.1.) ..|l .l v. .. .. bl. Talliw‘. bx . f . ,. xxx}? . a. .3: ... E. $3... ...;_uz.fi< i. .. . . 1.133“. .735 El"; ii... 1 1.1: 3... 21.2.1; .1 .. «331.. .51 4 ¢ ., . . ... a...‘.m...w.‘_§.\._..: . , , \ M w ' LIBRARIES HIGAN STATE UNIVERSITY Q EXIST LANSING, MICH 48824-1048 This is to certify that the thesis entitled INVESTIGATION OF THE ACUTE INJURY RESPONSE OF ARTICULAR CARTILAGE IN VITRO AND IN VIVO: ANALYSIS OF VARIOUS THERAPEUTIC TREATMENTS presented by Steven Anthony Rundell has been accepted towards fulfillment of the requirements for the MS. degree in Mechanical Engineen’nL flmczy/md' Major Professors Signature MM 9, 2005’ d I Date MSU is an Affirmative Action/Equal Opportunity Institution - F‘- - - ._.. -_ ~ ‘ . . —v— v fl v ‘- — —' v v ‘ -v v‘ -‘ 'v v‘, 'Vv‘.’ .- v - ‘v’ - v - v -v-‘ PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE alas—cit “WWW .l 15‘ INVESTIGATION OF THE ACUTE INJURY RESPONSE OF ARTICULAR CARTILAGE IN VITRO AND IN V1 V0: ANALYSIS OF VARIOUS THERAPEUTIC TREATMENTS By Steven Anthony Rundell A THESIS Submitted to Michigan State University In partial fulfillment of the requirements For the degree of MASTER OF SCIENCE Department of Mechanical Engineering 2005 ABSTRACT INVESTIGATION OF THE ACUTE INJURY RESPONSE OF ARTICULAR CARTILAGE IN VIT R0 AND IN V1 V0: ANALYSIS OF VARIOUS THERAPEUTIC TREATMENTS By Steven Anthony Rundell Excessive mechanical loading to a joint can lead to matrix damage and chondrocyte death in articular cartilage. These injuries have been associated with the subsequent development of osteoarthritis. Understanding the mechanisms responsible for acute damage of cartilage is essential in the development of therapeutic methods to either prevent or treat these early alterations. The research presented in the current thesis uses both the in vitro chondral explant and in viva rabbit models to examine the acute response of articular cartilage to blunt impact loading. Chapter 1 addressed the issue of injury severity in chondral explants exposed to a culture medium prior to mechanical loading versus explants taken directly from the joint. This study hypothesized that excess fluid present in explants allowed to bathe in a culture medium would result in an artificially high amount of surface fissuring and associated chondrocyte death. Chapter 2 describes experiments in which knee joints of giant Flemish rabbits were subjected to a 6 Joule intensity impact and retro-patellar cartilage was studied in terms of surface fissuring as well as chondrocyte death. This study also evaluated the efficacy of a mild non-ionic surfactant, poloxamer 188, which has been found to reduce chondrocyte death in vitro by repairing damaged cells after an impact. Chapter 3 evaluated the efficacy of the nutraceutical glucosamine in enhancing the mechanical integrity of cartilage explants when pre-treated for a period of 6 days prior to an injurious unconfined compression test. DEDICATION I would like to thank my parents who have always supported me in every decision I have made. I would also like to thank them for instilling a good work ethic in me and exposing me to so many great opportunities. iii ACKNOWLEDGMENTS I would like to acknowledge my professor, mentor and good friend Dr. Roger Haut. I am extremely gratefiil to Dr. Mike Orth and Dr. Neil Wright for serving on my committee. I would like to express my gratitude towards Clifford Becket who not only contains a vast amount of knowledge but was always willing to share it with me. I would also like to thank Jane Walsh and Jean Atkinson for all their hard work and dedication. I would like to thank the undergrads that supplied me with great data and did a lot of hard work: Austin McPhillamy, Aaron Stewart, Jo Ewen, Michelle Foncannon, and Zach Kaltz Last but not least I would like to acknowledge my fellow graduate students for their help and friendship: Derek Baars, Jill Krueger, Dan Phillips, Eric Meyer, Lynn Martin, Mike Sinnot, and Eugene Kepich. iv TABLE OF CONTENTS LIST OF TABLES ...................................................................................................... vii LIST OF FIGURES ..................................................................................................... viii RESEARCH PUBLICATIONS ................................................................................... .ix INVESTIGATIONS INTO THE ACUTE INJURY RESPONSE OF ARTICULAR CARTILAGE IN VI T R0 AND IN VIVO: ANALYSIS OF VARIOUS THERAPEUTIC TREATMENTS Introduction ............................................................................................................ 1 References .............................................................................................................. 6 CHAPTER 1 TISSUE EQUILIBRATION ALTERS THE RESPONSE OF CARTILAGE EXPLANTS TO UNCONFINED COMPRESSION Abstract .................................................................................................................. 9 Introduction ............................................................................................................ 10 Methods ................................................................................................................. 13 Results .................................................................................................................... 19 Discussion .............................................................................................................. 21 References .............................................................................................................. 28 Figure Captions ...................................................................................................... 33 Figures ................................................................................................................... 37 CHAPTER 2 THE LIMITATION OF ACUTE NECROSIS IN RETRO-PATELLAR CARTILAGE AFTER A SEVERE BLUNT IMPACT TO THE IN VI V 0 RABBIT PATELLO- FEMORAL JOINT Abstract .................................................................................................................. 45 Introduction ............................................................................................................ 46 Methods ................................................................................................................. 49 Results .................................................................................................................... 52 Discussion .............................................................................................................. 54 References .............................................................................................................. 61 Figure Captions ...................................................................................................... 64 Figures ................................................................................................................... 