25% deformation of the cells and returns to baseline levels in 3-5 minutes. The immediate and transient increase of intracellular calcium waves is abolished by removing Ca2+ from the culture medium and is significantly reduced by the presence of gadolinium and amiloride, agents known to block mechanosensitive ion channels [53-56]. Inhibitors of intracellular Ca2+ release or agents known to cause cytoskeletal disruption including cytochalasin D and colchicine had no significant effect on the Ca2+ waves. The results indicate that mechanosensitive ion channels are upstream in the mechanotransduction signaling pathway, consistent with results obtained using electrophysiological parameters in the assessment of the cell response to cyclical mechanical strain [56].

The effects of fluid-induced shear stress on articular chondrocyte morphology and metabolism in vitro have also been investigated [57]. Fluid-induced shear stress (1.6 Pa = 16 dynes/cm2) was applied by cone viscometer to both normal human and bovine articular chondrocytes. Shear stress for 48 and 72 hours caused individual chondrocytes to elongate and align tangential to the direction of cone rotation. Glycosaminoglycan synthesis was increased 2-fold. After 48 hours of shear stress, the release of prostaglandin E2 (PGE2) was increased 10 to 20-fold. In human articular chondrocytes, mRNA levels for tissue inhibitor of metalloproteinase (TIMP) increased 9-fold in response to shear stress compared to controls. In contrast, mRNA levels for the neutral metalloproteinases, collagenase, stromelysin, and gelatinase, did not show significant changes [57].

Advertisement

5. Chondrocyte mechanoreceptors

The mechanisms by which chondrocytes recognize and respond to the various mechanical stresses encountered in mechanically loaded cartilage continue to be elaborated. A number of potential mechanoreceptors, sensory receptors that respond to vibration, stretching, pressure, or other mechanical stimuli have been identified in chondrocytes. In cartilage the extracellular matrix (ECM) transmits mechanical signals to the cell interior through changes in tension on the cell membrane. Integrins, stretch-activated ion channels, connexins, and primary cilia have each been identified as candidate mechanoreceptors [58-61].

5.1. Integrins

Integrins were first isolated, characterized, and sequenced from chick embryo fibroblast cDNA clones which encoded one subunit of the complex of membrane glycoproteins [62]. The name ‘integrin’ was proposed as the consequence of its role as an integral membrane complex involved in the transmembrane association between the ECM and the cytoskeleton. Integrins are a large family of α/β heterodimeric cell surface adhesion receptors that can bind a wide variety of ECM and cell surface ligands [63-69]. Most integrins bind ligands that are components of ECM, e.g. fibronectin, collagen, and vitronectin [70]; certain integrins can bind to soluble ligands (such as fibrinogen) or to counter receptors (such as intercellular adhesion molecules) on adjacent cells, leading to homo- or heterotypic aggregation [71,72]. Some of the integrin recognition sites in the ligands and counter receptors have been defined [73,74]. The first binding site to be defined was the Arg-Gly-Asp (RGD) containing sequence present in fibronectin, vitronectin, and a variety of other adhesive proteins [75,76].

Integrin expression by human chondrocytes has been investigated utilizing several techniques including immunohistochemistry, flow cytometry, immunoprecipitation, and northern blotting [77-79]. Normal adult human articular chondrocytes express α1β1, α3β1, α5β1, αVβ5, αVβ3, α6β1, α10β1, and α2β1, in which the α1β1 (receptor for collagen), α5β1 (receptor for fibronectin) and the αVβ5 (receptor for vitronectin) heterodimers are consistently expressed. In cartilage, integrin-ECM interactions are thought to be important in many aspects, physiological and pathological, of chondrocyte function, including adhesion, spreading, proliferation, signal transduction, biomechanical regulation, chondrogenesis, and gene expression [80]. Chondrocyte β1 integrin-ECM interactions are required for chondrocyte survival, matrix deposition and differentiation in models of chondrocyte development [81]. Integrins mediate chondrocyte adhesion to many ECM proteins including type II collagen, type VI collagen, fibronectin, laminin, and osteopontin [82-85]. Chondrocyte spreading and migration on type II collagen, type VI collagen, or fibronectin substrates in vitro is mediated by interactions with β1 integrins [86,87]. The interaction of the α5β1 integrin with fibronectin is necessary for adhesion, spreading, and proliferation of both chicken and rabbit chondrocytes [88,89]. In vitro chondrogenesis (the differentiation of blastemal cells to chondroblasts and the formation of cartilage matrix) is inhibited by the function blocking anti-β1 integrin antibodies [90]. Laminin-α3β1/α6β1 interactions are regulated by the ligand trend (depletion/reconstitution or competition experiments) during early chondrocyte differentiation [89]. Other integrin-mediated chondrocyte-matrix interactions include β1-matrix Gla protein, β3-bone sialoprotein II, β3-osteopontin, and α2β1-chondroadherin associations [84,86].

Integrins are involved in regulation of both cartilage matrix synthesis and integrity. Loss of integrin function inhibits type II collagen synthesis by chondrocytes in culture [91]. However, integrins are also involved in cartilage breakdown processes. Fibronectin fragments stimulate chondrolysis and decrease proteoglycan synthesis in cartilage explants through fibronectin-integrin dependent interactions [92]. Ligation of α5β1 integrin with fibronectin in cultured chondrocytes results in the formation of focal adhesion complexes comprising actin, focal adhesion kinase (FAK) and the G protein Rho [93]. Nitric oxide (NO), a potential mediator of events occurring in osteoarthritis, can inhibit the assembly of the intracellular activation complex and the subsequent upregulation of proteoglycan synthesis that occurs following ligation of α5β1 integrin to fibronectin [93].

Integrins also act as mechanoreceptors and transmit mechanical signals from the extracellular environment to the cytoskeleton [94-96]. Integrins can provide a gating function for signal transduction, by either supporting or prohibiting force transmission between ECM and the cytoskeleton [94]. Wang et al. [97], using a magnetic twisting device applied mechanical forces directly to cell surface receptors. They showed that the integrin subunit β1 induced focal adhesion formation and supported a force-dependent stiffening response, whereas nonadhesion receptors did not [97]. Maniotis et al. [98] reported that living cells and nuclei are hard-wired. When integrins were stimulated by micromanipulating bound microbeads or micropipettes, cytoskeletal filaments reoriented, nuclei distorted, and nucleoli redistributed along the axis of the applied tension field. These effects were specific for integrins, independent of cortical membrane distortion, and were mediated by direct linkages between the cytoskeleton and nucleus [98]. Using a similar magnetic drag force device, intracellular Ca2+ concentration was shown to increase when the α2 or the β1 integrin subunits were stressed, whereas mechanical loading of the transferrin receptor produced a significantly reduced effect [99]. An increase in tyrosine phosphorylation was observed as a reaction to mechanical stress on the β1-subunits of the integrin family, whilst stress to the transferrin or low density lipoprotein receptors which have no connection to the cytoskeleton did not produce this reaction [100,101].

Integrins are also modulated by mechanical stress [102]. When chondrosarcoma cells were exposed to mechanical stimulation, mRNA expression of the α5 integrin subunit was found to increase whilst expression of the β1, α2, and αV did not increase significantly [102]. The effect of mechanical stress on integrin subunit expression has also been investigated in cells cultured on type II collagen-coated dishes with a flexible base. Mechanical stress increased mRNA expression of the α2 integrin subunit whilst the levels of mRNA for integrin subunits β1, α1, α5, and αV showed no or only small changes [102]. It is likely that mechanical induced regulation of integrins is closely regulated and may be dependent on the nature of the mechanical force acting on the cell and specific mechanoreceptor stimulated.

