2 - Etiopathogenesis of Osteoarthritis

Authors: Moskowitz, Roland W.; Altman, Roy D.; Hochberg, Marc C.; Buckwalter, Joseph A.; GoldberG, Victor M.

Title: Osteoarthritis: Diagnosis and Medical/Surgical Management, 4th Edition

Copyright 2007 Lippincott Williams & Wilkins

> Table of Contents > I - Basic Considerations > 2 - Etiopathogenesis of Osteoarthritis

function show_scrollbar() {}

2

Etiopathogenesis of Osteoarthritis

A. Robin Poole

Farshid Guilak

Steven B. Abramson

Osteoarthritis (OA) is a condition that represents a pathological imbalance of degradative and reparative processes involving the whole joint and its component parts, with secondary inflammatory changes, particularly in the synovium, but also in the articular cartilage itself (Fig. 2-1). Idiopathic primary OA may involve one particular joint, or it may be generalized or involve multiple joints in erosive inflammatory forms1,2 (Table 2-1). The presentation of this pathological condition in joints may be a consequence of the biomechanics within the joint which reveal otherwise masked systemic genetically determined changes. The mechanical pressures within the joint may therefore reveal weaknesses in tissue maintenance that are more widespread than previously considered.

Most forms of OA fall into two categories, depending on the predominant background: those that are primary, and often idiopathic, with abnormalities of joint biomaterial and biomechanically faulty joint structure that may result from a recognizable mutation, and those that are secondary and result from superimposed risk factors affecting distribution and severity of loading forces acting on specific joints, such as joint injury.

Risk Factors for Osteoarthritis

Risk Factors Associated with Abnormalities in Joint Biomaterial and Biomechanical Properties, Chemical Injury, or Endocrine Disorders

Familial OA (Table 2-1) may result from abnormal cartilage structure and properties such as a consequence of a mutation in the type II collagen COL2A1 gene (localized on chromosome 12), which causes not only cartilage dysplasia but also a severe form of OA with defective collagen.3,4 Onset is usually soon after cessation of growth. (See Chapter 6 for other gene defects relevant to OA associated with dysplasias and joint laxity.) Chronic abnormalities of growth plate development and bone growth leading to altered congruity of articulating surfaces can also cause OA. OA can develop in patients after traumatic injury or damage to chondrocytes associated with abnormal deposits in the cartilage matrix found in metabolic diseases such as hemochromatosis, ochronosis or alkaptonuria, Wilson disease, and Gaucher disease. OA can also result from disturbances in cartilage metabolism caused by endocrine disorders, such as acromegaly5 (see Chapter 13).

Risk Factors Related to Abnormalities in the Quality and Distribution of Loading Forces on a Joint

Cartilage damage due to trauma, impact injuries, abnormal joint loading, and excessive wear, or as part of an aging process, can lead to changes in the composition, structure, and material properties of the tissue (Fig. 2-2). These alterations can compromise the ability of cartilage to function and survive in the strenuous mechanical environment normally found in load-bearing joints. Joint injury and subsequent joint instability, from loss of ligamental or meniscal support, are significant risk factors for OA.6 Injury and abnormal loading, due to overexercise or abnormal joint use, are a risk for OA.7,8 In regard to underloading, disuse immobilization from chronic inanition, neurologic disorders, or postoperative casting can lead to cartilage disuse atrophy, with increased vulnerability to cartilage injury, unless exercise programs for rehabilitation are carefully controlled. Evidence for this is available from studies of casted animals and those subsequently subjected to controlled or excessive treadmill exercise. Immobilized joint cartilage exhibits arrest of cartilage proteoglycan aggrecan

P.28


synthesis,8,9 altered biomechanical properties,10,11,12 and associated elevated metalloproteinases that may contribute to cartilage damage13(reviewed in Helminen et al.14). Deep ulcerative lesions with elevated protease activity can result from treadmill overexercise after disuse immobilization.15 Standard levels of exercise may cause site-specific changes in proteoglycan content and cartilage stiffness, although these changes are not believed to be deleterious16 and may potentially have a beneficial effect in the normal joint.17,18,19 Altered joint loading due to instability or injury of the joint is now well known to be a significant risk factor for the onset and progression of OA.8,20,21 Alterations in joint loading, brought about through ligament transection22 or meniscectomy,23 may lead to profound and repeatable changes in joint tissues which mimic changes seen in early human OA, including increased hydration and proteoglycan turnover, and decreased tissue stiffness in tension, compression, and shear.11,24,25,26,27,28 Obesity is a well-defined risk factor involving excessive joint loading29 as well as systemic metabolic changes that include low-grade chronic inflammation.30 In this respect, obesity is regarded as the number one preventable risk factor for OA.19,31 Evolutionary adaptations to the upright posture by humans have redistributed preponderant loading forces to new sites, predisposing to an increased risk for development of OA in hips, knees, bunion joints, and the lumbar spine.32

Figure 2-1 Gross pathologic changes observed in OA joints during many years of degenerative change.

Age-related changes in the magnitude and pattern of stresses on joint cartilage may arise from a number of factors, including altered gait, muscle weakness, changes in proprioception, and changes in body weight. A neuromus-cular control deficit is likely to contribute to the loss of normal attenuation of body weight-bearing forces during walking. This is manifested in OA patients as the digging of the heels on forceplate analysis.7 The development of experimental canine OA can be accelerated by posterior nerve root section at the spinal cord, which affects the afferent signals governing stereognostic control of the affected arthritic limb;33 this provides important insights into the pathophysiologic process of the Charcot joint. The mechanism of impact loading, in which energy is normally predominantly absorbed or attenuated by strong muscle groups in the thigh and leg, is also significantly impaired in OA patients,34 who exhibit decreased muscle strength. Quadriceps weakness is associated with knee pain, especially in OA.35 Biomechanical changes in the joint capsule as a consequence of genetic or post-translational changes, such as those affecting collagen fibers, may also influence joint loading.

Examination of nutritional factors in the etiopathology of OA has provided evidence for an increased risk of development of OA of the knee with vitamin E and C deficiencies,36 no doubt influencing matrix assembly and function. In Asia, Kashin-Beck disease, a form of OA, may be caused by dietary exposure to an endemic fungus.37 Hypothyroidism afflicts many of these patients because of a dietary selenium deficiency.38

Articular Cartilage

Composition

Normal articular cartilage consists of an extensive, hydrated extracellular matrix that is synthesized and maintained by a sparse population of specialized cells, the chondrocytes (Fig. 2-3). In the adult human, these cells may occupy as little as 2% of the total volume. The matrix consists mainly of collagen (mostly type II with lesser amounts of other collagen types) and proteoglycans, principally aggrecan, which is large and aggregates with hyaluronic acid (HA).39 Types II, IX, and XI collagens of cartilage combine to form a fibrillar network that provides a structural framework of the matrix in the form of an inhomogeneous and anisotropic meshwork of fibers, which is surrounded by a highly concentrated solution of the proteoglycan aggrecan. A secondary microfibrillar

P.29


network involving type VI collagen is concentrated in the pericellular matrix between the cell surface and the territorial matrix.

TABLE 2-1 CLASSIFICATION OF OSTEOARTHRITIS

PRIMARY (IDIOPATHIC)
Peripheral joints (single vs. multiple joints)
Interphalangeal joints (nodal) (e.g., distal interphalangeal, proximal interphalangeal)
Other small joints (e.g., first carpometacarpal, first metacarpophalangeal)
Large joints (e.g., hip, knee)
SPINE
Apophyseal joints
Intervertebral joints
VARIANT SUBSETS
Erosive inflammatory osteoarthritis
Generalized osteoarthritis
Chondromalacia patellae
Diffuse idiopathic skeletal hyperostosis (DISH, ankylosing hyperostosis)
SECONDARY
Trauma
Acute
Chronic (occupational, sports, obesity)
OTHER JOINT DISORDERS
Local (fracture, avascular necrosis, infection)
Diffuse (rheumatoid arthritis, hypermobility syndrome, hemorrhagic diatheses)
SYSTEMIC METABOLIC DISEASE
Ochronosis (alkaptonuria)
Hemochromatosis
Wilson disease
Kashin-Beck disease
ENDOCRINE DISORDERS
Acromegaly
Hyperparathyroidism
Diabetes mellitus
CALCIUM CRYSTAL DEPOSITION DISEASES
Calcium pyrophosphate dihydrate
Calcium apatite
NEUROPATHIC DISORDERS (CHARCOT JOINTS)
(e.g., tabes dorsalis, diabetes mellitus, intra-articular steroid overuse)
FAMILIAL OSTEOARTHRITIS (associated with skeletal dysplasias such as multiple epiphyseal dysplasia, spondyloepiphyseal dysplasia)
MISCELLANEOUS
Frostbite
Long-leg arthropathy

Proteoglycans are complex macromolecules with a protein core to which are attached glycosaminoglycan chains, primarily of chondroitin sulfate and keratan sulfate. These glycosaminoglycans, which are negatively charged in solution, are responsible for the hydration and large swelling pressure of this tissue.40,41 Many other smaller proteoglycans and other molecules are present within the tissue, some of which may be directly associated with the collagen fibrils. For more details, please see Chapter 4.

Biomechanics of Articular Cartilage

Articular cartilage serves as a low-friction, wear-resistant surface for load support, load transfer, and motion between the bones of the diarthrodial joint. Under normal circumstances, this tissue, bathed in synovial fluid, is able to withstand millions of cycles of loading each year at stresses that may reach 18 MPa,42 while exhibiting little or no wear. These unique properties are endowed by the composition and structure of the major constituents of articular cartilage, and their associated spatial variations provide a complex and inhomogeneous set of material properties.

From a biomechanical standpoint, articular cartilage can be viewed as a fiber-reinforced, porous, and permeable composite material that is saturated with an interstitial fluid, water.43 These characteristics provide for mechanical properties that are viscoelastic (time- or rate-dependent), anisotropic (dependent on direction), and nonlinear (dependent on magnitude of strain). Accordingly, the biomechanical properties of cartilage have been widely studied using models that take into account the multiple phases (collagen solid matrix, interstitial fluid, and mobile and fixed ionic groups) and the interactions among these phases.43,44,45

The largest constituent, water, provides load support through fluid pressurization and energy dissipation by fluid flow in response to applied loading.43 The second largest constituent, the collagen fibril, provides articular cartilage with its nonlinear properties in tension,46,47 which are inhomogeneously distributed with depth from the tissue surface. However, these tensile properties decrease progressively with increasing age after 30 years in articular cartilages, such as of the hip, that most commonly develop OA.48 Age-related changes in cartilage tensile properties are believed to arise in part from the accumulation of advanced glycation end products (AGE) which increase the brittleness of the cartilage, increasing the potential for tissue fracture and aging-associated biomechanical dysfunction.49

The high concentration of negative charges associated with the proteoglycans in cartilage has a significant effect on the mechanical behavior of the tissue. The negatively charged nature of cartilage arises primarily from the glycosaminoglycans, namely, the many chondroitin sulfate and keratan sulfate chains that are present on aggrecan molecules.50 The large size of aggrecan and its interactions with HA to form macromolecular aggregates serve to retain aggrecan molecules in the extracellular matrix so that their negative charges are fixed. These fixed charges are responsible for the physicochemical and electrokinetic properties of this tissue, such as a large osmotic pressure and associated propensity to swell and exhibit streaming potentials, streaming currents, and electro-osmotic effects.41,45 As a result, the mechanical behavior of articular cartilage exhibits an interdependence on physicochemical factors, such as the structural organization and properties

P.30


of the collagen fibrillar network, the proteoglycan aggrecan concentration, and the type or distribution of counterions.44 These physicochemical properties significantly contribute to the function of articular cartilage.45 The swelling pressure is believed to largely contribute to residual stresses in the cartilage matrix that enhance the support and distribution of applied loads.51,52 The loss of tensile properties and swelling pressure as collagen and proteoglycans are degraded and lost in OA results in mechanical decompensation and an ensuing increase in the magnitude of tissue strains under similar magnitudes of physiologic loading,10,28,53 particularly at the cellular level.54

Figure 2-2 Potential mechanisms involved in the etiopathogenesis of OA.

Figure 2-3 Changes observed in articular cartilage in OA involving chondrocytes and extracellular matrix.

Articular cartilage endows the synovial joint with frictional properties that remain unmatched by man-made joints. The coefficient of friction (ratio of frictional force to compressive force) of cartilage gliding on cartilage has been reported to be as low as 0.002.55 The mechanisms by which the synovial joint achieves these properties involve a combination of biomechanical and biomolecular factors.56

P.31


For example, joint congruity and cartilage deformation serve to distribute loads over a larger contact area, thereby minimizing contact stresses. The high water content and low hydraulic permeability of the cartilage extracellular matrix provide a unique mechanism of supporting loads whereby fluid pressurization within the tissue can bear nearly 90% of the load and thus minimize stresses on the solid extracellular matrix.57,58 Synovial fluid plays multiple roles in joint lubrication by providing a high viscosity squeeze film layer that delays cartilage-to-cartilage contact under compression59 but also serves as a source of boundary lubricant molecules within the joint.

Boundary lubricants are molecules that may be adsorbed or bound to the cartilage surface and decrease friction in cases when the tissue surfaces actually contact one another. In the past two decades, there have been significant advances in our understanding of the composition, structure, and properties of biomolecular components of the synovial fluid that serve as the primary boundary lubricants for the joint. Lubricin, a ~ 227 kDa glycoprotein, was first identified from bovine synovial fluid by density gradient sedimentation and gel-permeation chromatography.60 This molecule was first identified as a product of synovial fibroblasts expressing the gene for megakaryocyte stimulating factor (MSF, also termed PRG4). Subsequent investigations have identified a 345 kDa protein, termed superficial zone protein (SZP), as an additional product of the MSF gene.61,62 The importance of these molecules in the health of the joint have been demonstrated by the rare autosomal recessive disorder, camptodactyly-arthropathy-coxa vara-pericarditis (CACP) syndrome, which has been attributed to mutations in the PRG4 gene.63 The loss of SZP function in CACP is associated with a severe juvenile-onset joint failure that is noninflammatory in nature.

Biomechanical Regulation of Cartilage Metabolism

A number of in vivo studies have emphasized the relationship between mechanical loading and the health of the joint, and suggest that a critical level and pattern of mechanical stress is required to maintain the normal balance of cartilage synthesis and breakdown. Under normal physiological conditions, the components of the extracellular matrix are in a state of slow turnover, retained in a homeostatic balance between the catabolic and anabolic events of the chondrocytes. These activities are controlled through the processing of both genetic and environmental information, which includes the action of soluble mediators (e.g., growth factors and cytokines), extracellular matrix composition, and physical factors such as mechanical stress. Indeed, many studies have shown that the mechanical stress environment of the joint is an important factor that influences (and presumably regulates) the activity of the chondrocytes in vivo.14,64

Considerable research effort has been directed toward understanding the processes by which physical signals are converted to a biochemical signal by the chondrocyte population. Clarification of the specific signaling mechanisms in normal and inflamed cartilage would not only provide a better understanding of the processes which regulate the physiology of cartilage, but would also be expected to yield new insights on the pathogenesis of OA. In this respect, in vitro explant models of mechanical loading provide a model system in which the biomechanical and biochemical environments can be better controlled as compared to the in vivo situation. Explant models of cartilage loading have been utilized in a number of different loading configurations, including unconfined compression, indentation, tension, shear, and osmotic and hydrostatic pressure (reviewed in Guilak et al.64). The general consensus of these studies is that static compression suppresses matrix biosynthesis, and cyclic and intermittent loading stimulate chondrocyte metabolism (e.g., Gray et al.,65 Sah et al.,66 Guilak et al.,67 Torzilli et al.68). These responses have been reported over a wide range of loading magnitudes, and exhibit a stress-dose dependency. Excessive loading (e.g., high magnitude, long duration) seems to have a deleterious effect, resulting in cell death, tissue disruption, and swelling.69,70,71

The cellular transduction mechanisms responsible for converting mechanical stress into a biochemical response are the subject of much study but are not fully understood. Currently, there is significant evidence that chondrocytes may transduce mechanical signals into biochemical responses through various intracellular and intercellular signaling pathways, including activation of the traditional second messenger pathways such as cyclic AMP (cAMP), inositol trisphosphate, or calcium ion (see Guilak et al.64 and Stockwell72 for reviews).