66 CHAPTER 3 GLUCOSAMINE SUPPLEMENTATION CAN HELP LIMIT MATRIX DAMAGE AND ADJACENT CELL DEATH IN TRAUMATIZED EXPLANT S Abstract .................................................................................................................. 7O Introduction ............................................................................................................ 71 Methods ................................................................................................................. 74 Results .................................................................................................................... 79 Discussion .............................................................................................................. 82 References .............................................................................................................. 88 Tables ..................................................................................................................... 91 Figure Captions ...................................................................................................... 92 Figures ................................................................................................................... 95 CONCLUSIONS AND RECONHVIENDATIONS FOR FUTURE WORK .................. 103 APPENDIX A RAW DATA FROM CHAPTER 1 .............................................................................. 107 Mechanical Data ..................................................................................................... 107 Equilibrated Weight Data ....................................................................................... 109 Surface Fissure Lengths .......................................................................................... 110 Cell Viability Data .................................................................................................. 111 Increased Variability Mechanical Data ................................................................... 113 APPENDIX B RAW DATA FROM CHAPTER 2 .............................................................................. 116 Mechanical Data ..................................................................................................... 116 Surface Fissure Data ............................................................................................... 117 Cell viability counts for Time 0 .............................................................................. 118 Cell viability counts for 4 Day No Poloxamer ......................................................... 120 Cell viability counts for 4 Day Poloxamer .............................................................. 122 APPENDIX C RAW DATA FROM CHAPTER 3 .............................................................................. 124 Mechanical Data ..................................................................................................... 124 Surface Fissure Data ............................................................................................... 128 Cell Viability Data .................................................................................................. 130 Theoretical Material Property Table ....................................................................... 134 APPENDIX D RABBIT IMPACT SOP .............................................................................................. 135 vi LIST OF TABLES CHAPTER 3 Table 1 ................................................................................................................... 91 vii LIST OF FIGURES CHAPTER 1 Figure 1 .................................................................................................................. 37 Figure 2 .................................................................................................................. 37 Figure 3 .................................................................................................................. 38 Fgmm4. .................................................................................................................. 38 Figure 5 .................................................................................................................. 39 Figure 6 .................................................................................................................. 39 Figure 7 .................................................................................................................. 40 Figure 8 .................................................................................................................. 40 Figure 9 .................................................................................................................. 41 Figure 10 ................................................................................................................ 41 Figure 11 ................................................................................................................ 42 Figure 12 ................................................................................................................ 42 Figure 13 ................................................................................................................ 43 Figure 14 ................................................................................................................ 44 CHAPTER 2 Figure l .................................................................................................................. 66 Figure 2 .................................................................................................................. 66 F1gure3 .............................................................................. 67 Figure 4 .................................................................................................................. 67 Figure 5 .................................................................................................................. 68 Figure 6 .................................................................................................................. 69 Figure 7 .................................................................................................................. 69 CHAPTER 3 Figure l .................................................................................................................. 95 Figure 2 .................................................................................................................. 96 Figure 3 .................................................................................................................. 96 Figure 4 .................................................................................................................. 97 Figure 5 .................................................................................................................. 