5.2. Stretch-activated Ion channels

Stretch-activated or stretch-sensitive ion channels (SACs) open as a consequence of mechanical deformation of the cell membrane [103]. SACs are directly activated by mechanical forces applied along the plane of the cell membrane that induce membrane tension and distortion of the lipid bilayer. These result in conformational changes which alter opening or closing rates of the channels permitting ion flux [104]. Application of mechanical forces perpendicular to the cell membrane, as seen with hydrostatic pressure, appears to be less effective in activating SACs [103]. Activation of calcium permeable SACs leads to local increase in intracellular calcium levels and stimulation of downstream calcium-dependent intracellular signal cascades. SACs sensitive to gadolinium are necessary for load and fluid flow related cellular responses in both chondrocytes and bone cells.

5.3. Connexins

Connexins are widely expressed in connective tissue where networks of cells are seen such as in bone, tendon and, meniscus. They probably act to allow propagation of a mechanical stimulus through a tissue. They are a superfamily of twenty-one transmembrane proteins that form gap junctions and hemichannels [105]. Gap junctions allow continuity between cells permitting diffusion of ions, metabolites and small signaling molecules such as cyclic nucleotides and inositol derivatives. Cx43 is the most abundant connexin present in skeletal tissue. Conditional deletion of Cx43 reduces mineral apposition rate to mechanical loading and Cx43 hemichannels are important for fluid shear induced PGE2 and ATP release in osteocytic cells [106]. Connexins and gap junctions are present at the tip of osteocyte dendritic processes and between these processes and osteoblasts indicating their potential importance in permitting cell–cell communication among the osteocytic network although propagation may be only directed from osteocytes to osteoblasts [107]. Cx43 hemichannels are activated and mediate small molecule exchange between cells and matrix under mechanical stimulation in rat temporomandibular joint (TMJ) chondrocytes [108]. Cx32 and Cx43 are important in tenocyte mechanotransduction [109]. Cx32 junctions form a communication network arranged along the line of principal loading and stimulate collagen production in response to strain. Connexin dependent mechanotransducation may be important in adaptation of subchondral bone to mechanical loading of joints rather than having a major role in chondrocyte dependent mechanotransduction. Nevertheless recent studies suggest that primary cilia associated connexins may be functional in responses of chondrocytes to mechanical loading.

5.4. Primary Cilia

Primary cilia are solitary, immotile cilium present in most cells including chondrocytes and bone cells. They are microtubule-based organelles, growing from the centrosome to extend from the cell surface and contain large concentrations of cell membrane receptors, including integrins [110]. They function both as chemosensors and mechanosensors [111,112]. Bending of the cilium upon matrix deformation or with fluid flow is thought to cause cilium bending, pulling on associated matrix receptors and activation of the mechanoreceptors [113]. In addition to integrins, Cx43 hemichannels are also present on primary cilia and by regulating ATP release cilia and activation of purine receptors cilia-associated connexins may also be involved in mechanotransduction.

Advertisement

6. Chondrocyte mechanotransduction

Mechanoresponsiveness is a fundamental feature of all living cells [94,114,115]. Studies with cultured cells confirm that mechanical stresses can directly alter many cellular processes, including signal transduction, gene expression, growth, differentiation, and survival [116]. Wright et al. [117] investigated the effects of applied hydrostatic pressure on the transmembrane potentials of articular chondrocytes. These studies have been pivotal in identifying potential mechanotransduction pathways in both normal and osteoarthritic human chondrocytes. In this system cells in monolayer culture were exposed to an increase in hydrostatic pressure by placing culture dishes in a sealed perspex pressure chamber with a gas inlet and outlet. Nitrogen or helium gas was used to pressurize the cultures. A hyperpolarization of the chondrocyte plasma membrane was induced by cyclic pressurization (0.33 Hz, 120 mmHg for 20 min) whilst depolarization was induced by continuous pressure (120 mmHg, 20 min). For the frequencies tested, the maximum values for chondrocyte hyperpolarization occurred at approximately 0.3-0.4 Hz. The mechanical stimulation regime (0.33 Hz, 120 mmHg, 20 min), similar to that used by Veldhuijzen et al. [35], allowed identification of a number of integral components of the electrophysiological response providing insight into molecules and pathways activated in chondrocyte upon mechanical stimulation. By the use of pharmacological inhibitors, it was shown that the hyperpolarization response in cultured human chondrocytes induced by cyclic pressurization involved Ca2+ activated K+ channels and L-type calcium channels. Hyperpolarization was also produced by addition of the calcium ionophore A23187 to the culture medium showing that a rise in intracellular Ca2+ concentration within the cell could induce the response. Plasma membrane histamine H1 and H2 receptors, and β-adrenoreceptors did not appear to be involved in the hyperpolarization response. The studies also showed that the actin cytoskeleton, but not microtubules, was involved in the chondrocyte hyperpolarization response [117].

Subsequent studies identified that the electrophysiological response to cyclical pressurization was the result of deformation of the base of the culture dishes to which the chondrocytes were attached and therefore deformation (strain) on the chondrocytes rather than a direct effect of the increased hydrostatic pressure on chondrocytes [56]. The hyperpolarization response was proportional to the microstrain to which cells were subjected and did not occur when chondrocytes were subjected to cyclical pressurization in rigid glass culture dishes or when the plastic dishes were positioned in the pressurization chamber so as to avoid deformation of the base of the culture dish [56].

Experiments undertaken to identify the source of intracellular calcium that activated the SK channels leading to hyperpolarization demonstrated a requirement for extracellular calcium and activity of L-type calcium channels [118-120]. Thapsigargin which raises intracellular Ca2+ by inhibition of Ca2+-ATPase in endoplasmic reticulum [121-123] caused hyperpolarization independent of mechanical strain but further hyperpolarization of the cells occurred after cyclical pressurization further supporting the idea that mechanically induced chondrocyte hyperpolarization is dependent on intracellular free Ca2+ levels [56]. In addition, TRPV4-mediated Ca2+ signaling has been demonstrated to play a central role in the transduction of mechanical signals to support cartilage extracellular matrix maintenance [124].

6.1. Intracellular signal cascades activated by mechanical stimulation

Stimulation of connective tissue cell mechanoreceptors is followed by generation of the secondary messenger molecules and activation of a cascade of downstream signaling events that regulate gene expression and cell function. Many intracellular signaling pathways are known to be activated by mechanical forces applied to tissues and cells including heterotrimeric guanine nucleotide binding proteins (G-proteins), protein kinases and transcription factors. These pathways that regulate tissue modelling/remodelling may be activated directly as a consequence of mechanoreceptor signaling or indirectly following production of autocrine/paracrine acting molecules.

PKB/Akt is a protein family of serine/threonine kinases that have multiple roles including inhibition of apoptosis by phosphorylation and inactivation of pro-apoptotic factors. Integrin-dependent activation of phosphoinositide3 kinase (PI3 kinase) by mechanical forces regulates PKB activity and can inhibit cell death. Inactivation of the PI3-K/PKB pathway may be important in deleterious effects of mechanical overloading of cartilage and bone loss in response to withdrawal of loading [125]. The activity of mammalian target of rapamycin (mTOR) may be an essential mechanotransduction component modulated by SH2-containing protein tyrosine phosphatase 2 and is required for cartilage development [126].

Mitogen-activated protein kinases (MAPKs) regulate multiple cellular activities, such as gene expression, mitosis, differentiation, and cell survival/apoptosis. ERK1/2, JNK and p38, of critical importance in regulation of matrix protein and protease gene expression have each been shown to be activated in chondrocytes following mechanical stimulation [127]. Mechanical stimuli may activate different MAPKs and through this mechanism differential cellular responses may occur. MAPK responses may also be cell type dependent. Mechanical stimulation induced ERK1/2 activation in bone cells requires calcium-dependent ATP release whilst in cartilage activation, under certain circumstances, is dependent on FGF-2 rather than through integrin mechanoreceptors [128]. Tyrosine phosphorylation of focal adhesion kinase (pp125FAK), beta-catenin, and paxillin following mechanical stimulation is also recognized in human articular chondrocytes [129].