The Degeneration of Articular Cartilage in Osteoarthritis

The classic loss of articular cartilage observed in OA may be initiated as a focal process that first appears to manifest at the articular cartilage surface. These lesions, which are commonly observed in aging human populations, have all the biochemical features of OA cartilages at arthroplasty.73 They may progressively enlarge to involve specific compartments, inducing alterations in other articulating surfaces by producing changes in loading. Some of the clues as to whether these focal lesions enlarge may have been revealed by recent studies on these lesions in the ankle versus the knee joint. In knee lesions, the emphasis in the cartilage is on matrix degradation, whereas in the ankle it is on matrix synthesis.39 Moreover, in the ankle the response involving increased matrix turnover is more generalized, whereas in the knee it is associated with the lesion.39,73 These fundamental differences in the response and in the lesions could help explain why knee OA is much more common. More widespread changes in loading after traumatic injury may involve the whole joint cartilage, in which alterations in cartilage matrix turnover are detectable within days, weeks, or months after damage to the anterior cruciate ligament or meniscus.74,75 In particular, joint injuries that involve fracture of the articular surface tend to accelerate more rapidly to a degenerative state of post-traumatic arthritis.76

P.32


Cartilage degeneration is first observed at the articular surface in the form of fibrillation (Fig. 2-3). Splits are initially parallel to the articular surface; later, they vertically penetrate the damaged cartilage, eventually reaching subchondral bone. Cell cloning is observed early around the splits but is confined to more superficial sites. Many of these cells have become hypertrophic. Progressive loss of cartilage thickness, starting at the surface, is observed.

In large joints, idiopathic OA is ordinarily believed to be a slow process that may take as long as 20 to 30 years, but it is accelerated in cases of joint injury75,77,78,79 or may present clinically on cessation of growth as in familial OA.80 Degeneration is often more pronounced in the tibial cartilage, particularly in the medial compartment. This characteristic is observed in both human and in experimental OA.

Other examples of long-term development of OA come from experimental studies of anterior cruciate ligament transection in dogs, in which bone, ligament, and synovial changes are also observed early28,81,82 (see Chapter 5), presenting as an advanced lesion after a period of only 3 or more years. The earliest changes in the unstable knee that can be seen in articular cartilage in injured joints in animals and humans feature cartilage swelling (also called a hypertrophic reaction), with enhanced synthesis of matrix with an increased content of aggrecan.24 This is followed by a phase of increased matrix turnover, with net depletion of principal matrix components. Finally, severe damage to and loss of the collagen network is observed.11,28,75,83,84 The hypertrophic stage, not to be confused with chondrocyte hypertrophy, clearly precedes the occurrence of the lesional stage with its characteristic deep focal loss of cartilage. In the dog, the biomechanical properties correlate well with a reduction in proteoglycan aggregation and the loss of hyaluronan and link protein.84,85 The presence of smaller proteoglycan aggregates in uncovered or unprotected areas of tibial plateau cartilage in normal control subjects, coupled with the presence of elevated protease activity after surgery, may result in part from a shift of load-bearing sites during the development of OA in this model.86,87

Proteinases of Osteoarthritis Cartilage

Metalloproteinases (MMPs) are generally considered to play a principal role in the cleavage of matrix macromolecules, including type II collagen and the cartilage proteoglycan aggrecan. Collagenases cleave type II collagen, which contains a long triple-helical domain, at a specific site approximately three quarters from the amino terminus. This results in unwinding (denaturation) of the -chains, which are then susceptible to secondary cleavage by collagenases and other MMPs such as stromelysin 1(MMP-3) and the gelatinases MMP-2 and MMP-9. Aggrecan is cleaved by different MMPs and by the aggrecanases (adamolysins) ADAMTS-4 and -5.

Four collagenases may be active in articular cartilage, namely, MMP-1, -8, -13, and -14 (Table 2-2). MMP-13 is the most efficient at cleaving type II collagen.88 It may be involved in type II collagen cleavage in articular cartilage more than any other collagenase.89,90 Overexpression of constitutively active MMP-13 in cartilage induces the early onset of OA in mice.91 MMP-2 and -9 have also been reported to have collagenase activity. Most MMPs are secreted as latent proenzymes and are then activated. MMP-1 and -13, for example, can bind to collagen fibrils where they can be activated. MMP-3 and MMP-14 are involved in activating other MMPs. Like other MMPs, these activators may themselves be activated by plasmin generated from plasminogen by urokinase-type activator produced by chondrocytes or by cysteine proteinases such as cathepsin B,92 or by MT1-MMP (MMP-14), which possesses a furin cleavage site, which means that it is usually activated within the cell before secretion. Cathepsin K, a cysteine proteinase, can also cleave triple-helical collagen but in primary sites different from the primary cleavage produced by collagenases.93 It is expressed by superficial chondrocytes in increased amounts in OA cartilage.94 There is unpublished evidence thatthis proteinase is involved in collagen cleavage in OA articular cartilage in culture (V. Dejica, J.S. Mort, and A.R. Poole, unpublished data).

In OA, there is increased expression and content of various MMPs, including MMP-1, MMP-3, MMP-13, and MMP-28 (epilysin), but not ADAMTS-4 (a disintegrin and metalloproteinase with thrombospondin motifs) or ADAMTS-5, the latter being downregulated.95 Others have observed that increased MMP-13 expression is the most dominant of the collagenases, being stronger in late stage disease when MMP-3 is downregulated; of the aggrecanases, only ADAMTS-5 was clearly expressed.96 This increased expression is first observed at or close to the articular surface very early in the degenerative process.97,98 Similarly there is increased expression of MMP-2, MMP-9, MMP-8,99,100 MMP-11,101 MMP-14,102,103 and matrilysin.104 Plasminogen activator105 and cathepsin B106,107 are also upregulated in human and experimental OA cartilage.

Cleavage and denaturation of type II collagen by collagenases are first observed around chondrocytes in these superficial sites where MMP-1 and -13 are present in increased amounts.97,108,109 This change is also seen with aging but more so in OA. It extends into the territorial and interterritorial matrix and then progressively extends into the mid and deep zones with lesion development.97 Collagenase activity, the denaturation of type II collagen, and the associated loss of type II collagen73,110 are much increased in OA cartilages.73,89,108 These changes in articular cartilage are seen within months of anterior cruciate ligament rupture, which is a major risk factor for OA development in a joint.75,111

In established OA of the knee, the synthesis of type II collagen is simultaneously increased markedly.39,112,113,114 However, these new molecules, as well as resident pre-existing collagen, are the subject of excessive proteolysis.90 The same applies to aggrecan.115

The core protein of aggrecan is cleaved by different proteinases.50 Cleavage sites are common in the interglobular domain between the G1 and G2 domains as well as in the chondroitin sulfate-rich region. In the interglobular domain, two principal cleavage sites have been identified: the aggrecanase site, where proteinases such as cell surface-associated

P.33


P.34


aggrecanases or ADAMTS-4 and -5 can cleave;116 the MMP cleavage site is a target for multiple MMPs.92 There is evidence for enhanced degradation of aggrecan at both these cleavage sites in cartilage matrix in OA.117

TABLE 2-2 EARLY CHANGES IN THE STRUCTURE AND COMPOSITION OF ARTICULAR CARTILAGE IN HUMAN OSTEOARTHRITIS

Molecule Synthesis, Content, and Activity Reference
COLLAGENS
Type IIB
Content and expression - 108, 112
Denaturation + 108, 109
Cleavage by collagenase + 89, 90
Synthesis + 114
Type IIA + (mid zone) 161*, 162
Type III
mRNA and content + (surface zone) 112
Type VI + (pericellular) 164*, 165, 166
Type X + 195
PROTEOGLYCANS
Aggrecan - Surface zone; + (mid and deep zones) 50, 92
Decorin + (mid and deep zones); 160
Biglycan - (surface zone) 160
OTHER PROTEINS
Cartilage oligomeric protein Altered distribution 176
Tenascin + 178
Fibronectin + 131
Cartilage matrix protein + 177
Osteonectin + 143
PROTEINASES
Metalloproteinases
Collagenases 1, 2, 3 + 88, 206, 207
Stromelysin 1 + 50, 92
Gelatinase A + 102
Gelatinase B + 314
Matrilysin + 104
MT-MMP-1 + 102, 103
Proteinases, other
Plasmin, plasminogen + 105, 155
Cathepsin B + 106, 107*
INHIBITORS
TIMP - 154, 155
CYTOKINES, NITRIC OXIDE, GROWTH FACTORS
IL-1 and IL-1 receptor + 135, 136
TNF- receptor + 136, 137
Inducible nitric oxide synthase + 136, 140
IGF-1 + 173, 174
CHONDROCYTE HYPERTROPHY
Annexin V + 201
Type X + 195
Mineralization + 212, 213
Apoptosis + 112, 197
Parathyroid hormone-related peptide + 202, 203
These changes are typical of early degeneration with limited fibrillation at the articular surface and are mainly observed in the superficial and mid zones unless otherwise indicated. Increases (+) and decreases (-) are indicated. Experimental studies (*) are shown. These changes are reflective primarily of large joint disease.

Of the aggrecanases, ADAMTS-5 has been shown to be the major enzyme responsible for aggrecan loss in experimental murine OA118 and inflammatory joint disease.119 These proteinases are also activated outside the cell, probably at the cell surface.119,120

A variety of cleavage products of aggrecan accumulate in OA cartilages. Early in cartilage degeneration (Mankin score of 1 to 6), the excessive proteolysis leads to a reduction in molecular size of aggrecan fragments that have accumulated with aging as a result of degradation, probably during a period of many years.115 With increased degeneration, the sizes of aggrecan fragments increase.115 This may reflect altered proteolysis, but is more likely the partial degradation of more recently synthesized larger molecules. That new aggrecan synthesis and incorporation occur is reflected by the appearance of epitopes on chondroitin sulfate that are normally only commonly found in actively biosynthetic fetal cartilages.115,121,122,123 Some of these epitopes that increase in OA in synovial fluid, such as the 846 epitope,124,125 are thought to present on newly synthesized molecules that have been released by matrix degradation.126

The Causes and Regulation of Cartilage Matrix Degradation

The reasons for this increased proteolysis have been widely studied. Degradation products of matrix molecules may themselves stimulate degradation through chondrocyte and synovial cell receptor-mediated activation, forming a chronic cycle (Fig. 2-4). Different fragments of fibronectin can stimulate chondrocyte-mediated cartilage resorption by cell surface receptor activation127,128,129 just as in synovial fibroblasts, where RGD-integrin receptor activation is involved.130 Fibronectin is produced in increased amounts in OA cartilage.131 Its degradation may therefore play an important role in establishing positive feedback generation of proteolysis. Cellular responses in OA cartilage involve the production of cytokines such as interleukin (IL)-1, which are known to stimulate degradation50 and also play an essential autocrine and paracrine role132 in fibronectin fragment-mediated degradation (Fig. 2-4). Fragments of type II collagen can, when present in sufficient concentration, also induce matrix resorption.133,134

Figure 2-4 Responses of chondrocytes to mechanical forces, cytokines, and matrix degradation products generated by MMPs.

In OA, there is increased expression on chondrocytes of the receptors for IL-1135 and of IL-1 itself, even more than in rheumatoid arthritis,136 as well as the receptor for tumor necrosis factor a (TNF- )99,137 (Table 2-2). The

P.35


presence of the TNF- p55 receptor (but not the p75) on OA chondrocytes correlates with the susceptibility of cartilage explants to TNF- -induced proteoglycan loss.137

IL-1 and TNF- are potent activators of cartilage degradation in vitro.50 In combination with oncostatin M, IL-1 is even more potent in causing cartilage resorption, but levels of oncostatin M, a member of the IL-6 family, are not usually elevated in synovial fluid in OA.138 Inhibition of IL-1 or TNF- by biologic antagonists can suppress cartilage matrix resorption in articular cartilage in culture. It can also stimulate an increase in aggrecan content,139 probably by inhibiting degradation of newly synthesized molecules.

Inducible nitric oxide synthase (iNOS, or NOS2) is upregulated in OA chondrocytes140 more so than in rheumatoid articular cartilage,136 leading to an increased generation of nitric oxide. IL-1 , IL-1b , and TNF- are potent stimulators of nitric oxide production in cartilage141 in a manner that can be arrested by osteopontin,142 which is also upregulated in OA cartilages.143 The expression by chondrocytes of iNOS, TNF-a, IL-1 , and IL-1 are correlated in arthritis136 (Fig. 2-4). Nitric oxide mediates the inhibition of aggrecan synthesis induced by IL-1.144 However, protease activity and proteoglycan degradation are enhanced when nitric oxide production is blocked,145,146 suggesting that nitric oxide may also play a protective role. Nitric oxide can also induce apoptosis in chondrocytes,147 but only in the presence of other reactive oxygen species. Inhibition of iNOS reduces the progression of experimental OA.148 Nonexpression of IL-1 or IL-1 converting enzyme or iNOS can also accelerate the development of surgically-induced murine OA. These observations together point to the importance of physiological amounts of these molecules to maintain a healthy joint,146 and suggest that complete suppression of activity could be detrimental from a therapeutic dosing standpoint.

Changes in matrix loading can also induce the production of a variety of pro-inflammatory mediators and promote matrix degradation as well as alter the synthesis of matrix molecules50,78 (Fig. 2-4). Mechanical compression causes a dose-dependent increase in the synthesis of nitric oxide and prostaglandin E2 through the activation of NOS2 and cyclooxygenase 2 (COX2), respectively, with significant interaction between the NOS2 and COX2 pathways.149,150 Injurious mechanical compression of cartilage explants at high loading rates and magnitudes results in significant upregulation of MMPs and aggrecanases151,152 that are also associated with increased loss of proteoglycans as well as other biomarkers of OA.153

Figure 2-5 Changes observed in cartilage and bone in OA and factors that protect against, accelerate, or are associated with the OA process.

The pathologic changes in cartilage matrix structure in OA are likely to cause fundamental disturbances in the normal balance resulting from controlled mechanical loading and cytokine and growth factor signaling, which leads to changing gene expression of matrix macromolecules, signaling molecules, and enzymes (Fig. 2-5). Activities of these MMPs are regulated not only at the levels of transcriptional activation, translation, and extracellular proenzyme activation, but also extracellularly at the level of inhibition by tissue inhibitors of metalloproteinases (TIMPs). Four such inhibitors have been described, namely, TIMP-1, -2, -3, and -4. These react with the active MMP in a 1:1 molar ratio. In OA, there is a deficiency of TIMP activity154,155 favoring excessive proteolysis (Table 2-2). TIMP-3, which is the only TIMP that can bind to extracellular matrix, is capable of inhibiting aggrecan degradation in hyaline cartilage.156

P.36


Its expression is upregulated in OA cartilage whereas TIMP-1 and -4 are downregulated.95

TGF- 1 can downregulate MMP-1 and -13 as well as IL-1 and TNF receptors on OA chondrocytes.99 Yet, it can stimulate ADAMTS-4 expression and aggrecan degradation.157 TGF- 2 can rather selectively suppress the cleavage of type II collagen by collagenases in OA cartilage in culture and reduce the expression of MMPs and proinflammatory cytokines: this also involves suppression of hypertrophy associated genes.158 The Chitinase 3-like protein glycoprotein 39 (GP39), which is upregulated in OA cartilage, can suppress chondrocyte induction of proinflammatory cytokines, chemokines, and collagenases by IL-1 and TNF- 159

General Changes in Cartilage Matrix Protein Content and Gene Expression

The early damage to the more superficial matrix in early OA is accompanied by an increased content of biglycan and decorin160 and aggrecan (A.R. Poole and A. Reiner, unpublished data) in the mid and deep zones, no doubt to compensate for the increased loading on the chondrocytes and the damage to and loss of these molecules from the more superficial cartilage. This accompanies the marked increase in the synthesis of type II procollagen in these deeper sites,114 mainly type IIB as revealed by experimental studies113 but also some type IIA,161,162 which is normally observed only before chondroblast differentiation early in development.163 Overall there is a loss of type II collagen, starting in the more superficial cartilage.108 There is limited expression and synthesis of type III collagen.112 Type VI content is, however, increased164,165,166 and its filamentous structure is altered,167 presumably as a result of pericellular remodeling which frequently results in an enlargement of the pericellular matrix.168 These changes are also reflected as a loss of mechanical properties of the pericellular matrix,169,170 which results in significant alterations in the biomechanical environment of the chondrocytes.54

Aggrecan and link protein synthesis or expression are upregulated in response to the increased damage.50,115,121,171 There is a general increased expression and synthesis of the proteoglycans versican, fibromodulin, lumican, decorin, and biglycan.171,172 This is associated with an increase in insulin-like growth factor 1 (IGF-1)173 and its receptor.174 IGF-1 is a potent stimulant of aggrecan synthesis and may be responsible for this increase. This increased synthesis is often seen in the same sites where degradation is enhanced. Although bone morphogenetic proteins play key roles in promoting matrix synthesis, BMP-3 is markedly downregulated in OA.175

There are changes in the distributions and contentsof many other matrix molecules in OA.50 Cartilage oligomeric protein (COMP) is altered in distribution.176 Contents of cartilage matrix protein,177 tenascin,178 osteonectin,143 and fibronectin131 and other molecules are increased with the result that the cartilage reverts in part to a more fetal tissue. These changes are summarized in Table 2-2 and Figure 2-3.