97 Figure 6 .................................................................................................................. 98 Figure 7 .................................................................................................................. 99 Figure 8 .................................................................................................................. 100 Figure 9 .................................................................................................................. 100 Figure 10 ................................................................................................................ 101 Figure 11 ................................................................................................................ 101 Figure 12 ................................................................................................................ 102 Figure 13 ................................................................................................................ 102 viii LIST OF PUBLICATIONS Peer Reviewed Manuscrimts Rundell, S., Haut RC, 2004. Tissue equilibration alters the response of cartilage explants to unconfined compression. Journal of Biomechanics, In Revision Jex, C.T., Rundell, S., Wan, C., MacDonald, B., Haut, RC, 2004. The effect of fixation types on the biomechanical properties of the Well osteotomy. Journal of Foot and Ankle Surgery, In Revision Rundell, S., Baars, D., Phillips, D., Haut, RC, 2004. Repair of damaged chondrocytes in the in vivo traumatized joint. Journal of Orthopaedic Research, In Revision. Baars, D., Rundell, S., Haut, RC, 2005. Treatment with the non-ionic surfactant Poloxamer pl 88 reduces TUNEL positive cells in bovine chondral explants exposed to injurious unconfined compression. Biomechanics and Modeling in Mechanobiology, In Review Peer Reviewed Abstrfl Rundell, S., Haut R.C., Tissue equilibration alters the response of cartilage explants to unconfined compression. 5 lst Annual Meeting of the Orthopaedic Research Society, 2005 Rundell, S., McPhilamy, A., Orth, M., Haut, R.C., Glucosamine supplementation can help limit matrix damage and adjacent cell death in traumatized explants. 5 lst Annual Meeting of the Orthopaedic Research Society, 2005 ix Introduction Lower extremity injuries account for approximately $21.5 billion each year in treatment, rehabilitation, and lost work day expenses (Miller, 1995). These injuries are a frequent outcome of automobile accidents, comprising nearly 25% of the total injtu'ies (Luchter et al., 1995). A more recent analysis of the National Accident Sampling System database has found that 10% of these injuries involve trauma to the knee joint, while only 5% of these injuries result in fracture. The Federal Safety Standard requires that femur forces generated during a knee-instrument panel collision must not exceed 10 kN, a force that has been shown to result in gross fracture of the patella, femur, or pelvis (Melvin et al., 1975; Patrick et al., 1965; Powell et al., 1975). However, the presence of chronic joint disease has been found in patients that have only suffered from a sub—fracture knee injury (States, 1970). Osteoarthritis (0A) is a chronic joint disease characterized by the loss of articular cartilage. Articular cartilage is the connective tissue that lines the ends of bones in diarthroidial joints. Its firnction is to provide a near frictionless surface on which bones can glide over, as well as absorb the shock from physical motion. Cartilage consists of a solid organic matrix and free interstitial fluid, which is mostly water. The solid matrix mainly consists of collagen (type II) and proteoglycans (PGs) (Mow et al., 1990). The PGs attract water and repel each other due to their electro-negativity. The network of collagen fibers resists this swelling (Grodzinsky et al., 1978), and gives cartilage its compressive stiffness. Chondrocytes (cartilage cells) are imbedded in the solid matrix and are responsible for synthesis and degradation of the solid matrix constituents. It has been suggested that the pathogenesis of 0A is the result of an imbalance in solid matrix turnover with degradation exceeding synthesis (Goldring, 2000; Malemud 1999). Clinically 0A is characterized by joint pain and narrowing of the joint, as diagnosed by radiological examination (Flores and Hochber, 1998). Pathologically, the disease exhibits a loss of cartilage and sclerosis of underlying bone. While the etiology of DA is not fully understood the risk of developing this disease is increased significantly in joints suffering a major injury (Felson et al., 2004). However, establishing a cause and effect relationship between joint injury and CA can be difficult as after an osteochondral injury, joint disease may not be diagnosed for 2 to 5 years, while less severe joint injuries may not be diagnosed for 10 or more years after the trauma (Wright, 1990). Animal models have been developed in order to study the associations between blunt impact trauma to a joint and the subsequent development of chronic joint disease. In a recent study a rigid impact mass was dropped with a 6 Joule intensity onto the flexed patello-femoral joint of a giant Flemish rabbit (Ewers et al., 2002). This study documented a progressive increase in the degradation of retro-patellar surface cartilage and the thickening of underlying subchondral bone after 3 years. A separate study that used a New Zealand white rabbit model and a 10 Joule impact intensity, found significantly advanced OA-like changes in the patello-femoral joint, with fibrillation, ulceration and erosion of retro-patellar cartilage within 6 months post-contusion (Mazieres et al., 1987)). Early OA-like changes have also been described using the flexed canine patello-femoral joint subjected to approximately 2.2 kN of impact force delivered with a gravity-dropped rigid mass (Thompson et al., 1991). While these studies indicate a link between blunt impact trauma to a joint and a subsequent degradation of joint tissue, the mechanisms responsible are still unclear. Studies that examine the acute response of articular cartilage after a blunt impact to a joint both in viva and in situ document gross surface lesions on the cartilage surface (Atkinson and Haut, 2001; Newberry et al., 1998). Also, In vitra chondral and osteochondral explant models have been developed in order to examine the acute response of injurious loading on articular cartilage. Studies that have subjected either chondral or osteochondral explants to an injurious level of unconfined compression have documented surface fissuring across the surface, as well as associated chondrocyte death (Ewers et al., 2001; Krueger et al., 2003). Acute chondrocyte death, or necrosis, has also been linked to degradation of articular cartilage after 1 year in viva (Simon et al., 1976). These studies found that a severe impact to a joint will result in gross fissuring of the articular cartilage, as well