In bone cells NF-κB, a protein complex that acts as a transcription factor, is directly stimulated by mechanical stimulation is dependent on intracellular calcium release [130]. In chondrocytes biomechanical signals within the physiological range block NF-κB activity and proinflammatory chondrocyte responses [131]. Mechanical stimuli that induce catabolic rather than anabolic responses in chondrocytes induce rapid nuclear translocation of NF-κB subunits p65 and p50 in a similar manner to IL-1β [132].

6.2. Growth factors and autocrine/paracrine signaling in mechanotransduction

As part of the cellular response to mechanical stimulation mechanosensitive connective tissue cells release a range of soluble mediators. These may be present in the cell and available for immediate release, or secretion may depend de novo synthesis by enzymatic activity or transcriptional activation and protein production. These mediators, including prostaglandins, nitric oxide, cytokines, growth factors, and neuropeptides are involved in downstream tissue modelling and remodelling responses initiated by the mechanosensitive cells or other effector cells. Production of soluble mediators by connective tissue cells in response to mechanical stimulation however may also be intrinsic to mechanotransduction pathways. Autocrine and paracrine activity allows increased regulation of the cellular response to mechanical stimuli by permitting cross talk between different components of a mechanotransduction cascade. As the cellular responses to mechanical stimuli and soluble mediators activate similar signal cascades inducing either anabolic or catabolic responses, it would be expected that they may be antagonistic, additive or synergistic. Anabolic cytokines and growth factors enhance production of matrix under mechanical loading conditions whilst anabolic mechanical stimuli antagonize the effects of catabolic cytokines such as IL-1β [133].

Prostaglandins, predominantly PGE2, NO and ATP are produced when bone cells and chondrocytes are mechanically stimulated. Prostaglandin production is integrin dependent requiring an intact cytoskeleton and activation of SACs, PKC, and PLA2. In cartilage PGE2 induced by mechanical loads is catabolic. Mechanical loading of chondrocytes by physiological stimuli inhibits production of PGE2 and NO whereas damaging loading induces PGE2 release [134]. Following mechanical stimulation bone cells and chondrocytes release ATP which can bind and activate purinergic receptors on these and adjacent cells. Both metabotropic P2Y receptors and ionotropic P2X receptors, have been shown to be involved in mechanical load activated signal cascades in chondrocytes and bone cells and may have physiological roles [135].

IL-4 and IL-1β autocrine/paracrine activity is seen in the integrin-dependent mechanotransduction cascade of chondrocytes (IL-4 and IL-1β) and bone cells (IL-1β) to mechanical stimulation [136,137]. These molecules are secreted within 20 minutes of the onset of mechanical stimulation, suggesting release from preformed stores. IL-4 release relies on secretion of the neuropeptide substance P which binds to its NK1 receptor. Both IL-4 and substance P are necessary but not sufficient for the increased expression of aggrecan mRNA and decrease in MMP3 mRNA induced by the mechanical stimulus suggesting cross talk with other mechanosensitive signaling pathways. IL-1β is involved in the early mechanotransduction pathway of both osteoarthritic chondrocytes and human trabecular bone derived cells [138]. Mechanical loading may also induce release or activation of sequestered growth factors in extracellular matrix which will then act on near-by resident connective tissue cells. Basic fibroblast growth factor (FGF2) is a possible mediator of mechanical signaling in cartilage through such a mechanism [128]. Dynamic compression of porcine cartilage induces release of FGF2 with activation of ERK MAP kinase, synthesis and secretion of TIMP-1. In contrast FGF2 production by bovine cartilage is inhibited by 1 hour of compressive stress of 20 MPa [139]. This mechanical induced suppression of FGF2 is blocked by IL-4 indicating further roles for this pleiotropic cytokine in the regulation of chondrocyte responses to mechanical stimulation.

Advertisement

7. Mechanical loading and osteoarthritis

Abnormal mechanical loading is associated with osteoarthritis [140]. Most animal models of OA are mechanically induced, for example, by introducing joint instability by anterior cruciate ligament section [22] or by altering the loading across the joint by menisectomy [141]. These changes in joint loading affect cartilage structure and chondrocyte activity within days of the procedure, and may eventually result in complete loss of cartilage [142]. When cartilage matrix is lost or made deficient as a consequence of direct physical effects or proteolytic digestion the articular cartilage loses its mechanical function. The tensile modulus has been shown to decrease by as much as 90%, reflecting damage to the cartilage matrix network [26]. Animal studies, for example, have shown that the tensile modulus of canine knee articular cartilage was reduced after one month of immobilization [143]. In the dog, severe OA lesions in the knee joint have been produced by treadmill exercise after the limb was immobilized for several weeks [144]. The compressive modulus also decreases with increasing severity of degeneration [27]. Other joint tissues (e.g., anterior cruciate ligament) also undergo similar changes of tensile and compressive modulus in an experimental OA model [145].

Mechanical loading out with that which joint tissues can normally withstand or are required to maintain a healthy state is central to the development of OA. OA arises when there is an imbalance between the mechanical forces within a joint and the ability of the cartilage to withstand these forces. This arises in two situations. In the first normal articular cartilage is exposed to abnormal mechanical loads whereas in the other the articular cartilage is fundamentally defective with biomaterial properties that are insufficient to withstand normal load bearing. Risk factors associated with development of OA may have influences in either one or both of these scenarios. The accumulation of advanced glycation end products (AGEs) in ECM with age results in a more brittle collagen network that is less able to withstand normal loads, again leading to cartilage degeneration.

The mechanisms by which abnormal mechanical loading may influence chondrocyte function to promote cartilage breakdown are beginning to be understood. Chondrocytes from osteoarthritic cartilage share mechanoreceptors with chondrocytes from normal chondrocytes [146,147]. Whilst activation of these receptors and downstream signaling cascades such as FAK, PKC, JAK/STAT and MAP kinases[148,149] result in pro-anabolic activity in normal chondrocytes activity through release of locally acting mediators that include the anti-inflammatory cytokine IL-4 and the neuropeptide substance P [136,150]. However the response of chondrocytes is aberrant with production of proinflammatory cytokines such as IL-1 and TNF-α which increase production of MMPs and aggrecanases, further accelerating disease progression and attenuating cartilage repair [151-156].

7.1. Altered responses to mechanical stimulation in osteoarthritic chondrocytes

Chondrocytes from osteoarthritic cartilage show a membrane depolarization response to IL-4 that is inhibited by functional receptor antibodies. It is unclear why chondrocytes from osteoarthritic cartilage should show differences in their response to 0.33 Hz mechanical stimulation and recombinant IL-4. This may be result of a general phenotypic change seen in OA chondrocytes in which the cells are resident in a pro-inflammatory, catabolic environment. Indeed the observation that mechanical stimulation in osteoarthritis may result in production of proinflammatory mediators is supported by the findings that α5β1 integrin ligation increases production of IL-1β by osteoarthritic human chondrocytes with subsequent induction of nitric oxide, PGE2, IL-6, and IL-8 [157]. These cytokines will inhibit anabolic responses and increase cartilage matrix breakdown by MMPs. This may be through direct mechanisms or by interfering with integrin signaling. Expression and function of molecules such as members of the SOCS (suppressors of cytokine signaling) which regulate cytokine signaling pathways [158] may also be implicated. These molecules modulate intracellular signals stimulated by IL-4 including JAK/STAT activation. SOCS-1 has been shown to bind to and inhibit kinase activity of JAK family members and inhibit IL-4 induced activation of JAK1 and STAT6. SOCS-3 has been shown mediate IL-1β inhibition of STAT5 activity. These and other regulators of STAT transcription factor signaling may be responsible for modulation of IL-4 dependent responses of chondrocytes in osteoarthritic cartilage to mechanical stimulation.