The Chitinase 3-like protein GP 39 (YKL-39), but not YKL-40, is also markedly upregulated in OA cartilage. This protein can modulate IL-1 activity as well as suppress apoptosis and stimulate cell division. It therefore seems to offer protection against degeneration.159,179 The significance of a variety of early changes in gene expression in OA detected in articular cartilage and of genes that are expressed in cartilage which change in expression in peripheral blood cells in early OA remains to be determined.180

The Regulation of Cartilage Matrix Assembly in Osteoarthritis

Throughout the development of OA, synthetic processes are probably much influenced by the degradative cytokines IL-1 and TNF- , the major anabolic growth factors IGF-1, the transforming growth factor- (TGF- ) family,181 the bone morphogenetic proteins (particularly BMP-2 and -7), and the anti-inflammatory or modulatory cytokines of the synovium, cartilage, and other tissues, including platelet-derived growth factor, fibroblast growth factor-2, IL-4, IL-6, IL-10, and IL-13. These and other regulatory molecules contribute in many different ways to anabolism.50,181,182 Cytokines such as IL-1 can inhibit matrix synthesis whereas IGF-1 can suppress this inhibition.183 IGF-1 and mRNA levels are increased in OA articular tissue.184 IGF-1 can decrease the degradation and inhibition of synthesis induced by IL-1.185 IGF-1 release from chondrocytes is in fact stimulated by IL-1 and TNF- 186 In spite of the increase in IGF-1 in OA, OA chondrocytes are hyporesponsive to this growth factor. This may be because IGF-1 activity is excessively restricted by IGFBPs, which are also upregulated174,187 by cytokines such as IL-1. Proteases can also cleave these binding proteins, regulating their activity.186,188 Fibroblast growth factor-2 stimulates cell proliferation in articular chondrocytes but does not stimulate synthesis of glycosaminoglycans.189,190 Synergistic relationships have been demonstrated between growth factors and cytokines in articular cartilage, which help regulate important cellular processes. Chondrocyte proliferation can be amplified by combinations of IL-6 and TGF- or of fibroblast growth factor and IGF.186

Moreover, TGF- stimulates type II collagen and aggrecan gene expression and inhibits metalloproteinase mRNA expression in synovial fibroblasts and chondrocytes.181,182 It can suppress IL-1 expression and type collagen cleavage in cartilage,158 stimulate production of plasminogen activator inhibitor protein 1,191 and also TIMP synthesis,50 thereby regulating proteolysis and enhancing synthesis. However, injection of TGF- in the knee joint of mice results in increased osteoarthritic degeneration, possibly due to increased bone formation and remodeling.192,193 The balance between these pro-proliferative, anabolic, and catabolic molecules is very complex and clearly fundamentally perturbed in OA and by injury that leads to an altered mechanical environment of the chondrocytes, affecting the biosynthesis of these molecules. Studies of Guinea pig strains with different rates of natural OA development have revealed from serum measurements of cartilage molecular markers of type II collagen synthesis and degradation that the ratio of the markers favors degradation over synthesis in more rapid progressors194 (N. Gerwin, K. Rudolphi, A.R. Poole, et al., unpublished data).

P.37


Changes in the Chondrocyte Phenotype in Relationship to Cell Death, Matrix Degradation, and Calcification in Osteoarthritis

In previously uncalcified articular cartilage in OA, there is induction of type X collagen expression and synthesis.195 Expression of this and other hypertrophy-related genes including MMP-13 is seen very early in micro-lesion development.98 Hypertrophy is normally only seen in the growth plate when the extracellular collagen network is partially resorbed by MMP-1350,97,196 and then calcified as cells die as a result of apoptosis. In OA, these events including apoptosis197,198 reappear in degenerate OA cartilage although some expression data argue against this.199 The increase in apoptosis corresponds to a reduced cell density and expression of caspase 3.200 These changes are in association with expression of the cell surface type II collagen receptor annexin V,201 also highly expressed by early hypertrophic chondrocytes. Parathyroid hormone-related peptide (PTHrP), which is also produced by prehypertrophic cells in the growth plate and suppresses hypertrophy, is also upregulated in OA cartilage.202,203 The calcium-sensing receptor, expressed by hypertrophic cells, is upregulated with the onset of OA in the guinea pig together with PTHrP.204

These changes are initially observed in OA cartilage mainly in the more superficial and mid zones where damage to collagen is most pronounced205 (Fig. 2-3), and may represent a chondrocyte response to a damaged extracellular matrix with the reversion to a more fetal phenotype. There is also a marked increase in expression of type II collagen112,113,114 and MMP-13,88,206,207 which are also a feature of the shift to hypertrophy in the growth plate.208 An inhibitor of MMP-13 can suppress hypertrophy,97 suggesting that MMP-13 is a key proteinase in this process and that excessive collagen cleavage is a trigger for further chondrocyte differentiation.

In the partially calcified OA cartilage delimited by the tidemark and bordered by the subchondral bone, there is a reactivation of endochondral ossification characterized by upregulation of type X collagen expression and duplication or replication of the tidemark, separating this zone from uncalcified cartilage. Vascular invasion from subchondral bone is reinitiated, resembling that seen in endochondral ossification. Moreover, osteophyte formation, an endochondral process, is initiated peripheral to the articular cartilages, leading to bone spurs capped with articular cartilage (Fig. 2-1). Thus, there is a major shift in the physiology of the articular cartilage characteristic of a superimposition of endochondral changes within and associated with articular cartilages. We know that TGF- signaling is essential to prevent hypertrophy since mice with a functionless type II receptor209 or a deletion in a Smad signaling component210 develop a rapid degeneration of articular cartilage associated with extensive chondrocyte hypertrophy and the formation of large osteophytes. PGE2, generated when TGF- 2 is added to OA cartilage, suppresses hypertrophy and collagenase activity.158 A transcription factor, early growth response protein-1 (Egr-1), which can stimulate TGF- expression and suppress apoptosis, is very down-regulated in OA, favoring a drop in TGF- content.211

Hypertrophy is accompanied by calcification of the normally uncalcified extracellular matrix of articular cartilage. More than 90% of OA subjects show evidence for limited calcification of articular cartilages;212 there is also a high incidence of hydroxyapatite crystals in joint fluids.213,214 In calcium pyrophosphate deposition (CPPD) disease, joints affected by mineral deposition, such as the shoulders, wrists, and metacarpophalangeal joints, are different from those affected in idiopathic OA215 (see Chapter 13).

Although CPPD and hydroxyapatite crystal aggregates can, under some conditions, be markedly phlogistic in both animal models and cell culture, there is equivocal evidence in cartilage that crystals can injure chondrocytes directly.216,217,218 As in growth cartilage matrix, matrix vesicles have been detected in OA articular cartilages, often with associated minerals as well as cellular generation of pyrophosphate.219 Apoptotic cell particles may also generate minerals in the presence of adenosine triphosphate.220 These changes in OA cartilage are summarized in Table 2-2 and Figure 2-3.

Physiologic concentrations of extracellular pyrophosphate (PPi) regulate calcification that results in basic calcium phosphate crystal deposition. Hydrolysis of excess PPi promotes crystal formation in OA via elevated inorganic phosphate generation.221 PPi is generated from nucleoside triphosphates by nucleotide pyrophosphatase phosphodiesterase PC-1. The extracellular concentration of PPi is regulated via the trans-membrane protein ANK. In OA cartilage, ANK is upregulated.222 This would this favor the elevation of extracellular PPi. Interestingly, increased ANK expression promotes MMP-13 expression,222 providing a link between calcification and increased MMP-13 expression, both of which are features of chondrocyte hypertrophy.50

The Remodelling of Subchondral Bone in Osteoarthritis

Pronounced changes in subchondral bone occur in OA, and may even occur in sites remote from the affected joints. Femoral neck bone is less stiff and less dense in hip OA, but not as reduced in density as in osteoporosis.223 Bone mineral density remote from degenerate joints in arms and spine is elevated.224,225 Bone loss is elevated in men with hip OA.225 Women with incident knee OA have a higher bone mineral density in the spine and hip than those without disease.226 In generalized OA, there is also hypermineralization;227 yet in hand OA in women, there is evidence that with increasing grade of OA, bone mass is decreased.228 Subchondral bone in hip OA is less mineralized and has more osteoid indicative of incomplete mineralization.229,230 The elevated content of osteoid is accompanied by increased synthesis and content of type I collagen and increased alkaline phosphatase content.230 Together, these changes are indicative of increased turnover of bone in hip OA. TGF- a bone growth factor, is also increased in content.230 Such changes are also seen in remote sites, such as in the iliac crest, where osteocalcin, IGF-1, IGF-2, and TGF- contents are increased.227

P.38


These observations raise the possibility that bone turnover may be different in people before the development of OA and that these changes may vary according to the type of OA. They may clearly contribute to or even cause the onset of OA. Analyses of type II procollagen production in OA articular cartilages reveal increased synthesis reflected by elevated C-propeptide concentration. Yet in the peripheral blood of OA patients, there is a significant reduction of the C-propeptide. This also suggests a systemic alteration in cartilage type II collagen synthesis, which again may be a risk factor for the onset of OA.114 Observations of this kind raise questions as to whether idiopathic OA may result more from specific fundamental differences wherein local factors (such as biomechanical) precipitate joint disease via interactions with systemic metabolic factors in any given individual. Thus systemic bone or cartilage changes alone may predispose people to these degenerative changes in joints where the mechanical environment may reveal such abnormalities. It is interesting to note that peak bone mass is increased in the hips of daughters of women with hand OA.231 Clearly there would appear to be genetic linkages in the etiology of this disease.

The degeneration of articular cartilage may be accompanied or possibly preceded by increased subchondral bone turnover, as suggested by bone scintigraphy. In the absence of MRI, it is not possible to be clear as to whether the bone changes preceded those in articular cartilages. Such bone changes are predictive of the development of hand OA,232 the progression of knee OA,233 and generalized nodal OA.234 Studies in humans and animals, such as the aging guinea pig, using histologic analyses and magnetic resonance imaging, suggest that degenerative changes in articular cartilages are accompanied by local changes in subchondral bone that involve cyst formation and altered trabecular and osteoid thickness and bone formation and turnover.235,236,237,238

The development of osteophytes so often seen peripherally in an OA joint involves the formation of a cap of new articular cartilage as well as new bone formation as part of an endochondral process. Experimentally, it has been shown to be correlated with articular cartilage loss elsewhere in the joint.239 These changes and relationships are often compartmental, suggesting localized events. The overall increase in bone turnover in OA and resultant changes are reflected by increases in bone-specific deoxypyridinoline cross-links in urine as a result of enhanced osteoclastic resorption of bone.126,240 These changes in urine may also result from more extensive systemic change.

Changes in both the articular cartilages and the subchondral bone of a degenerate joint no doubt reflect the interdependence of these tissues within the joint. Changes in one tissue would influence mechanical loading of the other and alter tissue turnover. Thus, it is not surprising that such changes may occur simultaneously. The presence of bone marrow abnormalities on MRI has been reported to be associated with cartilage remodeling based on a type II collagen biomarker assay.241 Whether this marker reflects changes in uncalcified cartilage or in the calcified cartilage adjacent to subchondral bone or both is not known. It is noteworthy that in the ankle, where OA is much less common than in the hip or knee, cartilage degeneration of the talar is not associated with an increase in subchondral bone density,242 indicating different responses in ankle joints that may influence OA development. Studies of bone changes in familial OA (caused by a cartilage collagen mutation, for example) compared with idiopathic OA may be of use in determining whether these cartilage changes occur together with or independently of those in bone.

In osteoporosis, a condition that results in excessive resorption and a net loss of bone mineral density, it is uncommon to observe OA (Fig. 2-5), and vice versa.243,244,245 The structure, turnover, and density of bone would therefore appear to influence articular cartilage turnover and play a significant role in the pathogenesis of OA. Thus, a controlled reduction in bone density may help in controlling joint degeneration.

The interrelationships between bone and cartilage changes in OA may also result from the molecular effects of the products of one tissue on another. Bone cell cultures from OA subchondral bone, but not from nonarthritic bone, can stimulate proteoglycan release from nonarthritic human articular cartilage.246 Further studies of this kind are needed.

Calcitonin suppresses osteoclastic resorption of bone, a key component of the remodeling of bone. It can also reduce both cartilage pathology85 and suppress bone changes247 in experimental dog OA. Whether this effect is achieved through the control of bone changes or also involves a more direct stimulation of proteoglycan synthesis by chondrocytes248 is unclear. Alendronate, another potent inhibitor of bone resorption, can also reduce cartilage degeneration and osteophyte formation in experimental OA in the rat,249 again suggesting that bone remodeling in OA plays an important role in the degenerative process.

Synovitis and Inflammation in Osteoarthritis

OA is not considered an inflammatory arthritis, and the synovial fluid leukocyte count is characteristically less than 3000 cells/mL. To the extent that acute synovial fluid leukocytic inflammation does occur in OA, it is often the result of secondary crystal-induced synovitis (either calcium apatite or calcium pyrophosphate dihydrate). However, low-grade inflammatory processes nevertheless occur in osteoarthritic synovial tissues that contribute to disease pathogenesis, and many of the clinical symptoms and signs seen in OA joints clearly reflect synovial inflammation (e.g., joint swelling and effusion, stiffness, occasionally redness). Synovial histological changes include synovial hypertrophy and hyperplasia with an increased number of lining cells, often accompanied by infiltration of the sublining tissue with scattered foci of lymphocytes. Cartilage breakdown products, derived from the articular surface as a result of mechanical or enzymatic destruction of the cartilage, can provoke the release of collagenase and other hydrolytic enzymes from synovial cells and macrophages.250

P.39


Cartilage breakdown products are also believed to result in mononuclear cell infiltration and vascular hyperplasia in the synovial membrane in OA.251 A consequence of these low-grade inflammatory processes is the induction of synovial IL-1 and TNF- , which are likely contributors to the degradative cascade.

Recent studies have also demonstrated that, even in the absence of clinical signs of synovitis, there is frequently localized synovitis in patients with OA, which may be most pronounced immediately adjacent to the OA lesion of the articular cartilage. For example, arthroscopy has demonstrated localized synovial proliferative and inflammatory changes in up to 50% of patients with knee OA.252,253 Proteases and cytokines produced by activated synovium have been suggested to accelerate deterioration of contiguous cartilage lesions.253 Areas of increased radionuclide uptake ( hot spots ) on bone scintigraphy have also been reported to identify joints more likely to progress by radiographic criteria or to require surgical intervention over a 5-year period.233

Synovial tissues and cartilage also synthesize anti-inflammatory cytokines, including IL-13, IL-4, and IL-10. Indeed, synovial fluid of OA patients contains increased levels of these factors, which can decrease PGE2 release, IL- , TNF- , and MMPs while upregulating TIMP.254,255 Another anti-inflammatory molecule produced by joint tissues is the IL-1 receptor antagonist (IL-1Ra), which, like soluble IL-1 receptors discussed above, competes with cell surface receptors for IL-1 and thereby reduces nitric oxide production, PGE2 synthesis, or protease secretion. Although the production of IL-1Ra is increased in OA, increased production is insufficient to reverse the catabolic effects of augmented IL- .