as a reduction in chondrocyte viability. While the exact mechanisms responsible for long term degradation of articular cartilage after a blunt impact are not fully understood, the acute development of surface fissuring and chondrocyte death are among the suspected factors. Matrix damage in the form of surface fissuring has been suggested to occur when the interstitial fluid pressurization in cartilage developed during loading becomes too great and exceeds the restraining capacity of the collagen network (Morel and Quinn, 2004; Pins et al., 1995). A study that exposed cartilage explants to varying rates of injurious loading documents an increased likelihood of surface fissuring at rates of loading that exceed the gel diffusion rate (Morel and Quinn, 2004). Theoretical models that simulate the effects of a blunt impact on an intact joint document the presence of high shear (distortional) strains near the articular surface, when cartilage is modeled as transversely isotropic and biphasic (Garcia et al., 1998; Donzelli et al., 1999). Interestingly, acute chondrocyte death has been found to occur predominately around surface cracks when cartilage is loaded at a high rate (Ewers et al., 2001; Lewis et al., 2003). However, when the rate of loading is slower and closer to the gel diffusion rate a more diffuse pattern of chondrocyte death throughout the thickness has been found (Morel and Quinn, 2004: Quinn et al., 2001). In a recent study that performed confined compression on chondral explants it was documented that cell death in the superficial zone occurred only when water was allowed to flow out (Milentijevic and Torzilli, 2005). This study suggests that chondrocyte death can either occur by compaction of the superficial zone, which occurs when water flows out, or by collagen tensile failure where chondrocyte death occurs around cracks. This study also suggests that hydrostatic pressure generated during a high rate of loading has a protective effect on cells, where as at slower rates water is allowed to flow which gives rise to distortional strains. Understanding the mechanisms responsible for acute damage of cartilage and associated chondrocyte death are essential in the firture development of therapeutic methods to either prevent or treat these early alterations. The research presented in the current thesis uses both the in vitra chondral explant and in viva rabbit models to examine the acute response of articular cartilage to blunt impact loading. Chapter 1 addressed the issue of injury severity in chondral explants exposed to a culture medium prior to mechanical loading versus explants taken directly from the joint. This study hypothesized that excess fluid present in explants allowed to bathe in a culture medium would result in an artificially high amount of surface fissuring and associated chondrocyte death. Chapter 2 describes experiments in which knee joints of giant Flemish rabbits were subjected to a 6 Joule intensity impact and retro-patellar cartilage was studied in terms of surface frssuring as well as chondrocyte death. No previous studies have attempted to document acute chondrocyte death as a result of a blunt impact in an in viva model. This study also evaluated the efficacy of a mild non-ionic surfactant, poloxamer 188, which has been found to reduced chondrocyte death in vitra (Phillips and Haut, 2004), in ‘repairing’ damaged cells after an impact in viva. Chapter 3 evaluated the efficacy of the nutraceutical glucosamine in enhancing the mechanical integrity of cartilage explants when pre-treated for a period of 6 days prior to an injurious unconfined compression test. The research presented in this thesis provides useful data in regards to the response of articular cartilage to blunt impact loading in both an in vivo and an in vitra setting. Furthermore, potential therapeutic treatments have been investigated and found to be effective in preventing damage that may result in long term degradation of articular cartilage. Future studies can utilize the data presented in this thesis to design experiments that will further help explain the mechanisms responsible for the efficacy of these treatments . References Atkinson PJ, Haut RC, 2001. Injuries produced by blunt trauma to the human patellofemoral joint vary with flexion angle of the knee. J Orthop Res Sep;19(5):827-33 Donzelli PS, Spilker RL, Ateshian GA, Mow VC, 1999. Contact analysis of biphasic transversely isotropic cartilage layers and correlations with tissue failure. J Biomech 32(10):]037-47 Ewers BJ, Dvoracek-Driksna D, Orth MW, Haut RC, 2001. The extent of matrix damage and chondrocyte death in mechanically traumatized articular cartilage explants depends on rate of loading. J Orthop Res 192779-84 Ewers BJ, Weaver BT, Sevensma ET, Haut RC, 2002. Chronic changes in rabbit retro- patellar cartilage and subchondral bone after blunt impact loading of the patellofemoral joint. J Orthop Res 20:545-50 Felson DT, 2004. An update on the pathogenesis and epidemiology of osteoarthritis. Radiol Clin North Am 42: 1-9 Flores R, Hochber M, 1998. Definition and classification of osteoarthritis. In: Brandt K, Doherty M, Lohmander S, editors. Osteoarthritis. Oxford: Oxford University Press: p. 1—12. Garcia JJ, Altiero NJ, Haut RC, 1998. An approach for the stress analysis of transversely isotropic biphasic cartilage under impact load. J Biomech Eng 120(5):608-13 Grodzinsky AJ, Lipshitz H, Glimcher MJ, 1978. Electromechanical properties of articular cartilage during compression and stress relaxation. Nature 275(5679):448-50 Goldring MB, 2000. The role of the chondrocyte in osteoarthritis. Arthritis Rheum 43:1916-26 Krueger JA, Thisse P, Ewers BJ, Dvoracek-Driksna D, Orth MW, Haut RC, 2003. The extent and distribution of cell death and matrix damage in impacted chondral explants varies with the presence of underlying bone. J Biomech Eng 125:114-9 Lewis, J.L, Deloria, LB, Oyen-Tiesma, M, Thompson, RC, Jr., Ericson, M, Oegema TR, Jr., 2003. Cell death after cartilage impact occurs around matrix cracks. J Orthop Res 21, 881-7 Luchter S, Walz MC, 1995. Long-term consequences of head injury. J Neurotrauma. 12(4):517-26 Malumed CJ, 1999. Fundamental pathways in osteoarthritis: an overview. Front Biosci 15: D659-61 Mazieres B, Blanckaert A, Thiechart M, 1987. Experimental post-contusive osteo- arthritis of the knee. Quantitative microscopic study of the patella and the femoral condyles. J Rheum 14:1 19-21 Melvin J, Stalnaker R, Alem N, Benson J, Mohan D, 1975. Impact response and tolerance of the lower extremities. 