Advertisement

8. Summary

A healthy synovial joint requires exposure to mechanical loads within a physiological range. Osteoarthritis develops when joints are subjected to mechanical loads which they are not biomechanically conditioned to withstand. This may be because the loads are excessive due to obesity or joint malalignment or a consequence of intrinsic or acquired biomechanical weakness of joint tissues such as seen when cartilage proteoglycans are depleted secondary to synovial inflammation. Abnormal mechanical loads may have direct physical effects on joint tissues including cartilage but increasingly knowledge of the pathological process within the osteoarthritic joint indicate that chondrocytes regulated catabolic processes are of prime importance in cartilage degradation. The mechanisms by which chondrocytes recognize mechanical loads and how these mechanical stimuli are transduced into biochemical responses which subsequently lead to altered gene expression and cell function is being increasing understood (Figure 1). Knowledge of how anabolic and catabolic signaling cascades are differentially regulated in response to physiological and pathological mechanical stimuli will enable future strategies to be developed to prevent and treat the progression of cartilage pathology in osteoarthritis.

Figure 1.

The major mechanotransduction components in chondrocytes. The integrin, connexin, and stretch-activated ion channel mechanoreceptors are stimulated by the mechanical forces transduced via the extracellular matrix (ECM). Downstream transduction pathways involve the cytoskeleton and signaling molecules, including FAK, PKC, PI3K, PKB, NF-kB, and MAPK, which act to regulate gene expression, cell function and survival/apoptosis. The release and paracrine/autocrine activity of the anti-inflammatory cytokine IL-4 has beneficial effects in regulating an anabolic response with enhanced expression of aggrecan and inhibition of MMP expression. In contrast production of IL-1β, as seen in OA, has a catabolic outcome with activation of pathways resulting in increased expression of COX2 and MMPs.