A number of investigations have revealed that serum HA is upregulated in patients with rheumatoid arthritis and is related to joint inflammation and radiologic progression.256 HA is also often increased in OA patients.257 Those with persistently elevated serum HA levels exhibit more rapid disease progression.258 HA levels also indirectly correlate with minimal joint space.258,259 COMP is synthesized by synovial cells as well as by chondrocytes. COMP synthesis is stimulated by TGF- 1, which is produced in inflammation.260 Patients who show serum elevations of COMP often exhibit progression of joint damage.261,262,263 Serum/plasma levels appear to be closely associated with hip synovitis241 and knee OA.264 Elevation of both HA and COMP is associated with a greater risk of progression in hip OA.265

A glycosylated derivative of the pyridinoline collagen cross-link is enriched in human synovium and present in low levels in cartilage and other soft tissues.266 It can be detected in urine. Its close association with the presence of knee OA and severity267 suggests the presence of synovitis, a feature of which is increased content of type I and III collagens and their turnover.

MMP-3 and MMP-9, which are elevated in synovial cells from patients with rapidly destructive hip OA, are also elevated in synovial fluid, plasma, and sera in these patients.268,269 The elevation of MMP-9 is markedly reduced a year after arthroplasty, pointing to the joint cartilages and/or bone as the cause of this elevation, probably by triggering a synovitis. This may involve an immune reaction since T lymphocyte immunity to cartilage aggrecan and link protein is enhanced in OA.270 In experimental OA, T cell immunity to fibrillar collagens also develops.271

Chemokines are also elevated. Macrophage inflammatory protein-1 is elevated in OA synovial fluid over serum as in RA.272 Eotaxin-1, RANTES, and MCP-1 areelevated in plasma of OA patients.273

Fas ligand, which induces apoptosis, is also present in synovial fluid in OA274 and may account in part for the enhanced apoptosis in articular cartilage in OA by engagement of the receptor for this ligand on chondrocytes and synovial cells. The presence of systemic inflammation in OA, albeit limited, is also revealed by the small but significant increase in the serum acute-phase molecule C-reactive protein,275 which is predictive of progression in early onset disease.276 These are but some examples of evidence for an inflammatory process in OA that in some cases may be more pronounced within a joint and may accelerate joint degeneration by the release of cytokines, chemokines, proteinases, and other mediators that cause joint degeneration.

Pain in Osteoarthritis

Arthritis pain is the most common cause of pain in aged populations277 and arguably the most debilitating aspect of OA. Although the presence of knee pain increases with radiographic disease severity in most studies, it is apparent that the severity of abnormalities by routine radiography does not correlate with pain severity in the individual patient. MRI studies have indicated that in patients with knee OA, knee pain severity was associated with subarticular bone attrition, bone marrow lesions, synovitis/effusion, and meniscal tears.278

Pain severity in persons with OA is not a simple phenomenon, and can arise from any of several innervated tissues. The joints of the appendicular skeleton are innervated by the peripheral nervous system in every tissue except cartilage where innervation is peripheral in the periosteum, synovium, capsule, ligaments, and subchondral bone. Here nociceptors monitor the environment. Neuro-innervation can determine disease onset. Limbs paralyzed by hemiplegia or poliomyelitis are often spared in the development of OA or RA.279,280 Patients with different joint involvement in OA can exhibit pronounced hyperalgesia to thermal and mechanical stimuli including uninvolved joints.281,282,283 Following total hip arthroplasty, there is a return to normal pain thresholds in the contralateral hip.284 The association of movement with pain suggests the contribution of central mechanisms.285 Nerve growth factor plays a major role in inflammatory hyperalgesia: IL-1 contributes to the upregulation of this neurotrophin and hyperalgesia.286 There is evidence for an association between pain, synovitis, and changes in subchondral bone.287 Neuropeptides generated by nerve fibers are messengers that link the peripheral nervous system and

P.40


inflammation. C-fiber stimulation causes Substance P release, which in turn induces mast cell and platelet degranulation with histamine, serotonin, bradykinin, and platelet activating factor release, capable of not only causing vasodilation and vasopermiabilisation but of stimulating nociceptors and amplifying and prolonging afferent discharge into the central nervous system.288 Prostaglandins can also contribute to painful afferent sensory responses.289,290 Substance P can directly activate synovial cells to induce collagenase and PGE2 release291 and stimulate monocytes/macrophages and other inflammatory cells to produce IL-1, TNF- and IL-6.292

Neuropeptides can also perform various regulatory functions. Calcitonin gene-related peptide and vasoactive intestinal peptide (VIP) can each inhibit synovial cell proliferation and the expression of proinflammatory cytokines and MMP-2.293 Somatostatin can modulate the production of MMP-1 and MMP-9.294 The parasympathetic vagus nerve, which innervates all major organs, can control TNF- synthesis. It has sensory (input) and motor (output) fibers to sense and suppress inflammation. The sensory nerve can detect IL-1 and respond by releasing acetylcholine which suppresses TNF- and IL-1 generation by macrophages. Electrical stimulation of this nerve prevents TNF release from macrophages. Surgical section removes this protection.295 This is because the acetylcholine normally released by the nerve activates 7 nicotinic acid receptors on macrophages which suppress the release of these proinflammatory cytokines.296 The sympathetic nervous system mediates unloading-induced bone loss through the suppression of bone formation by osteoblasts and enhanced resorption by osteoclasts.297 Leptin has an anti-osteogenic function that is mediated by sympathetic neural pathways that control bone remodeling. This occurs by sympathetic signaling of 2 adrenergic receptors on osteoblasts. In mice deficient in the receptor Adrb2, the sympathetic system favors bone resorption; this occurs by an increase in Rankl expression on osteoblast progenitor cells. Leptin controls the expression of the neuropeptide cocaine amphetamine regulated transcript 9CART, which inhibits bone resorption by modulating Rankl expression. Thus leptin regulated neural pathways control both aspects of bone remodeling.298

Neuropeptides can also help explain some of the effects of the nervous system on the skeleton. VIP potentiates IL-6 production by osteoblasts induced by proinflammatory osteotropic cytokines, including IL-1.299 Substance P stimulates osteoclast formation.300 Chondrocytes express functional -opioid receptors that can be activated with -endorphin301 causing increased IL-1 and TNF- generation.302

Ipsilateral and contralateral joint involvement whereby joint inflammation in a single joint can induce distal bilateral degeneration of articular cartilage has also been ascribed to neurogenic mechanisms and neuropeptides by using an antagonist of neurokinin-1 and spinal compression to inhibit the involvement of the other joint.280

Thus it can be seen with these examples that neurological mechanisms influence not only the perception of pain that is altered in OA but also regulate skeletal turnover and inflammation. There is still much to be learned of this much neglected area of research in our understanding of the pathogenesis of OA.

Therapeutic Targets for the Management of Osteoarthritis

Cartilage is clearly a principal target because of its paramount importance in joint articulation. Despite the tremendous impact of this disease, however, current therapies are palliative and there are no disease-modifying OA drugs (DMOADs) currently available for clinical use. Nonetheless, a number of promising molecular targets exist in the development of pharmacologic therapies for OA, and useful and comprehensive review of such therapeutic approaches has recently been published.303

For example, there has been significant emphasis on targeting MMPs that cleave collagen because collagen degradation is clearly indicated in OA pathogenesis. The collagenase MMP-13 is thought to be an important target for the control of collagen fibril damage in articular cartilage. To date, doxycycline is the only molecule that can regulate collagenase activities in vitro,304 control the progression of experimental OA,305 and be used without serious side effects in the treatment of knee OA.306

Cartilage proteoglycan aggrecan degradation by aggrecanases is also of very much interest as a target. Although the aggrecanase ADAMTS-5 may be rate limiting in the degradation of this molecule in animals (see section on Proteinases of OA Cartilage), it remains to be seen whether it is in humans. This is because aggrecan, unlike type II collagen, is susceptible to cleavage by many proteinases such as those of the ADAMTS family. Based upon experiments in knockout mice, ADAMTS-5 has been reported to be the primary aggrecanase responsible for aggrecan degradation in a murine model of OA, and could be a potential target for therapeutic intervention in human OA.119

Another approach is to restore the balance between synthesis and degradation by enhancing cartilage matrix synthesis. A clearer identification of the key growth factors involved in matrix assembly is a priority, particularly since some of these can suppress degradation. IGF-1 is of obvious importance in view of its potency and upregulation in OA,173 but there may be more potent growth factors or combinations thereof that can be used to renew matrix assembly and control degradation. A combination of MMP inhibitors and enhanced stimulation of matrix synthesis may prove most effective.

Work with bone morphogenetic protein 7 has revealed its capacity to promote matrix assembly in articular cartilage and to inhibit the degradative effects of IL-1.307 We have already discussed how TGF- 2 is a potent stimulant of matrix synthesis,50 and can suppress the cleavage of collagen by collagenases and chondrocyte hypertrophy in OA cartilage in culture.158 Still other approaches have targeted the pro-inflammatory mediators such as nitric oxide through selective inhibition of NOS2 and animal models of OA.148 These studies have shown significant promise for inhibiting disease progression by decreasing chondrocyte

P.41


apoptosis as well as cartilage catabolism via inhibition of NOS2. While the mechanisms for these findings are not fully understood, they may involve the local suppression of other inflammatory cytokines such as IL-1 or TNF- .

Brandt et al. recently reported that doxycyline slows joint space narrowing (JSN) in patients with OA of the medial tibiofemoral compartment.306 In this placebo-controlled trial, 431 obese women (ages 45 to 64 years) with unilateral radiographic knee OA were randomly assigned to receive 30 months of treatment with doxycycline 100 mg or placebo twice daily. Doxycycline reduced JSN in the OA knee by approximately 30% at 30 months; however, the mean progression of JSN in both groups was limited (0.30 0.60 mm vs. 0.45 0.70 mm).

IL-1 has attracted significant interest as a target for disease modification in OA. The intra-articular injection of recombinant human IL-1Ra attenuates the development of cartilage lesions and the expression of MMPs in the canine experimental OA model.308 Two groups have also demonstrated that in vivo transfer of the IL-1Ra gene prevents disease progression in the meniscectomy rabbit model of knee OA.309,310 In human OA, the benefit of intra-articular IL-1Ra injection for symptomatic knee OA has been reported in an open-label 12-week study;311 however, in a follow-up controlled trial by these authors, intra-articular injection of IL1-Ra reduced knee pain at day 4, but was not more effective than placebo at 1 month. In addition to biological IL-1 antagonists, small molecule drugs are in development that interfere with the conversion of intracellular precursor IL-1 to active IL-1 by inhibiting caspase-1, the iNOS. Fundamental changes in bone in OA and the relative absence of OA in patients with osteoporosis (and vice versa) raise issues as to the possibility of controlling cartilage degeneration by modifying bone turnover and density. This hasalso been reviewed in the section on bone changes. Would controlled loss of bone density slow the process of degeneration? To date, DMOAD trials using the bisphosphonate to slow progression of knee OA have had mixed results.312

Regulation of synovitis is another target. About 10% of the OA population is thought to have a pronounced synovitis, comparable in numbers to the total rheumatoid population. We know that the incidence of much of the hip and knee OA is greater in women than in men, but we know almost nothing of the reasons for this. Clearly, much needs to be done to determine whether sex hormones regulate cartilage turnover.

None of these treatments will be effective if we rely on existing clinical assessments for the degenerative process. When most patients present with an OA joint, the disease is usually advanced and much damage has been done. For the future, it is important to detect and treat the disease early by employing new detection systems such as new imaging and biomarker modalities be they protein or gene based. Only then will therapy be most likely to succeed. Inevitably, this will mean the introduction of screening programs to identify these early changes. Otherwise, we must be content for now with halting further degeneration and slowing OA development elsewhere. We must also not forget the opportunities to treat single large joints such as the knee. This is an attractive consideration in view of the experience gained with intra-articular therapy with HA preparations and steroid usage. It also offers an opportunity to avoid or minimize possible side effects.

If we can effectively control the pain of OA, we will have already made a great stride forward. Much research has pinpointed new targets to control the pain of arthritis. Pharmacologic stimulation of the vagus nerve cholinergic anti-inflammatory pathway offers new therapeutic opportunities.313 These discoveries are being very actively pursued. We will probably find that many disease modifying drugs lack symptomatic relief for the patient. Inevitably, combination therapy will be needed in such cases.

Conclusions

In the last two decades, we have made significant progress in our understanding of the pathophysiology of OA. This has revealed a complex series of molecular changes at the cell, matrix, and tissue levels and complex interactions between tissues that make up the joint. Some of these are summarized in Figure 2-2 and Figure 2-5. We can now better understand how these changes develop and progress, as well as the fact that tissues other than cartilage are involved in the disease process. Opportunities for the management of disease progression and targets for therapeutic intervention have been identified and can now be tested for the first time. The future of OA research is extremely promising, and some real opportunities for the effective therapeutic management of OA are now available.

Acknowledgments

The authors work has been funded by Shriners Hospitals for Children; the Canadian Institutes of Health Research and the Canadian Arthritis Network (A.R.P.); and the National Institutes of Health (A.R.P., F.G., S.B.A.).

References

1. Altman R, Asch E, Bloch D, et al. Development of criteria for the classification and reporting of osteoarthritis. Classification of osteoarthritis of the knee. Diagnostic and Therapeutic Criteria Committee of the American Rheumatism Association. Arthritis Rheum 29:1039 1049, 1986.

2. Brandt KD, Mankin HJ, Shulman LE. Workshop on etiopathogenesis of osteoarthritis. J Rheumatol 13:1126 1160, 1986.

3. Holderbaum D, Haqqi TM, Moskowitz RW. Genetics and osteoarthritis: exposing the iceberg. Arthritis Rheum 42: 397-405, 1999.

4. Olsen BR. Role of cartilage collagens in formation of the skeleton. Ann N Y Acad Sci 785:124 130, 1996.

5. Lieberman SA, Bjorkengren AG, Hoffman AR. Rheumatologic and skeletal changes in acromegaly. Endocrinol Metab Clin North Am 21:615 631, 1992.

6. Evans CH, Brown TD. Role of physical and mechanical agents in degrading the matrix. In: Woessner JFJ, Howell DS, eds. Joint Cartilage Degradation; Basic and Clinical Aspects. New York: Marcel Dekker; 1993:187 208.

7. Radin EL, Burr DB, Caterson B, et al. Mechanical determinants of osteoarthrosis. Semin Arthritis Rheum 21:12 21, 1991.

P.42


8. Buckwalter JA. Osteoarthritis and articular cartilage use, disuse, and abuse: experimental studies. J Rheumatol Suppl 43:13 15, 1995.

9. Palmoski MJ, Brandt KD. Running inhibits the reversal of atrophic changes in canine knee cartilage after removal of a leg cast. Arthritis Rheum 24:1329 1337, 1981.

10. Setton LA, Elliott DM, Mow VC. Altered mechanics of cartilage with osteoarthritis: human osteoarthritis and an experimental model of joint degeneration. Osteoarthritis Cartilage 7:2 14, 1999.

11. Setton LA, Mow VC, Muller FJ, et al. Mechanical properties of canine articular cartilage are significantly altered following transection of the anterior cruciate ligament. J Orthop Res 12:451 463, 1994.

12. Setton LA, Mow VC, Muller FJ, et al. Mechanical behavior and biochemical composition of canine knee cartilage following periods of joint disuse and disuse with remobilization. Osteoarthritis Cartilage 5:1 16, 1997.

13. Grumbles RM, Howell DS, Howard GA, et al. Cartilage metalloproteases in disuse atrophy. J Rheumatol Suppl 43:146 148, 1995.