19th annual Stapp Car Crash Conf. 192543-559 Milentijevic D, Torzilli PA, 2005. Influence of stress rate on water loss, matrix deformation and chondrocyte viability in impacted articular cartilage. J Biomech 38, 493-502 Miller TR, Levy DT, 1995. The effect of regional trauma care systems on costs. Arch Surg 130(2):]88-93. Morel V, Quinn TM, 2004. Cartilage injury by ramp compression near the gel diffusion rate. J Orthop Res 22:145-151 Mow VC, 1990. Fundamentals of articular cartilage and meniscus biomechanics, in: Articular Cartilage and Knee Joint Function: Basic Science and Arthroscopy, J.W. Ewing, ed., Raven Press, Ltd., New York, ppl-18. Newberry WN, Garcia JJ, Mackenzie CD, Decamp CE, Haut RC, 1998. Analysis of acute mechanical insult in an animal model of post-traumatic osteoarthrosis. J Biomech Eng 120(6):704-9 Patrick LM, Koell CK, Mertz HJ Jr., 1965. Forces on the human body in simulated crashes. 9th Annual Stapp Car Crash Conf. 12:237-259 Phillips DM, Haut RC, 2004. The use of a non-ionic surfactant (P188) to save chondrocytes from necrosis following impact loading of chondral explants. J Orthop Res 22:1135-42 Pins GD, Huang EK, Christiansen DL, Silver FH, 1995. Effects of static axial strain on the tensile properties and failure mechanisms of self-assembled collagen fibers. J Appl Polym Sci 63:1429—40 Powell WR, Ojala SJ, Advani SH, 1975. Cadaver femur responses to longitudinal impacts. 19'h Annual Stapp Car Crash Conf. 19:561-579 Quinn TM, Allen RG, Schalet BJ, 2001. Matrix and cell injury due to sub-impact loading of adult bovine articular cartilage explants: effects of strain rate and peak stress. J Orthop Res 19(2):242—9 Simon WH, Richardson S, Herman W, Parsons JR, Lane J, 1976. Long-term effects of chondrocyte death on rabbit articular cartilage in vivo. J Bone Joint Surg Am 58:517-26 States JD, 1970. Traumatic arthritis: a medical dilemma, In: Proceedings of the 14th Annual Conference of the American Association for Automotive Medicine. 14:21- 28 Thompson RC, Oegema T, Lewis J, Wallace L, 1991. Osteoarthritic changes after acute transarticular load. J Bone Joint Surg Am 73:990-1001 Wright V, 1990. Post-traumatic osteoarthritis- A medico-legal minefield. Br J Rheum 29:474-8 CHAPTER 1: TISSUE EQUILIBRATION ALTERS THE RESPONSE OF CARTILAGE EXPLANTS TO UNCONFINED COMPRESSION Abstract: Excessive mechanical loading can lead to matrix damage and chondrocyte death in articular cartilage. Previous studies on chondral explants have not clearly distinguished to what extent the degree and the distribution of cell death are dependent on the amount of free swelling seen during tissue equilibration. The current study hypothesized that increased fluid content inside equilibrated chondral explants, when subjected to injurious compression, would lead to greater matrix damage. Equilibrated and non-equilibrated chondral explants were loaded to 3OMPa at a fast rate of loading (~600MPa/s) in an unconfined compression experiment. Stress-strain curves were created for each explant, and fit with a hyperelastic Ogden firnction using a least squares method. After 24 hours in a culture medium, matrix damage was assessed by the total length of surface fissures. The explants were also sectioned and stained for cell viability in the various layers of the cartilage. A separate group of explants was exposed to a displacement controlled unconfined compression experiment at a fast rate of loading (1200%c/s). These explants were dried out immediately after loading in order to separate contributions of the solid and fluid phase to the mechanical response. The stiffness of the equilibrated specimens was less than non-equilibrated specimens, and it correlated with the amount of fluid taken in during equilibration. More matrix damage in the form of surface fissuring and cell death in the superficial zone was documented in equilibrated explants. Matrix damage correlated positively with fluid gained during equilibration. The amount of cell death in the deep layer of the cartilage was less in equilibrated versus non-equilibrated explants, and also depended on the amount of fluid the specimens took in during equilibration. Correlations between peak stress and the percentage of water in the explant as well as dry weight in the displacement controlled tests were documented to decrease when explants were allowed to equilibrate. This study indicated that equilibration alters explant response to mechanical loading in terms of stiffiress, matrix damage and cell viability. The alterations in response depend on percentage uptake of fluid during tissue equilibration. These data may have relevance to the applicability of experimental data from equilibrated chondral explants to the in viva condition. Introduction: In early stages of osteoarthritis (OA), matrix damage is often seen as fissures on the articular surface (Brandt et al., 1986). Chondrocyte death, resulting from a reduction in tissue cellularity or chondrocyte malfunction, has been suggested to facilitate development of OA (Blanca et al., 1998; Buckwalter 1995). Excessive levels of mechanical load can generate acute matrix damage and chondrocyte death in articular cartilage (Ewers et al., 2001). However, establishing a cause and effect relationship between impact trauma and development of CA has been difficult as joint injuries with bone fi'acture may require 2-5 years to develop clinical symptoms, while less severe injuries may not be diagnosed for 10 or more years after trauma (Wright, 1990). Therefore, animal models have been used to study the potential association of trauma and disease. For example, blunt impact to the rabbit patella-femoral joint has indicated acute surface fissuring and a subsequent chronic degradation of the traumatized articular cartilage with thickening of the underlying subchondral bone (Newberry et al., 1998). Computational models of the joint have been developed to help investigate correlations 10 between the acute state of stress and strain with the subsequent changes in joint tissues (N ewberry et al., 1998), but these models are complex and not yet validated. To help clarify potential associations between mechanical events and early alterations in joint tissue, chondral (Ewers et al., 2001; Milentijevic et al., 2003; D’Lima et al., 2001) and osteochondral (Kreuger et al., 2003; Clements et al., 2001) explant models have also been developed for the laboratory setting. While these experimental models provide more ideal geometries for mechanical analyses, they are in vitro models that may not adequately represent the in vivo situation. In a study that compared the amounts of matrix damage and cell death in chondral versus osteochondral explants when exposed to 30 MPa of pressure, the presence of underlying bone reduced matrix damage and cell death (Krueger et al., 2003). Similarly in situ cadaveric human joint studies show minimal surface fissuring at contact pressures in the range of 20 to 30 MPa (Atkinson and Haut 2001, Atkinson and Haut 1995). Also, an in vivo rabbit study documented minimal or no occurrence of surface fissuring on the retropatellar surface at pressures of approximately 30 MPa delivered in roughly 50 milliseconds (Ewers et al., 2002). A study which used similar rates and amplitudes of loading on chondral explants documented excessive fissuring