References

  1. 1. Brand R A. Joint lubrication. In: Brand R A, Albright J A (eds) The Scientific basis of orthopaedics. Appleton & Lange: Norwalk, Conn.; 1987. 373-389.
  2. 2. McCutchen C E. Lubrication of joints. In: Sokoloff L (ed.) The Joints and synovial fluid. Academic Press: New York; 1987. 437-483.
  3. 3. Harris E D, Parker H G, Radin E L, Krane S M. Effects of proteolytic enzymes on structural and mechanical properties of cartilage. Arthritis and rheumatism 1972;15(5): 497-503.
  4. 4. Kempson G E, Muir H, Pollard C, Tuke M. The tensile properties of the cartilage of human femoral condyles related to the content of collagen and glycosaminoglycans. Biochimica et biophysica acta 1973;297(2): 456-472.
  5. 5. Kempson G. The mechanical properties of articular cartilage. In: Sokoloff L (ed.) The Joints and synovial fluid. Academic Press: New York; 1978. 177-238.
  6. 6. Mayne R, Irwin M H. Collagen types in cartilage. In: Kuettner K E, Schleyerbach R, Hascall V C (eds) Articular cartilage biochemistry. Raven Press: New York; 1986. 23-38.
  7. 7. Comper W D, Laurent T C. Physiological function of connective tissue polysaccharides. Physiological reviews 1978;58(1): 255-315.
  8. 8. Greenwald R A, Moy W W, Seibold J. Functional properties of cartilage proteoglycans. Seminars in arthritis and rheumatism 1978;8(1): 53-67.
  9. 9. Hascall V C. Interaction of cartilage proteoglycans with hyaluronic acid. Journal of supramolecular structure 1977;7(1): 101-120.
  10. 10. Paul J. Joint kinetics. In: L S (ed.) The joints and synovial fluid. Academic press: NewYork; 1980. 139-176.
  11. 11. Seireg A, Arvikar. The prediction of muscular lad sharing and joint forces in the lower extremities during walking. Journal of biomechanics 1975;8(2): 89-102.
  12. 12. van Kampen G, van de Stadt R. Cartilage and chondrocyte responses to mechanical loading in vitro. In: HJ H, I K, M T, AM S, K P, J J (eds) Joint Loading: biology and health of articular structures; 1987. 112-125.
  13. 13. Helminen H, Jurvelin J, Kiviranta I, Paukkonen K, Saamanen A, Tammi M. Joint loading effects on articular cartilage: a historical review. In: Helminen H, Kiviranta I, Tammi M, Sasmanen A, Paukkonen K, Jurvelin J (eds) Joint Loading: biology and health of articular structures; 1987. 1-46.
  14. 14. Jurvelin J, Helminen H J, Lauritsalo S, Kiviranta I, Saamanen A M, Paukkonen K, Tammi M. Influences of joint immobilization and running exercise on articular cartilage surfaces of young rabbits. A semiquantitative stereomicroscopic and scanning electron microscopic study. Acta anatomica 1985;122(1): 62-68.
  15. 15. Eggli P S, Hunziker E B, Schenk R K. Quantitation of structural features characterizing weight- and less-weight-bearing regions in articular cartilage: a stereological analysis of medial femoral condyles in young adult rabbits. The Anatomical record 1988;222(3): 217-227.
  16. 16. Kiviranta I, Tammi M, Jurvelin J, Saamanen A M, Helminen H J. Moderate running exercise augments glycosaminoglycans and thickness of articular cartilage in the knee joint of young beagle dogs. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1988;6(2): 188-195.
  17. 17. Guilak F, Bachrach N M. Compression-induced changes in chondrocyte shape and volume determined in situ using confocal microscopy. Trans Orthopaedic Research Society 1993;18: 619.
  18. 18. Bjelle A. Content and composition of glycosaminoglycans in human knee joint cartilage. Variation with site and age in adults. Connective tissue research 1975;3(2): 141-147.
  19. 19. Roberts S, Weightman B, Urban J, Chappell D. Mechanical and biochemical properties of human articular cartilage in osteoarthritic femoral heads and in autopsy specimens. The Journal of bone and joint surgery. British volume 1986;68(2): 278-288.
  20. 20. Slowman S D, Brandt K D. Composition and glycosaminoglycan metabolism of articular cartilage from habitually loaded and habitually unloaded sites. Arthritis and rheumatism 1986;29(1): 88-94.
  21. 21. Palmoski M, Perricone E, Brandt K D. Development and reversal of a proteoglycan aggregation defect in normal canine knee cartilage after immobilization. Arthritis and rheumatism 1979;22(5): 508-517.
  22. 22. Muir H, Carney S. Pathological and biochemical changes in cartilage and other tissues of the canine knee resulting from induced joint instability. In: Helminen H, Kiviranta I, Tammi M, Sasmanen A, Paukkonen K, Jurvelin J (eds) Joint Loading: biology and health of articular structures; 1987. 47-63.
  23. 23. Caterson B, Lowther D A. Changes in the metabolism of the proteoglycans from sheep articular cartilage in response to mechanical stress. Biochimicaet Biophysica Acta 1978: 412-422.
  24. 24. O'Conor C J, Case N, Guilak F. Mechanical regulation of chondrogenesis. Stem cell research & therapy 2013;4(4): 61.
  25. 25. Hodge W A, Fijan R S, Carlson K L, Burgess R G, Harris W H, Mann R W. Contact pressures in the human hip joint measured in vivo. Proceedings of the National Academy of Sciences of the United States of America 1986;83(9): 2879-2883.
  26. 26. Akizuki S, Mow V C, Muller F, Pita J C, Howell D S, Manicourt D H. Tensile properties of human knee joint cartilage: I. Influence of ionic conditions, weight bearing, and fibrillation on the tensile modulus. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1986;4(4): 379-392.
  27. 27. Armstrong C G, Mow V C. Variations in the intrinsic mechanical properties of human articular cartilage with age, degeneration, and water content. The Journal of bone and joint surgery. American volume 1982;64(1): 88-94.
  28. 28. Athanasiou K A, Agarwal A, Dzida F J. Comparative study of the intrinsic mechanical properties of the human acetabular and femoral head cartilage. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1994;12(3): 340-349.
  29. 29. Zhu W, Mow V C, Koob T J, Eyre D R. Viscoelastic shear properties of articular cartilage and the effects of glycosidase treatments. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1993;11(6): 771-781.
  30. 30. Spirt A A, Mak A F, Wassell R P. Nonlinear viscoelastic properties of articular cartilage in shear. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1989;7(1): 43-49.
  31. 31. Setton L A, Mow V C, Howell D S. Changes in the shear properties of canine knee cartilage resulting from anterior cruciate transection. Journal of Orthopaedic Research 1995.
  32. 32. Mow V C, Wang C C, Hung C T. The extracellular matrix, interstitial fluid and ions as a mechanical signal transducer in articular cartilage. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 1999;7(1): 41-58.
  33. 33. Rodan G A, Mensi T, Harvey A. A quantitative method for the application of compressive forces to bone in tissue culture. Calcified tissue research 1975;18(2): 125-131.
  34. 34. Rodan G A, Bourret L A, Harvey A, Mensi T. Cyclic AMP and cyclic GMP: mediators of the mechanical effects on bone remodeling. Science 1975;189(4201): 467-469.
  35. 35. Veldhuijzen J P, Bourret L A, Rodan G A. In vitro studies of the effect of intermittent compressive forces on cartilage cell proliferation. Journal of cellular physiology 1979;98(2): 299-306.
  36. 36. Palmoski M J, Brandt K D. Effects of static and cyclic compressive loading on articular cartilage plugs in vitro. Arthritis and rheumatism 1984;27(6): 675-681.
  37. 37. Schaffer J L, Rizen M, L'Italien G J, Benbrahim A, Megerman J, Gerstenfeld L C, Gray M L. Device for the application of a dynamic biaxially uniform and isotropic strain to a flexible cell culture membrane. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1994;12(5): 709-719.
  38. 38. Banes A, Gilbert J, Taylor D, Monbureau O. A new vacuum-operated stress-providing instrument that applies static or variable duration cyclic tension or compression to cells in vitro. Journal of cell science 1985;75: 35-42.
  39. 39. Gilbert J, Banes A, Link G. Characterization of the surface strain applied to cyclically stretched cells in vitro. Trans Orthopaedic Research Society 1989;14: 249.
  40. 40. Vandenburgh H H. A computerized mechanical cell stimulator for tissue culture: effects on skeletal muscle organogenesis. In vitro cellular & developmental biology : journal of the Tissue Culture Association 1988;24(7): 609-619.
  41. 41. Hasegawa S, Sato S, Saito S, Suzuki Y, Brunette D M. Mechanical stretching increases the number of cultured bone cells synthesizing DNA and alters their pattern of protein synthesis. Calcified tissue international 1985;37(4): 431-436.
  42. 42. Andersen K L, Norton L A. A device for the application of known simulated orthodontic forces to human cells in vitro. Journal of biomechanics 1991;24(7): 649-654.