14. Helminen HJ, Jurvelin J, Kiviranta I, et al. Joint loading effects on articular cartilage: A historical review. In: Helminen HJ, Kiviranta I, Tammi M, et al., eds. Joint Loading: Biology and Health of Articular Structures. Bristol: Wright and Sons; 1-46, 1987.

15. Cheung HS, Setton LA, Guilak F, et al. Upregulation of metalloproteinases in articular cartilage resultant from disuse immobilization, as well as subsequent vigorous exercise. Trans Orthop Res Soc 24:40, 1999.

16. Lammi MJ, Hakkinen TP, Parkkinen JJ, et al. Effects of long-term running exercise on canine femoral head articular cartilage. Agents & Actions Supplements 39:95 99, 1993.

17. Lane NE, Buckwalter JA. Exercise: a cause of osteoarthritis? Rheum Dis Clin North Am 19:617 633, 1993.

18. Kraus VB. Pathogenesis and treatment of osteoarthritis. Med Clin North Am 81:85 112, 1997.

19. Griffin TM, Guilak F. The role of mechanical loading in the onset and progression of osteoarthritis. Exerc Sport Sci Rev 33: 195-200, 2005.

20. Lane JM, Chisena E, Black J. Experimental knee instability: Early mechanical property changes in articular cartilage in a rabbit model. Clin Orthop Rel Res 140:262 265, 1979.

21. Howell DS, Treadwell BV, Trippel SB. Etiopathogenesis of osteoarthritis. In: Moskowitz RW, Howell DS, Goldberg VM, et al., eds. Osteoarthritis, Diagnosis and Medical/Surgical Management. 2nd ed. Philadelphia: W.B. Saunders; pp 233-252, 1992.

22. Pond MJ, Nuki G. Experimentally induced osteoarthritis in the dog. Ann Rheum Dis 32:387 388, 1973.

23. Moskowitz RW, Davis W, Sammarco J. Experimentally induced degenerative joint lesions following partial meniscectomy in the rabbit. Arthritis Rheum 16:397 405, 1973.

24. McDevitt C, Gilbertson E, Muir H. An experimental model of osteoarthritis; early morphological and biochemical changes. J Bone Joint Surg 59B:24 35, 1977.

25. Eyre DR, McDevitt CA, Billingham ME, et al. Biosynthesis of collagen and other matrix proteins by articular cartilage in experimental osteoarthrosis. Biochem J 188:823 837, 1980.

26. Altman RD, Tenenbaum J, Latta L, et al. Biomechanical and biochemical properties of dog cartilage in experimentally induced osteoarthritis. Ann Rheum Dis 43:83 90, 1984.

27. Sandy JD, Adams ME, Billingham ME, et al. In vivo and in vitro stimulation of chondrocyte biosynthetic activity in early experimental osteoarthritis. Arthritis Rheum 27:388 397, 1984.

28. Guilak F, Ratcliffe A, Lane N, et al. Mechanical and biochemical changes in the superficial zone of articular cartilage in canine experimental osteoarthritis. J Orthop Res 12:474 484, 1994.

29. Messier SP. Osteoarthritis of the knee and associated factors of age and obesity: effects on gait. Med Sci Sports Exerc 26: 1446-1452, 1994.

30. Das UN. Obesity, metabolic syndrome X, and inflammation. Nutrition 18:430 432, 2002.

31. Powell A, Teichtahl AJ, Wluka AE, et al. Obesity: a preventable risk factor for large joint osteoarthritis which may act through biomechanical factors. Br J Sports Med 39:4 5, 2005.

32. Hutton CW. Generalised osteoarthritis: an evolutionary problem? Lancet 1:1463 1465, 1987.

33. O'Connor BL, Palmoski MJ, Brandt KD. Neurogenic acceleration of degenerative joint lesions. J Bone Joint Surg Am 67:562 572, 1985.

34. Yang KH, Riegger CL, Rodgers MM. Diminished lower limb deceleration as a factor in early stage osteoarthrosis. Trans Orthop Res Soc 14:52, 1989.

35. O'Reilly SC, Jones A, Muir KR, et al. Quadriceps weakness in knee osteoarthritis: the effect on pain and disability. Ann Rheum Dis 57:588 594, 1998.

36. McAlindon T, Felson DT. Nutrition: risk factors for osteoarthritis. Ann Rheum Dis 56:397 400, 1997.

37. Sokoloff L. The history of Kashin-Beck disease. N Y State J Med 89:343 351, 1989.

38. Utiger RD. Kashin-Beck disease expanding the spectrum of iodine-deficiency disorders. N Engl J Med 339:1156 1158, 1998.

39. Aurich M, Squires GR, Reiner A, et al. Differential matrix degradation and turnover in early cartilage lesions of human knee and ankle joints. Arthritis Rheum 52:112 119, 2005.

40. Hardingham TE, Fosang AJ, Dudhia J. The structure, function and turnover of aggrecan, the large aggregating proteoglycan from cartilage. Eur J Clin Chem Clin Biochem 32:249 257, 1994.

41. Maroudas A. Physicochemical properties of articular cartilage. In: Freeman M, ed. Adult Articular Cartilage. Tunbridge Wells: Pitman Medical; 1979, pp 215-290.

42. Hodge W, Fijan R, Carlson K, et al. Contact pressures in the human hip joint measured in vivo. Proc Natl Acad Sci USA 83:2879 2883, 1986.

43. Mow VC, Kuei SC, Lai WM, et al. Biphasic creep and stress relaxation of articular cartilage in compression: Theory and experiments. J Biomech Eng 102:73 84, 1980.

44. Frank EH, Grodzinsky AJ. Cartilage electromechanics I. Electrokinetic transduction and the effects of electrolyte pH and ionic strength. J Biomech Eng 20:615 627, 1987.

45. Lai WM, Hou JS, Mow VC. A triphasic theory for the swelling and deformation behaviors of articular cartilage. J Biomech Eng 113:245 258, 1991.

46. Kempson GE, Muir H, Pollard C, et al. The tensile properties of the cartilage of human femoral condyles related to the content of collagen and glycosaminoglycans. Biochim Biophys Acta 297: 456-472, 1973.

47. Roth V, Mow VC. The intrinsic tensile behavior of the matrix of bovine articular cartilage and its variation with age. J Bone Joint Surg 62:1102 1117, 1980.

48. Kempson GE. Age-related changes in the tensile properties of human articular cartilage: a comparative study between the femoral head of the hip joint and the talus of the ankle joint. Biochim Biophys Acta 1075:223 230, 1991.

49. Chen AC, Temple MM, Ng DM, et al. Induction of advanced glycation end products and alterations of the tensile properties of articular cartilage. Arthritis Rheum 46:3212 3217, 2002.

50. Poole AR. Cartilage in health and disease. In: Koopman WJ, Moreland LW, eds. Arthritis and Allied Conditions. A Textbook of Rheumatology. 15th ed. Baltimore: Williams & Wilkins; 2005, pp 223-269.

51. Setton LA, Tohyama H, Mow VC. Swelling and curling behaviors of articular cartilage. J Biomech Eng 120:355 361, 1998.

52. Narmoneva DA, Wang JY, Setton LA. Nonuniform swelling-induced residual strains in articular cartilage. J Biomech 32: 401-408, 1999.

53. Flahiff CM, Kraus VB, Huebner JL, et al. Cartilage mechanics in the guinea pig model of osteoarthritis studied with an osmotic loading method. Osteoarthritis Cartilage 12:383 388, 2004.

54. Alexopoulos LG, Setton LA, Guilak F. The biomechanical role of the chondrocyte pericellular matrix in articular cartilage. Acta Biomater 1:317 325, 2005.

55. Malcom LL. An experimental investigation of the frictional and deformational responses of articular cartilage interfaces to static and dynamic loading [dissertation]. San Diego, University of California, San Diego, 1976.

56. Mow VC, Ateshian GA. Lubrication and wear of diarthrodial joints. In: Mow VC, Hayes WC, eds. Basic Orthopaedic Biomechanics, 2nd Ed. Philadelphia: Lippincott-Raven; 1997, pp 275-315.

P.43


57. Park S, Krishnan R, Nicoll SB, et al. Cartilage interstitial fluid load support in unconfined compression. J Biomech 36: 1785-1796, 2003.

58. McCutchen CW. Joint lubrication. Bull Hosp Jt Dis Orthop Inst 43:118 129, 1983.

59. Hou JS, Mow VC, Lai WM, et al. An analysis of the squeeze-film lubrication mechanism for articular cartilage. J Biomech 25: 247-259, 1992.

60. Swann DA, Slayter HS, Silver FH. The molecular structure of lubricating glycoprotein-I, the boundary lubricant for articular cartilage. J Biol Chem 256:5921 5925, 1981.

61. Flannery CR, Hughes CE, Schumacher BL, et al. Articular cartilage superficial zone protein (SZP) is homologous to megakaryocyte stimulating factor precursor and is a multifunctional proteoglycan with potential growth-promoting, cytoprotective, and lubricating properties in cartilage metabolism. Biochem Biophys Res Commun 254:535 541, 1999.

62. Jay GD, Tantravahi U, Britt DE, et al. Homology of lubricin and superficial zone protein (SZP): products of megakaryocyte stimulating factor (MSF) gene expression by human synovial fibroblasts and articular chondrocytes localized to chromosome 1q25. J Orthop Res 19:677 687, 2001.

63. Marcelino J, Carpten JD, Suwairi WM, et al. CACP, encoding a secreted proteoglycan, is mutated in camptodactyly-arthropathy-coxa vara-pericarditis syndrome. Nat Genet 23:319 322, 1999.

64. Guilak F, Sah RL, Setton LA. Physical regulation of cartilage metabolism. In: Mow VC, Hayes WC, eds. Basic Orthopaedic Biomechanics. 2nd ed. Philadelphia: Lippincott-Raven, 1997, 179-207.

65. Gray ML, Pizzanelli AM, Grodzinsky AJ, et al. Mechanical and physiochemical determinants of the chondrocyte biosynthetic response. J Orthop Res 6:777 792, 1988.

66. Sah RL, Doong JY, Grodzinsky AJ, et al. Effects of compression on the loss of newly synthesized proteoglycans and proteins from cartilage explants. Arch Biochem Biophys 286:20 29, 1991.

67. Guilak F, Meyer BC, Ratcliffe A, et al. The effects of matrix compression on proteoglycan metabolism in articular cartilage explants. Osteoarthritis Cartilage 2:91 101, 1994.

68. Torzilli PA, Grigiene R, Huang C, et al. Characterization of cartilage metabolic response to static and dynamic stress using a mechanical explant test system. J Biomech 30:1 9, 1997.

69. Farquhar T, Xia Y, Mann K, et al. Swelling and fibronectin accumulation in articular cartilage explants after cyclical impact. J Orthop Res 14:417 423, 1996.

70. Chen C-T, Burton-Wurster N, Borden C, et al. Chondrocyte necrosis and apoptosis in impact damaged articular cartilage. J Orthop Res 19:703 711, 2001.

71. Loening AM, James IE, Levenston ME, et al. Injurious mechanical compression of bovine articular cartilage induces chondrocyte apoptosis. Arch Biochem Biophys 381:205 212, 2000.

72. Stockwell RA. Structure and function of the chondrocyte under mechanical stress. In: Helminen HJ, Kiviranta I, Tammi M, et al., eds. Joint Loading: Biology and Health of Articular Structures. Bristol: Wright and Sons, 1987, pp 126-148.

73. Squires GR, Okouneff S, Ionescu M, et al. The pathobiology of focal lesion development in aging human articular cartilage and molecular matrix changes characteristic of osteoarthritis. Arthritis Rheum 48:1261 1270, 2003.

74. Lohmander LS. Markers of cartilage metabolism in arthrosis. A review. Acta Orthop Scand 62:623 632, 1991.

75. Nelson F, Billinghurst RC, Pidoux I, et al. Early post-traumatic osteoarthritis-like changes in human articular cartilage following rupture of the anterior cruciate ligament. Osteoarthritis Cartilage 14:114 119, 2006.

76. Harris AM, Patterson BM, Sontich JK, et al. Results and outcomes after operative treatment of high-energy tibial plafond fractures. Foot Ankle Int 27:256 265, 2006.

77. Buckwalter JA, Brown TD. Joint injury, repair, and remodeling: roles in post-traumatic osteoarthritis. Clin Orthop Relat Res 7-16, 2004.

78. Guilak F, Fermor B, Keefe FJ, et al. The role of biomechanics and inflammation in cartilage injury and repair. Clin Orthop Relat Res 17-26, 2004.

79. Olson SA, Marsh JL. Posttraumatic osteoarthritis. Clin Orthop Relat Res 2, 2004.

80. Bleasel JF, Poole AR, Heinegard D, et al. Changes in serum cartilage marker levels indicate altered cartilage metabolism in families with the osteoarthritis-related type II collagen gene COL2A1 mutation. Arthritis Rheum 42:39 45, 1999.

81. Matyas JR, Atley L, Ionescu M, et al. Analysis of cartilage biomarkers in the early phases of canine experimental osteoarthritis. Arthritis Rheum 50:543 552, 2004.

82. Tiraloche G, Girard C, Chouinard L, et al. Effect of oral glucosamine on cartilage degradation in a rabbit model of osteoarthritis. Arthritis Rheum 52:1118 1128, 2005.

83. Adams ME, Brandt KD. Hypertrophic repair of canine articular cartilage in osteoarthritis after anterior cruciate ligament transection. J Rheumatol 18:428 435, 1991.

84. Brandt KD, Braunstein EM, Visco DM, et al. Anterior (cranial) cruciate ligament transection in the dog: a bona fide model of osteoarthritis, not merely of cartilage injury and repair. J Rheumatol 18:436 446, 1991.

85. Manicourt DH, Altman RD, Williams JM, et al. Treatment with calcitonin suppresses the responses of bone, cartilage, and synovium in the early stages of canine experimental osteoarthritis and significantly reduces the severity of the cartilage lesions. Arthritis Rheum 42:1159 1167, 1999.

86. Manicourt DH, Thonar EJ, Pita JC, et al. Changes in the sedimentation profile of proteoglycan aggregates in early experimental canine osteoarthritis. Connect Tissue Res 23:33 50, 1989.

87. Pelletier JP, Martel-Pelletier J, Altman RD, et al. Collagenolytic activity and collagen matrix breakdown of the articular cartilage in the Pond-Nuki dog model of osteoarthritis. Arthritis Rheum 26:866 874, 1983.

88. Mitchell PG, Magna HA, Reeves LM, et al. Cloning, expression, and type II collagenolytic activity of matrix metalloproteinase-13 from human osteoarthritic cartilage. J Clin Invest 97:761 768, 1996.

89. Billinghurst RC, Dahlberg L, Ionescu M, et al. Enhanced cleavage of type II collagen by collagenases in osteoarthritic articular cartilage. J Clin Invest 99:1534 1545, 1997.

90. Dahlberg L, Billinghurst RC, Manner P, et al. Selective enhancement of collagenase-mediated cleavage of resident type II collagen in cultured osteoarthritic cartilage and arrest with a synthetic inhibitor that spares collagenase 1 (matrix metalloproteinase 1). Arthritis Rheum 43:673 682, 2000.

91. Neuhold LA, Killar L, Zhao W, et al. Postnatal expression in hyaline cartilage of constitutively active human collagenase-3 (MMP-13) induces osteoarthritis in mice. J Clin Invest 107: 35-44, 2001.

92. Poole AR, Alini M, Hollander AP. Cellular biology of cartilage degradation. In: Henderson B, Edwards JCW, Pettipher ER, eds. Mechanisms and Models in Rheumatoid Arthritis. London: Academic Press, 1995, pp 163-204.

93. Kafienah W, Bromme D, Buttle DJ, et al. Human cathepsin K cleaves native type I and II collagens at the N-terminal end of the triple helix. Biochem J 331 (Pt 3):727 732, 1998.

94. Konttinen YT, Mandelin J, Li TF, et al. Acidic cysteine endoproteinase cathepsin K in the degeneration of the superficial articular hyaline cartilage in osteoarthritis. Arthritis Rheum 46: 953-960, 2002.