throughout the whole explant surface (Ewers et a1, 2001). Even studies using osteochondral explants exposed to the same level of high rate unconfined compression experienced substantial surface damage (Krueger et al., 2003). One possible explanation for differences in the extent of surface fissuring in in situ and in vitro models may be the amount of tissue fluid present in the in situ joint cartilage versus that of explanted tissue bathed for a period of time in a standard culture medium. This procedure is common in cartilage explant studies (Borelli et al., 1997; Bush and Hall, 11 2003; DiMicco et al., 2004: Sauerland et al., 2003) and is typically referred to as tissue equilibration. Interstitial fluid pressurization has been shown to contribute to the functional response of articular cartilage (Mansour and Mow, 1976; Mow et al. 1980; Zarek and Edward, 1963). This pressurization is the result of polyanionic glycosaminoglycans which create an electrostatic tissue swelling and resistance to compression. When collagen fibers are damaged, they lose their ability to resist swelling. This phenomenon has been documented in a number of studies. Grushko and co-workers ( 1989) noted that damaged or fibrillated human cartilage plugs swelled when placed in physiological saline. Sah and co-workers (1989) showed a swelling of excised bovine articular cartilage when placed in physiologic saline. Swelling of articular cartilage has been observed to be non-uniform and anisotropic (Myers et al., 1984, Maroudas et al., 1986) which characterizes the non-uniform organization of collagen (Clark, 1991; Clark, 1990) and the non-uniform distribution of negatively charged proteoglycans (Maroudas a al 1969). It follows that when explanting cartilage from a joint, the collagen matrix is disrupted allowing for changes in fluid content that arises when the tissue is bathed in a standard culture medium. Articular cartilage consists of a solid and fluid phase which, due to their interation, determines the mechanical characteristics of the tissue. Theoretical and computational analyses have separated the contributions of the fluid and solid phase in load bearing, and found that the fluid phase is responsible for approximately 90% of the load bearing (Ateshian et al., 1994; Ateshian and Wang, 1995; Macirowski et al., 1994). It is believed that interstitial fluid pressurization is the dominant mechanism of load 12 support (Park et al., 2003). Current studies, which induce or inhibit cartilage swelling by varying levels of NaCl concentrations in tissue baths (Maroudas, 1975; Lee et al., 1981), find significant effects occur in the tissue’s mechanical response characteristics. Given that the fluid content and the associated interstitial fluid pressurization has such a large effect on the load bearing capacity of this tissue, any changes in these parameters may result in an altered mechanical response during blunt immct. In addition, matrix damage in the form of surface fissuring, has been attributed to fluid pressurization and a subsequent tensile damage to the collagen network when the tissue is subjected to blunt impact loading (Morel and Quinn, 2004). Increases in cartilage explant fluid contentresulting from tissue equilibration, could potentially result in increased surface fissuring during high rates of unconfined compression. A previous study has shown that cell death in an intact bovine patella occurs exclusively around impact induced surface cracks (Lewis et al., 2003) suggesting that if in fact increased fluid results in a higher incidence of surface fissuring there will likely be an associated increase in chondrocyte death. The hypothesis of the current study was that excised chondral explants would swell in the medium during equilibration and the increase in fluid content would then alter the mechanical response of the explants resulting in greater matrix damage for equilibrated explants versus non-equilibrated tissue when exposed to a high rate of injurious compression. Methods and Materials: Three pairs of skeletally mature (12-24 months) bovine forelegs were obtained fi'om a local abattoir within 3 hours of slaughter. The legs were cut proximal to the 13 metacarpal surface leaving the joint intact. The legs were rinsed with distilled water, skinned, and rinsed again prior to opening the joint under a laminar flow hood. A 6-mm biopsy punch was used to make 52 cartilage plugs. These plugs were removed fiom the underlying bone with a scalpel and divided evenly into two groups, equilibrated and non- equilibrated. All explants were weighed immediately after removal from the joint. Twenty-six explants were assigned to a so-called “equilibrated” group. These were placed in Dulbecco’s Modified Eagle’s Medium (DMEM): F12 (Gibco, USA #12500-039) supplemented with 10 % fetal bovine serum, additional amino acids, and antibiotics (penicillin 100 units/ml, streptomycin lug/ml, amphotericin B 0.25 rig/ml). They were then washed three times, and allowed to equilibrate for one day in a humidity- controlled incubator (37° C, 7.2% C02, NuAire, Plymouth, MN). After equilibration these specimens were subjected to mechanical loading and retrn'ned to incubation for another 24 hours. They were then evaluated in terms of surface damage and cell viability. Another group of twenty-six were assigned to the “non-equilibrated” group. These explants were taken directly to a servo-hydraulic machine (Instron, model 1331 with 8500 upgraded electronics) for immediate mechanical testing of the explants. Tests were performed immediately after removal from the bone for each explant in order to avoid potential drying out. These specimens were then placed in supplemented DMEM:F12 and incubated for twenty-four hours before being examined for cell viability and matrix damage. Six equilibrated and non-equilibrated specimens were chosen as non- impact controls. These explants followed the same procedure as their respective groups with the exception of receiving impact. 