  43. 43. Brighton C T, Strafford B, Gross S B, Leatherwood D F, Williams J L, Pollack S R. The proliferative and synthetic response of isolated calvarial bone cells of rats to cyclic biaxial mechanical strain. The Journal of bone and joint surgery. American volume 1991;73(3): 320-331.
  44. 44. Williams J L, Chen J H, Belloli D M. Strain fields on cell stressing devices employing clamped circular elastic diaphragms as substrates. Journal of biomechanical engineering 1992;114(3): 377-384.
  45. 45. Belloli D M, Williams J L, Park J Y. Calculated strain fields on cell stressing devices employing circular diaphragms as substrates. Trans Orthopaedic Research Society 1991;16: 399.
  46. 46. Hung C T, Williams J L. A method for inducing equi-biaxial and uniform strains in elastomeric membranes used as cell substrates. Journal of biomechanics 1994;27(2): 227-232.
  47. 47. Schaffer J, Newman N, Gray M, Gerstenfeld L. Osteoblast phenotypic expression after mechanical perturbation with a dynamic and isotropic and biaxially uniform strain. Journal of Bone and Mineral Research 1993;8: S367.
  48. 48. Lee R C, Rich J B, Kelley K M, Weiman D S, Mathews M B. A comparison of in vitro cellular responses to mechanical and electrical stimulation. The American surgeon 1982;48(11): 567-574.
  49. 49. De Witt M T, Handley C J, Oakes B W, Lowther D A. In vitro response of chondrocytes to mechanical loading. The effect of short term mechanical tension. Connective tissue research 1984;12(2): 97-109.
  50. 50. Fujisawa T, Hattori T, Takahashi K, Kuboki T, Yamashita A, Takigawa M. Cyclic mechanical stress induces extracellular matrix degradation in cultured chondrocytes via gene expression of matrix metalloproteinases and interleukin-1. Journal of biochemistry 1999;125(5): 966-975.
  51. 51. Chano T, Tanaka M, Hukuda S, Saeki Y. Mechanical stress induces the expression of high molecular mass heat shock protein in human chondrocytic cell line CS-OKB. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2000;8(2): 115-119.
  52. 52. Guilak F, Zell R A, Erickson G R, Grande D A, Rubin C T, McLeod K J, Donahue H J. Mechanically induced calcium waves in articular chondrocytes are inhibited by gadolinium and amiloride. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1999;17(3): 421-429.
  53. 53. Jorgensen F, Ohmori H. Amiloride blocks the mechano-electrical transduction channel of hair cells of the chick. The Journal of physiology 1988;403: 577-588.
  54. 54. Yang X C, Sachs F. Block of stretch-activated ion channels in Xenopus oocytes by gadolinium and calcium ions. Science 1989;243(4894 Pt 1): 1068-1071.
  55. 55. Hamill O P, Lane J W, McBride D W, Jr. Amiloride: a molecular probe for mechanosensitive channels. Trends in pharmacological sciences 1992;13(10): 373-376.
  56. 56. Wright M, Jobanputra P, Bavington C, Salter D M, Nuki G. Effects of intermittent pressure-induced strain on the electrophysiology of cultured human chondrocytes: evidence for the presence of stretch-activated membrane ion channels. Clinical science 1996;90(1): 61-71.
  57. 57. Smith R L, Donlon B S, Gupta M K, Mohtai M, Das P, Carter D R, Cooke J, Gibbons G, Hutchinson N, Schurman D J. Effects of fluid-induced shear on articular chondrocyte morphology and metabolism in vitro. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1995;13(6): 824-831.
  58. 58. Roemer F W, Guermazi A, Niu J, Zhang Y, Mohr A, Felson D T. Prevalence of magnetic resonance imaging-defined atrophic and hypertrophic phenotypes of knee osteoarthritis in a population-based cohort. Arthritis and rheumatism 2012;64(2): 429-437.
  59. 59. Zeng Q Y, Chen R, Darmawan J, Xiao Z Y, Chen S B, Wigley R, Le Chen S, Zhang N Z. Rheumatic diseases in China. Arthritis research & therapy 2008;10(1): R17.
  60. 60. Lin J, Li R, Kang X, Li H. Risk factors for radiographic tibiofemoral knee osteoarthritis: the wuchuan osteoarthritis study. International journal of rheumatology 2010;2010: 385826.
  61. 61. Pai Y C, Rymer W Z, Chang R W, Sharma L. Effect of age and osteoarthritis on knee proprioception. Arthritis and rheumatism 1997;40(12): 2260-2265.
  62. 62. Tamkun J W, DeSimone D W, Fonda D, Patel R S, Buck C, Horwitz A F, Hynes R O. Structure of integrin, a glycoprotein involved in the transmembrane linkage between fibronectin and actin. Cell 1986;46(2): 271-282.
  63. 63. Albelda S M, Buck C A. Integrins and other cell adhesion molecules. FASEB journal : official publication of the Federation of American Societies for Experimental Biology 1990;4(11): 2868-2880.
  64. 64. Arnaout M A. Structure and function of the leukocyte adhesion molecules CD11/CD18. Blood 1990;75(5): 1037-1050.
  65. 65. Hemler M E. VLA proteins in the integrin family: structures, functions, and their role on leukocytes. Annual review of immunology 1990;8: 365-400.
  66. 66. Springer T A. Adhesion receptors of the immune system. Nature 1990;346(6283): 425-434.
  67. 67. Springer T A. The sensation and regulation of interactions with the extracellular environment: the cell biology of lymphocyte adhesion receptors. Annual review of cell biology 1990;6: 359-402.
  68. 68. Ruoslahti E. Integrins. The Journal of clinical investigation 1991;87(1): 1-5.
  69. 69. Aplin A E, Howe A, Alahari S K, Juliano R L. Signal transduction and signal modulation by cell adhesion receptors: the role of integrins, cadherins, immunoglobulin-cell adhesion molecules, and selectins. Pharmacological reviews 1998;50(2): 197-263.
  70. 70. Humphries M J, Newham P. The structure of cell-adhesion molecules. Trends in cell biology 1998;8(2): 78-83.
  71. 71. Kuhn K, Eble J. The structural bases of integrin-ligand interactions. Trends in cell biology 1994;4(7): 256-261.
  72. 72. Gahmberg C G, Valmu L, Fagerholm S, Kotovuori P, Ihanus E, Tian L, Pessa-Morikawa T. Leukocyte integrins and inflammation. Cellular and molecular life sciences : CMLS 1998;54(6): 549-555.
  73. 73. Komoriya A, Green L J, Mervic M, Yamada S S, Yamada K M, Humphries M J. The minimal essential sequence for a major cell type-specific adhesion site (CS1) within the alternatively spliced type III connecting segment domain of fibronectin is leucine-aspartic acid-valine. The Journal of biological chemistry 1991;266(23): 15075-15079.
  74. 74. Newham P, Craig S E, Seddon G N, Schofield N R, Rees A, Edwards R M, Jones E Y, Humphries M J. Alpha4 integrin binding interfaces on VCAM-1 and MAdCAM-1. Integrin binding footprints identify accessory binding sites that play a role in integrin specificity. The Journal of biological chemistry 1997;272(31): 19429-19440.
  75. 75. Pierschbacher M D, Ruoslahti E. Cell attachment activity of fibronectin can be duplicated by small synthetic fragments of the molecule. Nature 1984;309(5963): 30-33.
  76. 76. Hynes R O. Integrins: versatility, modulation, and signaling in cell adhesion. Cell 1992;69(1): 11-25.
  77. 77. Salter D M, Hughes D E, Simpson R, Gardner D L. Integrin expression by human articular chondrocytes. British journal of rheumatology 1992;31(4): 231-234.
  78. 78. Woods V L, Jr., Schreck P J, Gesink D S, Pacheco H O, Amiel D, Akeson W H, Lotz M. Integrin expression by human articular chondrocytes. Arthritis and rheumatism 1994;37(4): 537-544.
  79. 79. Camper L, Hellman U, Lundgren-Akerlund E. Isolation, cloning, and sequence analysis of the integrin subunit alpha10, a beta1-associated collagen binding integrin expressed on chondrocytes. The Journal of biological chemistry 1998;273(32): 20383-20389.
  80. 80. Hering T M. Regulation of chondrocyte gene expression. Frontiers in bioscience : a journal and virtual library 1999;4: D743-761.
  81. 81. Hirsch M S, Lunsford L E, Trinkaus-Randall V, Svoboda K K. Chondrocyte survival and differentiation in situ are integrin mediated. Developmental dynamics : an official publication of the American Association of Anatomists 1997;210(3): 249-263.
  82. 82. Durr J, Goodman S, Potocnik A, von der Mark H, von der Mark K. Localization of beta 1-integrins in human cartilage and their role in chondrocyte adhesion to collagen and fibronectin. Experimental cell research 1993;207(2): 235-244.
  83. 83. Enomoto M, Leboy P S, Menko A S, Boettiger D. Beta 1 integrins mediate chondrocyte interaction with type I collagen, type II collagen, and fibronectin. Experimental cell research 1993;205(2): 276-285.
  84. 84. Camper L, Heinegard D, Lundgren-Akerlund E. Integrin alpha2beta1 is a receptor for the cartilage matrix protein chondroadherin. The Journal of cell biology 1997;138(5): 1159-1167.
  85. 85. Enomoto-Iwamoto M, Iwamoto M, Pacifici M, Boettiger D, Kurisu K, Suzuki F. The interaction of alpha 5 beta 1 integrin with fibronectin is required for proliferation of chondrocytes. Trans Orthopaedic Research Society 1995;20: 396.