95. Kevorkian L, Young DA, Darrah C, et al. Expression profiling of metalloproteinases and their inhibitors in cartilage. Arthritis Rheum 50:131 141, 2004.

96. Bau B, Gebhard PM, Haag J, et al. Relative messenger RNA expression profiling of collagenases and aggrecanases in human articular chondrocytes in vivo and in vitro. Arthritis Rheum 46: 2648-2657, 2002.

97. Wu CW, Tchetina EV, Mwale F, et al. Proteolysis involving matrix metalloproteinase 13 (collagenase-3) is required for chondrocyte differentiation that is associated with matrix mineralization. J Bone Miner Res 17:639 651, 2002.

98. Tchetina EV, Squires G, Poole AR. Increased type II collagen degradation and very early focal cartilage degeneration is associated with upregulation of chondrocyte differentiation related genes in early human articular cartilage lesions. J Rheumatol 32:876 886, 2005.

99. Shlopov BV, Gumanovskaya ML, Hasty KA. Autocrine regulation of collagenase 3 (matrix metalloproteinase 13) during osteoarthritis. Arthritis Rheum 43:195 205, 2000.

P.44


100. Cole AA, Kuettner KE. MMP-8 (neutrophil collagenase) mRNA and aggrecanase cleavage products are present in normal and osteoarthritic human articular cartilage. Acta Orthop Scand Suppl 266:98 102, 1995.

101. Aigner T, Zien A, Gehrsitz A, et al. Anabolic and catabolic gene expression pattern analysis in normal versus osteoarthritic cartilage using complementary DNA-array technology. Arthritis Rheum 44:2777 2789, 2001.

102. Imai K, Ohta S, Matsumoto T, et al. Expression of membrane-type 1 matrix metalloproteinase and activation of progelatinase A in human osteoarthritic cartilage. Am J Pathol 151:245 256, 1997.

103. Buttner FH, Chubinskaya S, Margerie D, et al. Expression of membrane type 1 matrix metalloproteinase in human articular cartilage. Arthritis Rheum 40:704 709, 1997.

104. Ohta S, Imai K, Yamashita K, et al. Expression of matrix metalloproteinase 7 (matrilysin) in human osteoarthritic cartilage. Lab Invest 78:79 87, 1998.

105. Martel-Pelletier J, Faure MP, McCollum R, et al. Plasmin, plasminogen activators and inhibitor in human osteoarthritic cartilage. J Rheumatol 18:1863 1871, 1991.

106. Baici A, Horler D, Lang A, et al. Cathepsin B in osteoarthritis: zonal variation of enzyme activity in human femoral head cartilage. Ann Rheum Dis 54:281 288, 1995.

107. Mehraban F, Tindal MH, Proffitt MM, et al. Temporal pattern of cysteine endopeptidase (cathepsin B) expression in cartilage and synovium from rabbit knees with experimental osteoarthritis: gene expression in chondrocytes in response to interleukin-1 and matrix depletion. Ann Rheum Dis 56:108 115, 1997.

108. Hollander AP, Heathfield TF, Webber C, et al. Increased damage to type II collagen in osteoarthritic articular cartilage detected by a new immunoassay. J Clin Invest 93:1722 1732, 1994.

109. Hollander AP, Pidoux I, Reiner A, et al. Damage to type II collagen in aging and osteoarthritis starts at the articular surface, originates around chondrocytes, and extends into the cartilage with progressive degeneration. J Clin Invest 96:2859 2869, 1995.

110. Poole AR. Biochemical/immunochemical biomarkers of osteoarthritis: utility for prediction of incident or progressive osteoarthritis. Rheum Dis Clin North Am 29:803 818, 2003.

111. Price JS, Till SH, Bickerstaff DR, et al. Degradation of cartilage type II collagen precedes the onset of osteoarthritis following anterior cruciate ligament rupture. Arthritis Rheum 42: 2390-2398, 1999.

112. Aigner T, Bertling W, Stoss H, et al. Independent expression of fibril-forming collagens I, II, and III in chondrocytes of human osteoarthritic cartilage. J Clin Invest 91:829 837, 1993.

113. Matyas JR, Adams ME, Huang D, et al. Major role of collagen IIB in the elevation of total type II procollagen messenger RNA in the hypertrophic phase of experimental osteoarthritis. Arthritis Rheum 40:1046 1049, 1997.

114. Nelson F, Dahlberg L, Laverty S, et al. Evidence for altered synthesis of type II collagen in patients with osteoarthritis. J Clin Invest 102:2115 2125, 1998.

115. Rizkalla G, Reiner A, Bogoch E, et al. Studies of the articular cartilage proteoglycan aggrecan in health and osteoarthritis. Evidence for molecular heterogeneity and extensive molecular changes in disease. J Clin Invest 90:2268 2277, 1992.

116. Tortorella MD, Burn TC, Pratta MA, et al. Purification and cloning of aggrecanase-1: a member of the ADAMTS family of proteins. Science 284:1664 1666, 1999.

117. Lark MW, Bayne EK, Flanagan J, et al. Aggrecan degradation in human cartilage. Evidence for both matrix metalloproteinase and aggrecanase activity in normal, osteoarthritic, and rheumatoid joints. J Clin Invest 100:93 106, 1997.

118. Glasson SS, Askew R, Sheppard B, et al. Deletion of active ADAMTS5 prevents cartilage degradation in a murine model of osteoarthritis. Nature 434:644 648, 2005.

119. Stanton H, Rogerson FM, East CJ, et al. ADAMTS5 is the major aggrecanase in mouse cartilage in vivo and in vitro. Nature 434:648 652, 2005.

120. Gao G, Plaas A, Thompson VP, et al. ADAMTS4 (aggrecanase-1) activation on the cell surface involves C-terminal cleavage by glycosylphosphatidyl inositol-anchored membrane type 4-matrix metalloproteinase and binding of the activated proteinase to chondroitin sulfate and heparan sulfate on syndecan-1. J Biol Chem 279:10042 10051, 2004.

121. Cs-Szabo G, Roughley PJ, Plaas AH, et al. Large and small proteoglycans of osteoarthritic and rheumatoid articular cartilage. Arthritis Rheum 38:660 668, 1995.

122. Visco DM, Johnstone B, Hill MA, et al. Immunohistochemical analysis of 3-B-(-) and 7-D-4 epitope expression in canine osteoarthritis. Arthritis Rheum 36:1718 1725, 1993.

123. Carlson CS, Loeser RF, Johnstone B, et al. Osteoarthritis in cynomolgus macaques. II. Detection of modulated proteoglycan epitopes in cartilage and synovial fluid. J Orthop Res 13: 399-409, 1995.

124. Lohmander LS, Ionescu M, Jugessur H, et al. Changes in joint cartilage aggrecan after knee injury and in osteoarthritis. Arthritis Rheum 42:534 544, 1999.

125. Poole AR, Ionescu M, Swan A, et al. Changes in cartilage metabolism in arthritis are reflected by altered serum and synovial fluid levels of the cartilage proteoglycan aggrecan. Implications for pathogenesis. J Clin Invest 94:25 33, 1994.

126. Poole AR. Skeletal and inflammation markers in aging and osteoarthritis. In: Hamerman D, ed. Osteoarthritis and the Ageing Population. Baltimore: Johns Hopkins University Press, 1997, pp 187-214.

127. Homandberg GA, Meyers R, Xie DL. Fibronectin fragments cause chondrolysis of bovine articular cartilage slices in culture. J Biol Chem 267:3597 3604, 1992.

128. Yasuda T, Poole AR. A fibronectin fragment induces type II collagen degradation by collagenase through an interleukin-1-mediated pathway. Arthritis Rheum 46:138 148, 2002.

129. Yasuda T, Poole AR, Shimizu M, et al. Involvement of CD44 in induction of matrix metalloproteinases by a COOH-terminal heparin-binding fragment of fibronectin in human articular cartilage in culture. Arthritis Rheum 48:1271 1280, 2003.

130. Werb Z, Tremble PM, Behrendtsen O, et al. Signal transduction through the fibronectin receptor induces collagenase and stromelysin gene expression. J Cell Biol 109:877 889, 1989.

131. Chevalier X, Claudepierre P, Groult N, et al. Presence of ED-A containing fibronectin in human articular cartilage from patients with osteoarthritis and rheumatoid arthritis. J Rheumatol 23:1022 1030, 1996.

132. Arner EC, Tortorella MD. Signal transduction through chondrocyte integrin receptors induces matrix metalloproteinase synthesis and synergizes with interleukin-1. Arthritis Rheum 38: 1304-1314, 1995.

133. Yasuda T, Tchetina E, Ohsawa K, et al. Peptides of type II collagen can induce the cleavage of type II collagen and aggrecan in articular cartilage. Matrix Biol, In press, 2006.

134. Fichter M, Korner U, Schomburg J, et al. Collagen degradation products modulate matrix metalloproteinase expression in cultured articular chondrocytes. J Orthop Res 24:63 70, 2006.

135. Martel-Pelletier J, McCollum R, DiBattista J, et al. The interleukin-1 receptor in normal and osteoarthritic human articular chondrocytes. Identification as the type I receptor and analysis of binding kinetics and biologic function. Arthritis Rheum 35:530 540, 1992.

136. Melchiorri C, Meliconi R, Frizziero L, et al. Enhanced and coordinated in vivo expression of inflammatory cytokines and nitric oxide synthase by chondrocytes from patients with osteoarthritis. Arthritis Rheum 41:2165 2174, 1998.

137. Webb GR, Westacott CI, Elson CJ. Chondrocyte tumor necrosis factor receptors and focal loss of cartilage in osteoarthritis. Osteoarthritis Cartilage 5:427 437, 1997.

138. Cawston TE, Curry VA, Summers CA, et al. The role of oncostatin M in animal and human connective tissue collagen turnover and its localization within the rheumatoid joint. Arthritis Rheum 41:1760 1771, 1998.

139. Kobayashi M, Squires GR, Mousa A, et al. Role of interleukin-1 and tumor necrosis factor alpha in matrix degradation of human osteoarthritic cartilage. Arthritis Rheum 52:128 135, 2005.

140. Amin AR, Di Cesare PE, Vyas P, et al. The expression and regulation of nitric oxide synthase in human osteoarthritis-affected chondrocytes: evidence for up-regulated neuronal nitric oxide synthase. J Exp Med 182:2097 2102, 1995.

P.45


141. Palmer RM, Hickery MS, Charles IG, et al. Induction of nitric oxide synthase in human chondrocytes. Biochem Biophys Res Commun 193:398 405, 1993.

142. Attur MG, Dave MN, Stuchin S, et al. Osteopontin: an intrinsic inhibitor of inflammation in cartilage. Arthritis Rheum 44: 578-584, 2001.

143. Nakamura S, Kamihagi K, Satakeda H, et al. Enhancement of SPARC (osteonectin) synthesis in arthritic cartilage. Increased levels in synovial fluids from patients with rheumatoid arthritis and regulation by growth factors and cytokines in chondrocyte cultures. Arthritis Rheum 39:539 551, 1996.

144. Taskiran D, Stefanovic-Racic M, Georgescu H, et al. Nitric oxide mediates suppression of cartilage proteoglycan synthesis by interleukin-1. Biochem Biophys Res Commun 200:142 148, 1994.

145. Stefanovic-Racic M, Morales TI, Taskiran D, et al. The role of nitric oxide in proteoglycan turnover by bovine articular cartilage organ cultures. J Immunol 156:1213 1220, 1996.

146. Clements KM, Price JS, Chambers MG, et al. Gene deletion of either interleukin-1beta, interleukin-1beta-converting enzyme, inducible nitric oxide synthase, or stromelysin 1 accelerates the development of knee osteoarthritis in mice after surgical transection of the medial collateral ligament and partial medial meniscectomy. Arthritis Rheum 48:3452 3463, 2003.

147. Blanco FJ, Ochs RL, Schwarz H, et al. Chondrocyte apoptosis induced by nitric oxide. Am J Pathol 146:75 85, 1995.

148. Pelletier JP, Jovanovic DV, Lascau-Coman V, et al. Selective inhibition of inducible nitric oxide synthase reduces progression of experimental osteoarthritis in vivo: possible link with the reduction in chondrocyte apoptosis and caspase 3 level. Arthritis Rheum 43:1290 1299, 2000.

149. Fermor B, Weinberg JB, Pisetsky DS, et al. The influence of oxygen tension on the induction of nitric oxide and prostaglandin E2 by mechanical stress in articular cartilage. Osteoarthritis Cartilage 13:935 941, 2005.

150. Fermor B, Weinberg JB, Pisetsky DS, et al. Induction of cyclooxygenase-2 by mechanical stress through a nitric oxide-regulated pathway. Osteoarthritis Cartilage 10:792 798, 2002.

151. Lee JH, Fitzgerald JB, Dimicco MA, et al. Mechanical injury of cartilage explants causes specific time-dependent changes in chondrocyte gene expression. Arthritis Rheum 52:2386 2395, 2005.

152. Patwari P, Cook MN, DiMicco MA, et al. Proteoglycan degradation after injurious compression of bovine and human articular cartilage in vitro: interaction with exogenous cytokines. Arthritis Rheum 48:1292 1301, 2003.

153. Piscoya JL, Fermor B, Kraus VB, et al. The influence of mechanical compression on the induction of osteoarthritis-related biomarkers in articular cartilage explants. Osteoarthritis Cartilage 13: 1092-1099, 2005.

154. Dean DD, Martel-Pelletier J, Pelletier JP, et al. Evidence for metalloproteinase and metalloproteinase inhibitor imbalance in human osteoarthritic cartilage. J Clin Invest 84:678 685, 1989.

155. Yamada H, Nakagawa T, Stephens RW, et al. Proteinases and their inhibitors in normal and osteoarthritic articular cartilage. Biomed Res 8:289 300, 1987.

156. Gendron C, Kashiwagi M, Hughes C, et al. TIMP-3 inhibits aggrecanase-mediated glycosaminoglycan release from cartilage explants stimulated by catabolic factors. FEBS Lett 555:431 436, 2003.

157. Moulharat N, Lesur C, Thomas M, et al. Effects of transforming growth factor-beta on aggrecanase production and proteoglycan degradation by human chondrocytes in vitro. Osteoarthritis Cartilage 12:296 305, 2004.

158. Tchetina EV, Antoniou J, Tanzer M, et al. Transforming growth factor-beta2 suppresses collagen cleavage in cultured human osteoarthritic cartilage, reduces expression of genes associated with chondrocyte hypertrophy and degradation, and increases prostaglandin E(2) production. Am J Pathol 168:131 140, 2006.

159. Ling H, Recklies AD. The Chitinase 3-like protein human cartilage glycoprotein 39 inhibits cellular responses to the inflammatory cytokines interleukin-1 and tumour necrosis factor-alpha. Biochem J 380:651 659, 2004.

160. Poole AR, Rosenberg LC, Reiner A, et al. Contents and distributions of the proteoglycans decorin and biglycan in normal and osteoarthritic human articular cartilage. J Orthop Res 14: 681-689, 1996.

161. Lefkoe TP, Nalin AM, Clark JM, et al. Gene expression of collagen types IIA and IX correlates with ultrastructural events in early osteoarthrosis: new applications of the rabbit meniscectomy model. J Rheumatol 24:1155 1163, 1997.

162. Aigner T, Zhu Y, Chansky HH, et al. Reexpression of type IIA procollagen by adult articular chondrocytes in osteoarthritic cartilage. Arthritis Rheum 42:1443 1450, 1999.

163. Sandell LJ, Morris N, Robbins JR, et al. Alternatively spliced type II procollagen mRNAs define distinct populations of cells during vertebral development: differential expression of the amino-propeptide. J Cell Biol 114:1307 1319, 1991.

164. McDevitt CA, Pahl JA, Ayad S, et al. Experimental osteoarthritic articular cartilage is enriched in guanidine soluble type VI collagen. Biochem Biophys Res Commun 157:250 255, 1988.

165. Ronziere MC, Ricard-Blum S, Tiollier J, et al. Comparative analysis of collagens solubilized from human foetal, and normal and osteoarthritic adult articular cartilage, with emphasis on type VI collagen. Biochim Biophys Acta. 1038:222 230, 1990.