14 The unconfined compression experiments were performed with a high rate of loading (50 ms to peak, ~600MPa/s). Each explant was placed between two highly polished stainless steel plates (figure 1). Prior to the test, the plates were pressed together at 100 N and the location of the machine actuator was recorded. The thickness of each specimen was determined by finding the difference in the actuator’s location at 5 N of load during the test and when the plates were pressed together prior to loading. Each specimen received a 0.5 N preload before being compressed to a peak load of 857N (~30MPa) using a single haversine load-time pulse. This protocol has been established in previous experiments from this laboratory (Ewers et al., 2001; Krueger et al., 2003). The load, time, and actuator displacement were recorded during each experiment with an accuracy of 0.1 N, 0.001 s, and 0.01 mm, respectively. After loading, each explant was washed three times in media before being returned to pre-assigned wells with one ml of supplemented media and incubated for an additional 24 hours. Twenty-four hours after the mechanical compression experiment, each explant was examined for matrix damage and cell viability. The surface of each explant was wiped with India ink. The explants were immediately photographed at 25X under a dissection microscope (Wild MSA, Wild Heerbrugg Ltd., Switzerland) to determine the total length of the surface fissures (figure 2). The total fissure length was measured with digital imaging software (Sigma Scan, SPSS Inc., Chicago, IL). One observer (AM) digitally recorded the length of the surface fissures in each photograph. For cell viability, two 0.5-mm slices were taken through the thickness at the center of each explant using a customized cutting tool (Ewers et al., 2001). The sections were stained with a kit containing calcein and ethidium bromide homodimer (Live/Dead 15 &Viabflity/Cytotoxity, Molecular Probes, Oregon). Each section was viewed under a fluorescence microscope at 100X (Lecia DM LB (frequency: ,50-60 Hz), Lecia Mikroskopie and Systeme GmgH, Germany). Full thickness, digital images were taken of a 2.5-mm length at the center of each explant. These images were partitioned into the superficial zone (top 20%), intermediate zone (middle 50%), and deep zone (bottom 30%). The viable (green) and dead (red) cells were manually counted by one observer (AS) using image software (Sigma Scan, SPSS Inc., Chicago, IL) (Figure 3). The percent of cell death was computed for each layer. Stress-strain curves were determined using the collected load and displacement data from the unconfined compression tests of the equilibrated and non-equilibrated explants (Figure 4). Compressive stress and corresponding strain were determined at 1, 5, 10, 15, and 20 MPa and used to generate an average response curve for each group of explants. To help quantify the overall responses of each group of explants during unconfined compression the stress-strain data were curve fit to a hyperelastic Ogden function using a least squares method, O'(.ua a, 8) : Z’L—lkl + g)a—1 _(1+ 8)-0.5a—1] a where c is the compressive stress, a is compressive strain, p. is a shear modulus parameter, and or is the stiffening factor. In order to ensure that the values for p. and a converged, the initial guesses, determined from a preliminary analysis, were multiplied by a random constant between 0.1 and 10 before conducting the curve fitting procedure. The process was repeated 100 times for each curve fit in order to obtain a guess that resulted in the minimum amount of error (Abramowitch and Woo, 2004). All guesses 16 tried in this range resulted in the same values for p. and a. R squared values were calculated for each curve in order to quantify the goodness of fit. The overall error (SSE) was calculated by subtracting the theoretical stress from the actual stress at l, 5, 10, 15, and 20 MPa and summing all the values. In order to quantify the goodness of fit of the curve, the r squared value was calculated. The r squared value (R2) was calculated by taking the ratio of the sum of squares of the regression (SSR) and the total sum of squares (SST). SSR = 2(ath(flaaa£)i —0-avg 2 i=1 n _ 2 SST - Z (0(actual)i — aavg i=1 R 2 _ SSR SST Unpaired t-tests were used to compare the peak strain, fissure length, n, or, and total cell death for equilibrated and non-equilibrated explants. Repeated factors AN OVA was used to compare cell death between the different layers through the thickness. Pearson correlation tests were performed in order to determine statistical significance of correlations between parameters within each group of explants. All data were reported as mean :l: one standard deviation. Statistical significance was indicated at p<0.05. Controlled Displacement Tests A separate group of a total of 96 (48 equilibrated and 48 un-equilibrated) explants from 4 bovine forelegs were subjected to a displacement controlled unconfined l7 compression experiment. The unconfined compression experiments were performed with a high rate of loading (50 ms to peak displacement, ~1200%a/s). Each explant was placed between two highly polished stainless steel plates (Figure 1). Prior to the test, the plates were pressed together at 100 N and the location of the machine actuator was recorded. The thickness of each specimen was determined by finding the difference in the actuator’s location at a preload of -0.5 N and when the plates were pressed together. Using the thickness of each explant the actuator arm was programmed to travel a total displacement corresponding to 60% strain (a) for each individual explant. Each specimen received a 0.5 N preload before being compressed to a peak strain of 60% using a single haversine load-time pulse. The load, time, and actuator displacement were recorded during each experiment with an accuracy of 0.1 N, 0.001 s, and 0.01 mm, respectively. After loading, each explant was dried out in a vacuum oven for 24 hours in order to determine the dry weight (DW). Besides differences in mechanical loading equilibrated and non-equilibrated explants used in these tests underwent the same procedure as previously described. Dry weights were obtained for these specimens in order to calculate the percentage of fluid present in the explant at the time of mechanical loading. Data from these tests was used to separate contributions of the fluid and solid phase to the mechanical response of cartilage explants. Chondrocyte viability and matrix damage were not assessed because this does not allow for the firll explant to be dried out after mechanical loading. Pearson correlation tests were performed to find correlations between peak stress and explant thickness, explant weight, dry weight, percent fluid content, and percent fluid gained during equilibration. Statistical significance was indicated at p<0.05. l8 Results: The peak stress generated during loading for the equilibrated group (28.86 i 0.71 MPa, n=20) was not significantly different than that for the non-equilibrated group (29.08i 0.32 MPa, n=20). Nonlinear stress-strain curves were generated for explants with and without equilibration. Equilibration decreased the explant stiffness by producing a shift in the response curves to the right (Figure 5). The maximum strain for the equilibrated explants (62 i 13 %, n=20) was significantly different than that for the non-equilibrated explants (56 i 6 %, n=20). There was no significant differences in the average thickness of the equilibrated explants (0.56 i 0.11 mm, n=20) versus that of the non-equilibrated explants (0.59 i 0.09 mm, n=20). The Ogden function fit the data with a high degree of accuracy. The r2 values for both the equilibrated and non-equilibrated explant response curves were 0.99 i 0.01 (n=20). All r2 values for individual tests were above 0.95. The shear modulus parameter, p, for the