  86. 86. Loeser R F. Integrin-mediated attachment of articular chondrocytes to extracellular matrix proteins. Arthritis and rheumatism 1993;36(8): 1103-1110.
  87. 87. Loeser R F, Sadiev S, Tan L, Goldring M B. Integrin expression by primary and immortalized human chondrocytes: evidence of a differential role for alpha1beta1 and alpha2beta1 integrins in mediating chondrocyte adhesion to types II and VI collagen. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2000;8(2): 96-105.
  88. 88. Enomoto-Iwamoto M, Iwamoto M, Nakashima K, Mukudai Y, Boettiger D, Pacifici M, Kurisu K, Suzuki F. Involvement of alpha5beta1 integrin in matrix interactions and proliferation of chondrocytes. Journal of bone and mineral research : the official journal of the American Society for Bone and Mineral Research 1997;12(7): 1124-1132.
  89. 89. Tavella S, Bellese G, Castagnola P, Martin I, Piccini D, Doliana R, Colombatti A, Cancedda R, Tacchetti C. Regulated expression of fibronectin, laminin and related integrin receptors during the early chondrocyte differentiation. Journal of cell science 1997;110 (Pt 18): 2261-2270.
  90. 90. Shakibaei M. Inhibition of chondrogenesis by integrin antibody in vitro. Experimental cell research 1998;240(1): 95-106.
  91. 91. Beekman B, Verzijl N, Bank R A, von der Mark K, TeKoppele J M. Synthesis of collagen by bovine chondrocytes cultured in alginate; posttranslational modifications and cell-matrix interaction. Experimental cell research 1997;237(1): 135-141.
  92. 92. Homandberg G A, Hui F. Arg-Gly-Asp-Ser peptide analogs suppress cartilage chondrolytic activities of integrin-binding and nonbinding fibronectin fragments. Archives of biochemistry and biophysics 1994;310(1): 40-48.
  93. 93. Clancy R M, Rediske J, Tang X, Nijher N, Frenkel S, Philips M, Abramson S B. Outside-in signaling in the chondrocyte. Nitric oxide disrupts fibronectin-induced assembly of a subplasmalemmal actin/rho A/focal adhesion kinase signaling complex. The Journal of clinical investigation 1997;100(7): 1789-1796.
  94. 94. Ingber D. Integrins as mechanochemical transducers. Current opinion in cell biology 1991;3(5): 841-848.
  95. 95. Davies P F. Flow-mediated endothelial mechanotransduction. Physiological reviews 1995;75(3): 519-560.
  96. 96. Shyy J Y, Chien S. Role of integrins in cellular responses to mechanical stress and adhesion. Current opinion in cell biology 1997;9(5): 707-713.
  97. 97. Wang N, Butler J P, Ingber D E. Mechanotransduction across the cell surface and through the cytoskeleton. Science 1993;260(5111): 1124-1127.
  98. 98. Maniotis A J, Chen C S, Ingber D E. Demonstration of mechanical connections between integrins, cytoskeletal filaments, and nucleoplasm that stabilize nuclear structure. Proceedings of the National Academy of Sciences of the United States of America 1997;94(3): 849-854.
  99. 99. Pommerenke H, Schreiber E, Durr F, Nebe B, Hahnel C, Moller W, Rychly J. Stimulation of integrin receptors using a magnetic drag force device induces an intracellular free calcium response. European journal of cell biology 1996;70(2): 157-164.
  100. 100. Bierbaum S, Notbohm H. Tyrosine phosphorylation of 40 kDa proteins in osteoblastic cells after mechanical stimulation of beta1-integrins. European journal of cell biology 1998;77(1): 60-67.
  101. 101. Schmidt C, Pommerenke H, Durr F, Nebe B, Rychly J. Mechanical stressing of integrin receptors induces enhanced tyrosine phosphorylation of cytoskeletally anchored proteins. The Journal of biological chemistry 1998;273(9): 5081-5085.
  102. 102. Holmvall K, Camper L, Johansson S, Kimura J H, Lundgren-Akerlund E. Chondrocyte and chondrosarcoma cell integrins with affinity for collagen type II and their response to mechanical stress. Experimental cell research 1995;221(2): 496-503.
  103. 103. Martinac B. Mechanosensitive ion channels: molecules of mechanotransduction. Journal of cell science 2004;117(Pt 12): 2449-2460.
  104. 104. Ingber D E. Cellular mechanotransduction: putting all the pieces together again. FASEB journal : official publication of the Federation of American Societies for Experimental Biology 2006;20(7): 811-827.
  105. 105. Muller U. Cadherins and mechanotransduction by hair cells. Current opinion in cell biology 2008;20(5): 557-566.
  106. 106. Civitelli R. Cell-cell communication in the osteoblast/osteocyte lineage. Archives of biochemistry and biophysics 2008;473(2): 188-192.
  107. 107. Yellowley C E, Li Z, Zhou Z, Jacobs C R, Donahue H J. Functional gap junctions between osteocytic and osteoblastic cells. Journal of bone and mineral research : the official journal of the American Society for Bone and Mineral Research 2000;15(2): 209-217.
  108. 108. Zhang J, Zhang H Y, Zhang M, Qiu Z Y, Wu Y P, Callaway D A, Jiang J X, Lu L, Jing L, Yang T, Wang M Q. Connexin43 hemichannels mediate small molecule exchange between chondrocytes and matrix in biomechanically-stimulated temporomandibular joint cartilage. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2014;22(6): 822-830.
  109. 109. Waggett A D, Benjamin M, Ralphs J R. Connexin 32 and 43 gap junctions differentially modulate tenocyte response to cyclic mechanical load. European journal of cell biology 2006;85(11): 1145-1154.
  110. 110. Ruhlen R, Marberry K. The chondrocyte primary cilium. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2014;22(8): 1071-1076.
  111. 111. Whitfield J F. The solitary (primary) cilium--a mechanosensory toggle switch in bone and cartilage cells. Cellular signalling 2008;20(6): 1019-1024.
  112. 112. Poole C A, Jensen C G, Snyder J A, Gray C G, Hermanutz V L, Wheatley D N. Confocal analysis of primary cilia structure and colocalization with the Golgi apparatus in chondrocytes and aortic smooth muscle cells. Cell biology international 1997;21(8): 483-494.
  113. 113. Malone A M, Anderson C T, Tummala P, Kwon R Y, Johnston T R, Stearns T, Jacobs C R. Primary cilia mediate mechanosensing in bone cells by a calcium-independent mechanism. Proceedings of the National Academy of Sciences of the United States of America 2007;104(33): 13325-13330.
  114. 114. Ingber D E. Tensegrity: the architectural basis of cellular mechanotransduction. Annual review of physiology 1997;59: 575-599.
  115. 115. Chicurel M E, Chen C S, Ingber D E. Cellular control lies in the balance of forces. Current opinion in cell biology 1998;10(2): 232-239.
  116. 116. Chen C S, Ingber D E. Tensegrity and mechanoregulation: from skeleton to cytoskeleton. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 1999;7(1): 81-94.
  117. 117. Wright M O, Stockwell R A, Nuki G. Response of plasma membrane to applied hydrostatic pressure in chondrocytes and fibroblasts. Connective tissue research 1992;28(1-2): 49-70.
  118. 118. Tsunoo A, Yoshii M, Narahashi T. Block of calcium channels by enkephalin and somatostatin in neuroblastoma-glioma hybrid NG108-15 cells. Proceedings of the National Academy of Sciences of the United States of America 1986;83(24): 9832-9836.
  119. 119. Narahashi T, Tsunoo A, Yoshii M. Characterization of two types of calcium channels in mouse neuroblastoma cells. The Journal of physiology 1987;383: 231-249.
  120. 120. Bers D M. A simple method for the accurate determination of free [Ca] in Ca-EGTA solutions. The American journal of physiology 1982;242(5): C404-408.
  121. 121. Thastrup O. Role of Ca2(+)-ATPases in regulation of cellular Ca2+ signalling, as studied with the selective microsomal Ca2(+)-ATPase inhibitor, thapsigargin. Agents and actions 1990;29(1-2): 8-15.
  122. 122. Thastrup O, Cullen P J, Drobak B K, Hanley M R, Dawson A P. Thapsigargin, a tumor promoter, discharges intracellular Ca2+ stores by specific inhibition of the endoplasmic reticulum Ca2(+)-ATPase. Proceedings of the National Academy of Sciences of the United States of America 1990;87(7): 2466-2470.
  123. 123. Lytton J, Westlin M, Hanley M R. Thapsigargin inhibits the sarcoplasmic or endoplasmic reticulum Ca-ATPase family of calcium pumps. The Journal of biological chemistry 1991;266(26): 17067-17071.
  124. 124. O'Conor C J, Leddy H A, Benefield H C, Liedtke W B, Guilak F. TRPV4-mediated mechanotransduction regulates the metabolic response of chondrocytes to dynamic loading. Proceedings of the National Academy of Sciences of the United States of America 2014;111(4): 1316-1321.
  125. 125. Dufour C, Holy X, Marie P J. Transforming growth factor-beta prevents osteoblast apoptosis induced by skeletal unloading via PI3K/Akt, Bcl-2, and phospho-Bad signaling. American journal of physiology. Endocrinology and metabolism 2008;294(4): E794-801.