166. Hambach L, Neureiter D, Zeiler G, et al. Severe disturbance of the distribution and expression of type VI collagen chains in osteoarthritic articular cartilage. Arthritis Rheum 41:986 996, 1998.

167. Soder S, Hambach L, Lissner R, et al. Ultrastructural localization of type VI collagen in normal adult and osteoarthritic human articular cartilage. Osteoarthritis Cartilage 10:464 470, 2002.

168. Lee GM, Paul TA, Slabaugh M, et al. The pericellular matrix is enlarged in osteoarthritic cartilage. Trans Orthop Res Soc 22:495, 1997.

169. Alexopoulos LG, Williams GM, Upton ML, et al. Osteoarthritic changes in the biphasic mechanical properties of the chondrocyte pericellular matrix in articular cartilage. J Biomech 38: 509-517, 2005.

170. Alexopoulos LG, Haider MA, Vail TP, et al. Alterations in the mechanical properties of the human chondrocyte pericellular matrix with osteoarthritis. J Biomech Eng 125:323 333, 2003.

171. Cs-Szabo G, Melching LI, Roughley PJ, et al. Changes in messenger RNA and protein levels of proteoglycans and link protein in human osteoarthritic cartilage samples. Arthritis Rheum 40: 1037-1045, 1997.

172. Nishida Y, Shinomura T, Iwata H, et al. Abnormal occurrence of a large chondroitin sulfate proteoglycan, PG-M/versican in osteoarthritic cartilage. Osteoarthritis Cartilage 2:43 49, 1994.

173. Schneiderman R, Rosenberg N, Hiss J, et al. Concentration and size distribution of insulin-like growth factor-I in human normal and osteoarthritic synovial fluid and cartilage. Arch Biochem Biophys 324:173 188, 1995.

174. Dore S, Pelletier JP, DiBattista JA, et al. Human osteoarthritic chondrocytes possess an increased number of insulin-like growth factor 1 binding sites but are unresponsive to its stimulation. Possible role of IGF-1-binding proteins. Arthritis Rheum 37: 253-263, 1994.

175. Chen AL, Fang C, Liu C, et al. Expression of bone morphogenetic proteins, receptors, and tissue inhibitors in human fetal, adult, and osteoarthritic articular cartilage. J Orthop Res 22:1188 1192, 2004.

176. Di Cesare PE, Carlson CS, Stolerman ES, et al. Increased degradation and altered tissue distribution of cartilage oligomeric matrix protein in human rheumatoid and osteoarthritic cartilage. J Orthop Res 14:946 955, 1996.

177. Okimura A, Okada Y, Makihira S, et al. Enhancement of cartilage matrix protein synthesis in arthritic cartilage. Arthritis Rheum 40:1029 1036, 1997.

178. Chevalier X, Groult N, Larget-Piet B, et al. Tenascin distribution in articular cartilage from normal subjects and from patients with osteoarthritis and rheumatoid arthritis. Arthritis Rheum 37:1013 1022, 1994.

179. Recklies AD, White C, Ling H. The Chitinase 3-like protein human cartilage glycoprotein 39 (HC-gp39) stimulates proliferation of human connective-tissue cells and activates both extracellular signal-regulated kinase- and protein kinase B-mediated signalling pathways. Biochem J 365:119 126, 2002.

P.46


180. Marshall KW, Zhang H, Yager TD, et al. Blood-based biomarkers for detecting mild osteoarthritis in the human knee. Osteoarthritis Cartilage 13:861 871, 2005.

181. van den Berg WB, van der Kraan PM, van Beuningen HM. Synovial mediators of cartilage damage and repair in OA. In: Brandt KD, Doherty M, Lohmander LS, eds. Osteoarthritis. Oxford: Oxford University Press, 1998.

182. Goldring MB. Osteoarthritis and cartilage: the role of cytokines. Curr Rheumatol Rep 2:459 465, 2000.

183. Morales T, I. The role of signaling factors in articular cartilage homeostasis and osteoarthritis. In: Kuettner KE, Goldberg VM, eds. Osteoarthritic Disorders. Rosemont, IL: American Academy of Orthopaedic Surgeons, 1995, pp 261-270.

184. Middleton JF, Tyler JA. Upregulation of insulin-like growth factor I gene expression in the lesions of osteoarthritic human articular cartilage. Ann Rheum Dis 51:440 447, 1992.

185. Tyler JA. Insulin-like growth factor 1 can decrease degradation and promote synthesis of proteoglycan in cartilage exposed to cytokines. Biochem J 260:543 548, 1989.

186. Trippel SB, Corvol MT, Dumontier MF, et al. Effect of somatomedin-C/insulin-like growth factor I and growth hormone on cultured growth plate and articular chondrocytes. Pediatr Res 25:76 82, 1989.

187. Martel-Pelletier J, Di Battista JA, Lajeunesse D, et al. IGF/IGFBP axis in cartilage and bone in osteoarthritis pathogenesis. Inflamm Res 47:90 100, 1998.

188. Fowlkes JL, Serra DM, Rosenberg CK, et al. Insulin-like growth factor (IGF)-binding protein-3 (IGFBP-3) functions as an IGF-reversible inhibitor of IGFBP-4 proteolysis. J Biol Chem 270: 27481-27488, 1995.

189. Sachs BL, Goldberg VM, Moskowitz RW, et al. Response of articular chondrocytes to pituitary fibroblast growth factor (FGF). J Cell Physiol 112:51 59, 1982.

190. Kato Y, Gospodarowicz D. Sulfated proteoglycan synthesis by confluent cultures of rabbit costal chondrocytes grown in the presence of fibroblast growth factor. J Cell Biol 100:477 485, 1985.

191. Campbell IK, Wojta J, Novak U, et al. Cytokine modulation of plasminogen activator inhibitor-1 (PAI-1) production by human articular cartilage and chondrocytes. Down-regulation by tumor necrosis factor alpha and up-regulation by transforming growth factor-B basic fibroblast growth factor. Biochim Biophys Acta 1226:277 285, 1994.

192. Bakker AC, van de Loo FA, van Beuningen HM, et al. Overexpression of active TGF-beta-1 in the murine knee joint: evidence for synovial-layer-dependent chondro-osteophyte formation. Osteoarthritis Cartilage 9:128 136, 2001.

193. van Beuningen HM, Glansbeek HL, van der Kraan PM, et al. Osteoarthritis-like changes in the murine knee joint resulting from intra-articular transforming growth factor-beta injections. Osteoarthritis Cartilage 8:25 33, 2000.

194. Huebner JL, Kraus VB. Assessment of the utility of biomarkers of osteoarthritis in the guinea pig. Osteoarthritis Cartilage 2006.

195. von der Mark K, Kirsch T, Nerlich A, et al. Type X collagen synthesis in human osteoarthritic cartilage. Indication of chondrocyte hypertrophy. Arthritis Rheum 35:806 811, 1992.

196. Mwale F, Tchetina E, Wu CW, et al. The assembly and remodeling of the extracellular matrix in the growth plate in relationship to mineral deposition and cellular hypertrophy: an in situ study of collagens II and IX and proteoglycan. J Bone Miner Res 17:275 283, 2002.

197. Hashimoto S, Ochs RL, Komiya S, et al. Linkage of chondrocyte apoptosis and cartilage degradation in human osteoarthritis. Arthritis Rheum 41:1632 1638, 1998.

198. Blanco FJ, Guitian R, Vazquez-Martul E, et al. Osteoarthritis chondrocytes die by apoptosis. A possible pathway for osteoarthritis pathology. Arthritis Rheum 41:284 289, 1998.

199. Gebhard PM, Gehrsitz A, Bau B, et al. Quantification of expression levels of cellular differentiation markers does not support a general shift in the cellular phenotype of osteoarthritic chondrocytes. J Orthop Res 21:96 101, 2003.

200. Sharif M, Whitehouse A, Sharman P, et al. Increased apoptosis in human osteoarthritic cartilage corresponds to reduced cell density and expression of caspase-3. Arthritis Rheum 50:507 515, 2004.

201. Mollenhauer J, Mok MT, King KB, et al. Expression of anchorin CII (cartilage annexin V) in human young, normal adult, and osteoarthritic cartilage. J Histochem Cytochem 47:209 220, 1999.

202. Terkeltaub R, Lotz M, Johnson K, et al. Parathyroid hormone-related protein is abundant in osteoarthritic cartilage, and the parathyroid hormone-related protein 1-173 isoform is selectively induced by transforming growth factor beta in articular chondrocytes and suppresses generation of extracellular inorganic pyrophosphate. Arthritis Rheum 41:2152 2164, 1998.

203. Okano K, Tsukazaki T, Ohtsuru A, et al. Expression of parathyroid hormone-related peptide in human osteoarthritis. J Orthop Res 15:175 180, 1997.

204. Burton DW, Foster M, Johnson KA, et al. Chondrocyte calcium-sensing receptor expression is up-regulated in early guinea pig knee osteoarthritis and modulates PTHrP, MMP-13, and TIMP-3 expression. Osteoarthritis Cartilage 13:395 404, 2005.

205. Wu W, Billinghurst RC, Pidoux I, et al. Sites of collagenase cleavage and denaturation of type II collagen in aging and osteoarthritic articular cartilage and their relationship to the distribution of matrix metalloproteinase 1 and matrix metalloproteinase 13. Arthritis Rheum 46:2087 2094, 2002.

206. Shlopov BV, Lie WR, Mainardi CL, et al. Osteoarthritic lesions: involvement of three different collagenases. Arthritis Rheum 40:2065 2074, 1997.

207. Reboul P, Pelletier JP, Tardif G, et al. The new collagenase, collagenase-3, is expressed and synthesized by human chondrocytes but not by synoviocytes. A role in osteoarthritis. J Clin Invest 97:2011 2019, 1996.

208. Poole AR, Laverty S, Mwale F. Endochondral bone formation and development in the axial and appendicular skeleton. In: Henderson JE, Goltzman D, eds. The Osteoporosis Primer. Cambridge: Cambridge University Press, 2000, pp 3-17.

209. Serra R, Johnson M, Filvaroff EH, et al. Expression of a truncated, kinase-defective TGF-beta type II receptor in mouse skeletal tissue promotes terminal chondrocyte differentiation and osteoarthritis. J Cell Biol 139:541 552, 1997.

210. Yang X, Chen L, Xu X, et al. TGF-beta/Smad3 signals repress chondrocyte hypertrophic differentiation and are required for maintaining articular cartilage. J Cell Biol 153:35 46, 2001.

211. Wang FL, Connor JR, Dodds RA, et al. Differential expression of egr-1 in osteoarthritic compared to normal adult human articular cartilage. Osteoarthritis Cartilage 8:161 169, 2000.

212. Gordon GV, Villanueva T, Schumacher HR, et al. Autopsy study correlating degree of osteoarthritis, synovitis and evidence of articular calcification. J Rheumatol 11:681 686, 1984.

213. Ali SY, Rees JA, Scotchford CA. Microcrystal deposition in cartilage and in osteoarthritis. Bone Miner 17:115 118, 1992.

214. Swan A, Chapman B, Heap P, et al. Submicroscopic crystals in osteoarthritic synovial fluids. Ann Rheum Dis 53:467 470, 1994.

215. McCarty DJ. Calcium pyrophosphate dihydrate crystals deposition disease (pseudogout syndrome) clinical aspects. Clin Rheum Dis 3:61 89, 1977.

216. Schumacher HR, Jr. Synovial inflammation, crystals, and osteoarthritis. J Rheumatol Suppl 43:101 103, 1995.

217. Watanabe W, Baker DG, Schumacher HR, Jr. Comparison of the acute inflammation induced by calcium pyrophosphate dihydrate, apatite and mixed crystals in the rat air pouch model of a synovial space. J Rheumatol 19:1453 1457, 1992.

218. Fam AG, Morava-Protzner I, Purcell C, et al. Acceleration of experimental lapine osteoarthritis by calcium pyrophosphate microcrystalline synovitis. Arthritis Rheum 38:201 210, 1995.

219. Derfus B, Steinberg M, Mandel N, et al. Characterization of an additional articular cartilage vesicle fraction that generates calcium pyrophosphate dihydrate crystals in vitro. J Rheumatol 22:1514 1519, 1995.

220. Hashimoto S, Ochs RL, Rosen F, et al. Chondrocyte-derived apoptotic bodies and calcification of articular cartilage. Proc Natl Acad Sci USA 95:3094 3099, 1998.

221. Terkeltaub RA. What does cartilage calcification tell us about osteoarthritis? J Rheumatol 29:411 415, 2002.

P.47


222. Johnson K, Terkeltaub R. Upregulated ank expression in osteoarthritis can promote both chondrocyte MMP-13 expression and calcification via chondrocyte extracellular PPi excess. Osteoarthritis Cartilage 12:321 335, 2004.

223. Li B, Aspden RM. Mechanical and material properties of the subchondral bone plate from the femoral head of patients with osteoarthritis or osteoporosis. Ann Rheum Dis 56:247 254, 1997.

224. Nevitt MC, Lane NE, Scott JC, et al. Radiographic osteoarthritis of the hip and bone mineral density. The Study of Osteoporotic Fractures Research Group. Arthritis Rheum 38:907 916, 1995.

225. Burger H, van Daele PL, Odding E, et al. Association of radiographically evident osteoarthritis with higher bone mineral density and increased bone loss with age. The Rotterdam Study. Arthritis Rheum 39:81 86, 1996.

226. Hart DJ, Cronin C, Daniels M, et al. The relationship of bone density and fracture to incident and progressive radiographic osteoarthritis of the knee: the Chingford Study. Arthritis Rheum 46:92 99, 2002.

227. Dequeker J, Mokassa L, Aerssens J, et al. Bone density and local growth factors in generalized osteoarthritis. Microsc Res Tech 37:358 371, 1997.

228. Hochberg MC, Lethbridge-Cejku M, Scott WW, Jr., et al. Appendicular bone mass and osteoarthritis of the hands in women: data from the Baltimore Longitudinal Study of Aging. J Rheumatol 21:1532 1536, 1994.

229. Grynpas MD, Alpert B, Katz I, et al. Subchondral bone in osteoarthritis. Calcif Tissue Int 49:20 26, 1991.

230. Mansell JP, Bailey AJ. Abnormal cancellous bone collagen metabolism in osteoarthritis. J Clin Invest 101:1596 1603, 1998.

231. Naganathan V, Zochling J, March L, et al. Peak bone mass is increased in the hip in daughters of women with osteoarthritis. Bone 30:287 292, 2002.

232. McCarthy C, Cushnaghan J, Dieppe P. The predictive role of scintigraphy in radiographic osteoarthritis of the hand. Osteoarthritis Cartilage 2:25 28, 1994.

233. Dieppe P, Cushnaghan J, Young P, et al. Prediction of the progression of joint space narrowing in osteoarthritis of the knee by bone scintigraphy. Ann Rheum Dis 52:557 563, 1993.

234. Hutton CW, Higgs ER, Jackson PC, et al. 99mTc HMDP bone scanning in generalised nodal osteoarthritis. II. The four hour bone scan image predicts radiographic change. Ann Rheum Dis 45:622 626, 1986.

235. Huebner JL, Hanes MA, Beekman B, et al. A comparative analysis of bone and cartilage metabolism in two strains of guinea-pig with varying degrees of naturally occurring osteoarthritis. Osteoarthritis Cartilage 10:758 767, 2002.

236. Matsui H, Shimizu M, Tsuji H. Cartilage and subchondral bone interaction in osteoarthrosis of human knee joint: a histological and histomorphometric study. Microsc Res Tech 37:333 342, 1997.

237. Burr DB, Schaffler MB. The involvement of subchondral mineralized tissues in osteoarthrosis: quantitative microscopic evidence. Microsc Res Tech 37:343 357, 1997.

238. Watson PJ, Hall LD, Malcolm A, et al. Degenerative joint disease in the guinea pig. Use of magnetic resonance imaging to monitor progression of bone pathology. Arthritis Rheum 39:1327 1337, 1996.

239. van Osch GJ, van der Kraan PM, van Valburg AA, et al. The relation between cartilage damage and osteophyte size in a murine model for osteoarthritis in the knee. Rheumatol Int 16:115 119, 1996.