  126. 126. Guan Y, Yang X, Yang W, Charbonneau C, Chen Q. Mechanical activation of mammalian target of rapamycin pathway is required for cartilage development. FASEB journal : official publication of the Federation of American Societies for Experimental Biology 2014.
  127. 127. Fitzgerald J B, Jin M, Chai D H, Siparsky P, Fanning P, Grodzinsky A J. Shear- and compression-induced chondrocyte transcription requires MAPK activation in cartilage explants. The Journal of biological chemistry 2008;283(11): 6735-6743.
  128. 128. Vincent T, Saklatvala J. Basic fibroblast growth factor: an extracellular mechanotransducer in articular cartilage? Biochemical Society transactions 2006;34(Pt 3): 456-457.
  129. 129. Lee H S, Millward-Sadler S J, Wright M O, Nuki G, Salter D M. Integrin and mechanosensitive ion channel-dependent tyrosine phosphorylation of focal adhesion proteins and beta-catenin in human articular chondrocytes after mechanical stimulation. Journal of bone and mineral research : the official journal of the American Society for Bone and Mineral Research 2000;15(8): 1501-1509.
  130. 130. Chen N X, Geist D J, Genetos D C, Pavalko F M, Duncan R L. Fluid shear-induced NFkappaB translocation in osteoblasts is mediated by intracellular calcium release. Bone 2003;33(3): 399-410.
  131. 131. Anghelina M, Sjostrom D, Perera P, Nam J, Knobloch T, Agarwal S. Regulation of biomechanical signals by NF-kappaB transcription factors in chondrocytes. Biorheology 2008;45(3-4): 245-256.
  132. 132. Agarwal S, Deschner J, Long P, Verma A, Hofman C, Evans C H, Piesco N. Role of NF-kappaB transcription factors in antiinflammatory and proinflammatory actions of mechanical signals. Arthritis and rheumatism 2004;50(11): 3541-3548.
  133. 133. Chowdhury T T, Bader D L, Lee D A. Dynamic compression counteracts IL-1 beta-induced release of nitric oxide and PGE2 by superficial zone chondrocytes cultured in agarose constructs. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2003;11(9): 688-696.
  134. 134. Jeffrey J E, Aspden R M. Cyclooxygenase inhibition lowers prostaglandin E2 release from articular cartilage and reduces apoptosis but not proteoglycan degradation following an impact load in vitro. Arthritis research & therapy 2007;9(6): R129.
  135. 135. Ke H Z, Qi H, Weidema A F, Zhang Q, Panupinthu N, Crawford D T, Grasser W A, Paralkar V M, Li M, Audoly L P, Gabel C A, Jee W S, Dixon S J, Sims S M, Thompson D D. Deletion of the P2X7 nucleotide receptor reveals its regulatory roles in bone formation and resorption. Molecular endocrinology 2003;17(7): 1356-1367.
  136. 136. Salter D M, Millward-Sadler S J, Nuki G, Wright M O. Integrin-interleukin-4 mechanotransduction pathways in human chondrocytes. Clinical orthopaedics and related research 2001;(391 Suppl): S49-60.
  137. 137. Millward-Sadler S J, Wright M O, Lee H, Nishida K, Caldwell H, Nuki G, Salter D M. Integrin-regulated secretion of interleukin 4: A novel pathway of mechanotransduction in human articular chondrocytes. The Journal of cell biology 1999;145(1): 183-189.
  138. 138. Salter D M, Wallace W H, Robb J E, Caldwell H, Wright M O. Human bone cell hyperpolarization response to cyclical mechanical strain is mediated by an interleukin-1beta autocrine/paracrine loop. Journal of bone and mineral research : the official journal of the American Society for Bone and Mineral Research 2000;15(9): 1746-1755.
  139. 139. Fujiwara Y, Uesugi M, Saito T. Down-regulation of basic fibroblast growth factor production from cartilage by excessive mechanical stress. Journal of orthopaedic science : official journal of the Japanese Orthopaedic Association 2005;10(6): 608-613.
  140. 140. Woo S, Kwan M, Coutts R, Akeson W. Biomechanical considerations. In: Moskowitz R, Howell D, Goldberg V, Mankin H (eds) Osteoarthritis: Diagnosis and medical/surgical management. Philadelphia: WB Saunders; 1992. 191-211.
  141. 141. Hoch D H, Grodzinsky A J, Koob T J, Albert M L, Eyre D R. Early changes in material properties of rabbit articular cartilage after meniscectomy. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1983;1(1): 4-12.
  142. 142. Burton W N, Todhunter R J, Lust G. Animal models of osteoarthritis. In: JF W, DS H (eds) Joint cartilage degradation: basic and clinical aspects. Marcel Dekker: New York; 1993. 347-384.
  143. 143. Setton L, Zimmerman J, Mow V, Muller F, Pita J, Howell D. Effects of disuse on the tensile properties and composition of canine knee joint cartilage. Trans Orthopaedic Research Society 1990;15: 155.
  144. 144. Palmoski M J, Brandt K D. Running inhibits the reversal of atrophic changes in canine knee cartilage after removal of a leg cast. Arthritis and rheumatism 1981;24(11): 1329-1337.
  145. 145. Setton L A, Elliott D M, Mow V C. Altered mechanics of cartilage with osteoarthritis: human osteoarthritis and an experimental model of joint degeneration. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 1999;7(1): 2-14.
  146. 146. Wright M O, Nishida K, Bavington C, Godolphin J L, Dunne E, Walmsley S, Jobanputra P, Nuki G, Salter D M. Hyperpolarisation of cultured human chondrocytes following cyclical pressure-induced strain: evidence of a role for alpha 5 beta 1 integrin as a chondrocyte mechanoreceptor. Journal of orthopaedic research : official publication of the Orthopaedic Research Society 1997;15(5): 742-747.
  147. 147. Orazizadeh M, Lee H S, Groenendijk B, Sadler S J, Wright M O, Lindberg F P, Salter D M. CD47 associates with alpha 5 integrin and regulates responses of human articular chondrocytes to mechanical stimulation in an in vitro model. Arthritis research & therapy 2008;10(1): R4.
  148. 148. Millward-Sadler S J, Khan N S, Bracher M G, Wright M O, Salter D M. Roles for the interleukin-4 receptor and associated JAK/STAT proteins in human articular chondrocyte mechanotransduction. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2006;14(10): 991-1001.
  149. 149. Lee H S, Millward-Sadler S J, Wright M O, Nuki G, Al-Jamal R, Salter D M. Activation of Integrin-RACK1/PKCalpha signalling in human articular chondrocyte mechanotransduction. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2002;10(11): 890-897.
  150. 150. Millward-Sadler S J, Mackenzie A, Wright M O, Lee H S, Elliot K, Gerrard L, Fiskerstrand C E, Salter D M, Quinn J P. Tachykinin expression in cartilage and function in human articular chondrocyte mechanotransduction. Arthritis and rheumatism 2003;48(1): 146-156.
  151. 151. Fitzgerald J B, Jin M, Dean D, Wood D J, Zheng M H, Grodzinsky A J. Mechanical compression of cartilage explants induces multiple time-dependent gene expression patterns and involves intracellular calcium and cyclic AMP. The Journal of biological chemistry 2004;279(19): 19502-19511.
  152. 152. Murata M, Bonassar L J, Wright M, Mankin H J, Towle C A. A role for the interleukin-1 receptor in the pathway linking static mechanical compression to decreased proteoglycan synthesis in surface articular cartilage. Archives of biochemistry and biophysics 2003;413(2): 229-235.
  153. 153. Fanning P J, Emkey G, Smith R J, Grodzinsky A J, Szasz N, Trippel S B. Mechanical regulation of mitogen-activated protein kinase signaling in articular cartilage. The Journal of biological chemistry 2003;278(51): 50940-50948.
  154. 154. Li K W, Wang A S, Sah R L. Microenvironment regulation of extracellular signal-regulated kinase activity in chondrocytes: effects of culture configuration, interleukin-1, and compressive stress. Arthritis and rheumatism 2003;48(3): 689-699.
  155. 155. Salter D M, Millward-Sadler S J, Nuki G, Wright M O. Differential responses of chondrocytes from normal and osteoarthritic human articular cartilage to mechanical stimulation. Biorheology 2002;39(1-2): 97-108.
  156. 156. Millward-Sadler S J, Wright M O, Lee H S, Caldwell H, Nuki G, Salter D M. Altered electrophysiological responses to mechanical stimulation and abnormal signalling through alpha5beta1 integrin in chondrocytes from osteoarthritic cartilage. Osteoarthritis and cartilage / OARS, Osteoarthritis Research Society 2000;8(4): 272-278.
  157. 157. Attur M G, Dave M N, Clancy R M, Patel I R, Abramson S B, Amin A R. Functional genomic analysis in arthritis-affected cartilage: yin-yang regulation of inflammatory mediators by alpha 5 beta 1 and alpha V beta 3 integrins. Journal of immunology 2000;164(5): 2684-2691.
  158. 158. Boisclair Y R, Wang J, Shi J, Hurst K R, Ooi G T. Role of the suppressor of cytokine signaling-3 in mediating the inhibitory effects of interleukin-1beta on the growth hormone-dependent transcription of the acid-labile subunit gene in liver cells. The Journal of biological chemistry 2000;275(6): 3841-3847.

Written By

Herng-Sheng Lee and Donald M. Salter

Submitted: 29 August 2014 Published: 01 July 2015