240. Seibel MJ, Duncan A, Robins SP. Urinary hydroxy-pyridinium crosslinks provide indices of cartilage and bone involvement in arthritic diseases. J Rheumatol 16:964 970, 1989.

241. Garnero P, Mazieres B, Gueguen A, et al. Cross-sectional association of 10 molecular markers of bone, cartilage, and synovium with disease activity and radiological joint damage in patients with hip osteoarthritis: the ECHODIAH cohort. J Rheumatol 32:697 703, 2005.

242. Muehleman C, Berzins A, Koepp H, et al. Bone density of the human talus does not increase with the cartilage degeneration score. Anat Rec 266:81 86, 2002.

243. Marcelli C, Favier F, Kotzki PO, et al. The relationship between osteoarthritis of the hands, bone mineral density, and osteoporotic fractures in elderly women. Osteoporos Int 5:382 388, 1995.

244. Hart DJ, Mootoosamy I, Doyle DV, et al. The relationship between osteoarthritis and osteoporosis in the general population: the Chingford Study. Ann Rheum Dis 53:158 162, 1994.

245. Dequeker J. Inverse relationship of interface between osteoporosis and osteoarthritis. J Rheumatol 24:795 798, 1997.

246. Westacott CI, Webb GR, Warnock MG, et al. Alteration of cartilage metabolism by cells from osteoarthritic bone. Arthritis Rheum 40:1282 1291, 1997.

247. Behets C, Williams JM, Chappard D, et al. Effects of calcitonin on subchondral trabecular bone changes and on osteoarthritic cartilage lesions after acute anterior cruciate ligament deficiency.J Bone Miner Res 19:1821 1826, 2004.

248. Franchimont P, Bassleer C, Henrotin Y, et al. Effects of human and salmon calcitonin on human articular chondrocytes cultivated in clusters. J Clin Endocrinol Metab 69:259 266, 1989.

249. Hayami T, Pickarski M, Wesolowski GA, et al. The role of subchondral bone remodeling in osteoarthritis: reduction of cartilage degeneration and prevention of osteophyte formation by alendronate in the rat anterior cruciate ligament transection model. Arthritis Rheum 50:1193 1206, 2004.

250. Smith MD, Triantafillou S, Parker A, et al. Synovial membrane inflammation and cytokine production in patients with early osteoarthritis. J Rheumatol 24:365 371, 1997.

251. Pelletier JP, Martel-Pelletier J, Abramson SB. Osteoarthritis, an inflammatory disease: potential implication for the selection of new therapeutic targets. Arthritis Rheum 44:1237 1247, 2001.

252. Lindblad S, Hedfors E. Arthroscopic and immunohistologic characterization of knee joint synovitis in osteoarthritis. Arthritis Rheum 30:1081 1088, 1987.

253. Ayral X, Pickering EH, Woodworth TG, et al. Synovitis: a potential predictive factor of structural progression of medial tibiofemoral knee osteoarthritis results of a 1 year longitudinal arthroscopic study in 422 patients. Osteoarthritis Cartilage 13:361 367, 2005.

254. Martel-Pelletier J, Alaaeddine N, Pelletier JP. Cytokines and their role in the pathophysiology of osteoarthritis. Front Biosci 4: D694-703, 1999.

255. Alaaeddine N, Di Battista JA, Pelletier JP, et al. Inhibition of tumor necrosis factor alpha-induced prostaglandin E2 production by the antiinflammatory cytokines interleukin-4, interleukin-10, and interleukin-13 in osteoarthritic synovial fibroblasts: distinct targeting in the signaling pathways. Arthritis Rheum 42:710 718, 1999.

256. Poole AR, Dieppe P. Biological markers in rheumatoid arthritis. Semin Arthritis Rheum 23:17 31, 1994.

257. Goldberg RL, Huff JP, Lenz ME, et al. Elevated plasma levels of hyaluronate in patients with osteoarthritis and rheumatoid arthritis. Arthritis Rheum 34:799 807, 1991.

258. Sharif M, George E, Shepstone L, et al. Serum hyaluronic acid level as a predictor of disease progression in osteoarthritis of the knee. Arthritis Rheum 38:760 767, 1995.

259. Sharma L, Hurwitz DE, Thonar EJ, et al. Knee adduction moment, serum hyaluronan level, and disease severity in medial tibiofemoral osteoarthritis. Arthritis Rheum 41:1233 1240, 1998.

260. Recklies AD, Baillargeon L, White C. Regulation of cartilage oligomeric matrix protein synthesis in human synovial cells and articular chondrocytes. Arthritis Rheum 41:997 1006, 1998.

261. Sharif M, Saxne T, Shepstone L, et al. Relationship between serum cartilage oligomeric matrix protein levels and disease progression in osteoarthritis of the knee joint. Br J Rheumatol 34:306 310, 1995.

262. Petersson IF, Boegard T, Svensson B, et al. Changes in cartilage and bone metabolism identified by serum markers in early osteoarthritis of the knee joint. Br J Rheumatol 37:46 50, 1998.

263. Conrozier T, Saxne T, Fan CS, et al. Serum concentrations of cartilage oligomeric matrix protein and bone sialoprotein in hip osteoarthritis: a one year prospective study. Ann Rheum Dis 57:527 532, 1998.

P.48


264. Vilim V, Olejarova M, Machacek S, et al. Serum levels of cartilage oligomericl matrix protein (COMP) correlate with radiographic progression of knee osteoarthritis. Osteoarthritis Cartilage 10:707 713, 2002.

265. Mazieres B, Garnero P, Gueguen A, et al. Molecular markers of cartilage breakdown and synovitis at baseline as predictors of structural progression of hip osteoarthritis. The ECHODIAH Cohort. Ann Rheum Dis 65:354 359, 2006.

266. Gineyts E, Garnero P, Delmas PD. Urinary excretion of glucosyl-galactosyl pyridinoline: a specific biochemical marker of synovium degradation. Rheumatology (Oxford) 40:315 323, 2001.

267. Jordan KM, Syddall HE, Garnero P, et al. Urinary CTX-II and glucosyl-galactosyl-pyridinoline are associated with the presence and severity of radiographic knee osteoarthritis in men. Ann Rheum Dis 65:871 877, 2006.

268. Masuhara K, Nakai T, Yamaguchi K, et al. Significant increases in serum and plasma concentrations of matrix metalloproteinases 3 and 9 in patients with rapidly destructive osteoarthritis of the hip. Arthritis Rheum 46:2625 2631, 2002.

269. Tchetverikov I, Lohmander LS, Verzijl N, et al. MMP protein and activity levels in synovial fluid from patients with joint injury, inflammatory arthritis, and osteoarthritis. Ann Rheum Dis 64:694 698, 2005.

270. Guerassimov A, Zhang Y, Cartman A, et al. Immune responses to cartilage link protein and the G1 domain of proteoglycan aggrecan in patients with osteoarthritis. Arthritis Rheum 42:527 533, 1999.

271. Champion BR, Poole AR. Immunity to homologous type III collagen after partial meniscectomy and sham surgery in rabbits. Arthritis Rheum 25:274 287, 1982.

272. Koch AE, Kunkel SL, Shah MR, et al. Macrophage inflammatory protein-1 beta: a C-C chemokine in osteoarthritis. Clin Immunol Immunopathol 77:307 314, 1995.

273. Hsu YH, Hsieh MS, Liang YC, et al. Production of the chemokine eotaxin-1 in osteoarthritis and its role in cartilage degradation.J Cell Biochem 93:929 939, 2004.

274. Hashimoto H, Tanaka M, Suda T, et al. Soluble Fas ligand in the joints of patients with rheumatoid arthritis and osteoarthritis. Arthritis Rheum 41:657 662, 1998.

275. Sharif M, Elson CJ, Dieppe PA, et al. Elevated serum C-reactive protein levels in osteoarthritis. Br J Rheumatol 36:140 141, 1997.

276. Spector TD, Hart DJ, Nandra D, et al. Low-level increases in serum C-reactive protein are present in early osteoarthritis of the knee and predict progressive disease. Arthritis Rheum 40: 723-727, 1997.

277. Linaker CH, Walker-Bone K, Palmer K, et al. Frequency and impact of regional musculoskeletal disorders. Baillieres Clin Rheumatol 13:197 215, 1999.

278. Torres L, Dunlop DD, Peterfy C, et al. The relationship between specific tissue lesions and pain severity in persons with knee osteoarthritis. Osteoarthritis Cartilage 2006.

279. Segal R, Avrahami E, Lebdinski E, et al. The impact of hemiparalysis on the expression of osteoarthritis. Arthritis Rheum 41: 2249-2256, 1998.

280. Decaris E, Guingamp C, Chat M, et al. Evidence for neurogenic transmission inducing degenerative cartilage damage distant from local inflammation. Arthritis Rheum 42:1951 1960, 1999.

281. Farrell M, Gibson S, McMeeken J, et al. Pain and hyperalgesia in osteoarthritis of the hands. J Rheumatol 27:441 447, 2000.

282. Ordeberg G. Characterization of joint pain in human OA. Novartis Found Symp 260:105 115; discussion 115-121, 277-109,2004.

283. Kosek E, Ordeberg G. Lack of pressure pain modulation by heterotopic noxious conditioning stimulation in patients with painful osteoarthritis before, but not following, surgical pain relief. Pain 88:69 78, 2000.

284. Kosek E, Ordeberg G. Abnormalities of somatosensory perception in patients with painful osteoarthritis normalize following successful treatment. Eur J Pain 4:229 238, 2000.

285. Farrell MJ, Gibson SJ, McMeeken JM, et al. Increased movement pain in osteoarthritis of the hands is associated with A beta-mediated cutaneous mechanical sensitivity. J Pain 1:229 242, 2000.

286. Garabedian BV, Lemaigre-Dubreuil Y, Mariani J. Central origin of IL-1beta produced during peripheral inflammation: role of meninges. Brain Res Mol Brain Res 75:259 263, 2000.

287. Dieppe PA, Lohmander LS. Pathogenesis and management of pain in osteoarthritis. Lancet 365:965 973, 2005.

288. Richardson JD, Vasko MR. Cellular mechanisms of neurogenic inflammation. J Pharmacol Exp Ther 302:839 845, 2002.

289. Heppelmann B, Pfeffer A, Schaible HG, et al. Effects of acetylsalicylic acid and indomethacin on single groups III and IV sensory units from acutely inflamed joints. Pain 26:337 351,1986.

290. Dubois RN, Abramson SB, Crofford L, et al. Cyclooxygenase in biology and disease. FASEB J 12:1063 1073, 1998.

291. Lotz M, Carson DA, Vaughan JH. Substance P activation of rheumatoid synoviocytes: neural pathway in pathogenesis of arthritis. Science 235:893 895, 1987.

292. Kimball ES. Substance P, cytokines, and arthritis. Ann N Y Acad Sci 594:293 308, 1990.

293. Takeba Y, Suzuki N, Kaneko A, et al. Evidence for neural regulation of inflammatory synovial cell functions by secreting calcitonin gene-related peptide and vasoactive intestinal peptide in patients with rheumatoid arthritis. Arthritis Rheum 42: 2418-2429, 1999.

294. Takeba Y, Suzuki N, Takeno M, et al. Modulation of synovial cell function by somatostatin in patients with rheumatoid arthritis. Arthritis Rheum 40:2128 2138, 1997.

295. Borovikova LV, Ivanova S, Zhang M, et al. Vagus nerve stimulation attenuates the systemic inflammatory response to endotoxin. Nature 405:458 462, 2000.

296. Wang H, Yu M, Ochani M, et al. Nicotinic acetylcholine receptor alpha7 subunit is an essential regulator of inflammation. Nature 421:384 388, 2003.

297. Kondo H, Nifuji A, Takeda S, et al. Unloading induces osteoblastic cell suppression and osteoclastic cell activation to lead to bone loss via sympathetic nervous system. J Biol Chem 280: 30192-30200, 2005.

298. Elefteriou F, Ahn JD, Takeda S, et al. Leptin regulation of bone resorption by the sympathetic nervous system and CART. Nature 434:514 520, 2005.

299. Persson E, Lerner UH. The neuropeptide VIP potentiates IL-6 production induced by proinflammatory osteotropic cytokines in calvarial osteoblasts and the osteoblastic cell line MC3T3-E1. Biochem Biophys Res Commun 335:705 711, 2005.

300. Matayoshi T, Goto T, Fukuhara E, et al. Neuropeptide substance P stimulates the formation of osteoclasts via synovial fibroblastic cells. Biochem Biophys Res Commun 327:756 764, 2005.

301. Elvenes J, Andjelkov N, Figenschau Y, et al. Expression of functional mu-opioid receptors in human osteoarthritic cartilage and chondrocytes. Biochem Biophys Res Commun 311:202 207, 2003.

302. Andjelkov N, Elvenes J, Martin J, et al. Opiate regulation of IL-1beta and TNF-alpha in cultured human articular chondrocytes. Biochem Biophys Res Commun 333:1295 1299, 2005.

303. Wieland HA, Michaelis M, Kirschbaum BJ, et al. Osteoarthritis-an untreatable disease? Nat Rev Drug Discov 4:331 344, 2005.

304. Smith GN, Jr., Mickler EA, Hasty KA, et al. Specificity of inhibition of matrix metalloproteinase activity by doxycycline: relationship to structure of the enzyme. Arthritis Rheum 42: 1140-1146, 1999.

305. Yu LP, Jr., Smith GN, Jr., Brandt KD, et al. Reduction of the severity of canine osteoarthritis by prophylactic treatment with oral doxycycline. Arthritis Rheum 35:1150 1159, 1992.

306. Brandt KD, Mazzuca SA, Katz BP, et al. Effects of doxycycline on progression of osteoarthritis: results of a randomized, placebo-controlled, double-blind trial. Arthritis Rheum 52:2015 2025, 2005.

307. Flechtenmacher J, Huch K, Thonar EJ, et al. Recombinant human osteogenic protein 1 is a potent stimulator of the synthesis of cartilage proteoglycans and collagens by human articular chondrocytes. Arthritis Rheum 39:1896 1904, 1996.

P.49


308. Caron JP, Fernandes JC, Martel-Pelletier J, et al. Chondroprotective effect of intraarticular injections of interleukin-1 receptor antagonist in experimental osteoarthritis. Suppression of collagenase-1 expression. Arthritis Rheum 39:1535 1544, 1996.

309. Zhang X, Mao Z, Yu C. Suppression of early experimental osteoarthritis by gene transfer of interleukin-1 receptor antagonist and interleukin-10. J Orthop Res 22:742 750, 2004.

310. Fernandes J, Tardif G, Martel-Pelletier J, et al. In vivo transfer of interleukin-1 receptor antagonist gene in osteoarthritic rabbit knee joints: prevention of osteoarthritis progression. Am J Pathol 154:1159 1169, 1999.

311. Chevalier X, Giraudeau B, Conrozier T, et al. Safety study of intraarticular injection of interleukin 1 receptor antagonist in patients with painful knee osteoarthritis: a multicenter study. J Rheumatol 32:1317 1323, 2005.

312. Spector TD, Conaghan PG, Buckland-Wright JC, et al. Effect of risedronate on joint structure and symptoms of knee osteoarthritis: results of the BRISK randomized, controlled trial [ISRCTN01928173]. Arthritis Res Ther 7:R625-633, 2005.

313. Bernik TR, Friedman SG, Ochani M, et al. Pharmacological stimulation of the cholinergic antiinflammatory pathway. J Exp Med 195:781 788, 2002.

314. Mohtai M, Smith RL, Schurman DJ, et al. Expression of 92-kD type IV collagenase/gelatinase (gelatinase B) in osteoarthritic cartilage and its induction in normal human articular cartilage by interleukin 1 . J Clin Invest 92:179 185, 1993.



Osteoarthritis. Diagnosis and Medical. Surgical Management
Osteoarthritis: Diagnosis and Medical/Surgical Management
ISBN: 0781767075
EAN: 2147483647
Year: 2007
Pages: 19

flylib.com © 2008-2017.
If you may any questions please contact us: flylib@qtcs.net