equilibrated group (2.0 i 0.9 MPa, n=20) was significantly different than for the non-equilibrated group (2.9 i 0.9 MPa, n=20). The value of at for the equilibrated group (-2.1 i 0.8 MPa, n=20) was not significantly different than that for the non-equilibrated group (-2.5 i 0.7 MPa, n=20). The average fluid gain for the equilibrated group was 13 i 5 % (n=20). All values were positive. The shear modulus parameter, p, significantly correlated inversely (r2 = 0.44, p = 0.001) with fluid gain for the equilibrated group (Figure 6). The peak strain, on the other hand, significantly correlated positively (r2 = 0.53, p < 0.001) with fluid gain in the equilibrated group (Figure 7). l9 Equilibration of the chondral explants also significantly increased the degree of surface fissuring for a 30 MPa unconfined compression compared to non-equilibrated specimens (Figure 8). The total length of surface fissuring for the equilibrated group was 57.6 :1: 15.2m (n=20) versus 45.1 i 17.8mm (n=20) for the non-equilibrated group (p=0.022). In the equilibrated group, there was a significant positive correlation (r'2 = 0.40, p = 0.003) between the length of surface fissuring and the amount of fluid gained by the explant (Figure 9). There were no visible surface fissures observed on the surfaces of the non-impacted specimens for either group of chondral explants. There was no significant difference (p = 0.53) in the percentage of total cell death in equilibrated (17 i 6 %, n=20) and non-equilibrated groups of explants (19 i 9 %, n=20). Significantly (p=0.006) less cell death, however, was noted in the deep zone of the equilibrated explants (12 i 10 %, n=20) (Figure 10) versus the non-equilibrated explants (25 i 17 %, =20) (Figure 11). In contrast, a significant increase (p=0.016) in the percentage of cell death was documented in the superficial zone of the equilibrated group of explants (36 i 12 %, n=20) versus the non-equilibrated group (27 i 10 %, n=20). There was also a significant inverse correlation (r2 = 0.36, p = 0.005) between fluid gain and the percentage of cell death in the deep zone for the equilibrated group of explants (Figure 12). No cell death was documented in the non-impacted explants from either group. Controlled Displacement Tests The peak strain generated during loading for the equilibrated group (62.5 i 3.7 %, =48) was not significantly different than that for the non-equilibrated group (61.4 i 3.2 %, n=48). The average fluid gain for the equilibrated group was 12 i 9 % (n=48). The 20 average thickness of the equilibrated explants (0.71 i 0.11 mm, n=48) was significantly higher (p<0.001) versus that of the non-equilibrated explants (0.62 i 0.10 mm, n=48). Significant correlations were found between peak stress and thickness, explant weight, dry weight, and percent fluid content for the non-equilibrated explants (Figure 13). Significant correlations were found between peak stress and thickness, dry weight, percent fluid content, and percent fluid gain for the equilibrated explants (Figure 14). The strength of the correlations, as determined by the r value,» decreased in all instances for the equilibrated group of explants. The significance (p<0.001) of the correlation between explant weight and peak stress documented in the non-equilibrated explants was lost (p=0. 167) in the equilibrated explants. Discussion: The objective of this study was to document differences in the biomechanical responses of chondral explants subjected to 30 MPa of unconfined compression that were either taken directly from the joint or equilibrated in a standard culture medium for 24 hours before impact loading. The hypothesis of the study was that excised chondral explants would swell in the medium during equilibration, and the increase in fluid content would then alter the mechanical response of the explants resulting in greater matrix damage for equilibrated explants versus non-equilibrated tissue during a high rate of unconfined compression. The current study confirmed that equilibration of chondral explants in a standard culture medium resulted in a significant uptake of fluid over 24 hours. The study also confirmed that the increased fluid content resulted in more matrix damage resulting fi'om a 30 MPa unconfined compression, as documented by an increased extent of surface fissuring in equilibrated versus non-equilibrated explants. 21 All equilibrated explants gained weight during the 24 hour period of equilibration despite that the standard culture media used was hypertonic (300 mosM) with regards to an isotonic osmolarity of 150 mosM. Parsons and Black (1979) documented an increase in cartilage thickness in intact rabbit femoral condyles in hypotonic solutions (<150 mosM) and a corresponding decrease in hypertonic solutions (>150 mosM). However, many explant studies document swelling at both physiological and hypertonic osmolarities. A study by T orzilli et a1. (1997), for example, shows that bovine chondral explants swell 10.2 i 11.0% in the thickness direction and approximately 3.31 i 2.67% radially when exposed to physiological saline (~150 mosM) for 30 minutes. In another study on calf cartilage disks swelling was observed axially 25-40% above the cut tissue thickness of 1.0 mm, when incubated in culture media (~300 mosM) for 6 days (Sah et al., 1989). In the current study the controlled displacement test equilibrated specimens experienced approximately a 14% increase in thickness during 24 hours in culture media. The current study documented a range in the amount of weight from 3 — 22% of the original wet weight. This large range of swelling was similar to the variations mentioned above in the previous studies by others. A study by Appleyard et al. (2003) documents site-specific trends in the contents of collagen and proteoglycans across the surface of articular cartilage on the ovine tibia. Similar findings were documented by Brama et al. (2000) for mature equine metacarpophalangeal joints. Areas of cartilage rich in collagen may better restrain tissue swelling in the bovine explant, but additional studies are needed to more completely explain variations in the percentage of fluid gain for various explants taken from different areas on the same joint surface. 22 The percent of dead cells in the deep zone following impact loading for equilibrated explants was substantially lower than that for non-equilibrated explants. This difference in deep zone cell death contrasted with a higher percentage of cell death in the superficial zone for the equilibrated versus non-equilibrated explants. Previous studies have documented the appearance of dead cells predominately around surface cracks (Ewers et al., 2001; Lewis et al., 2003) resulting from high rates of injurious compression. In the current study the increased length of surface fissures documented in the equilibrated explants helps explain the increase in the percentage of dead cells noted in the superficial zone. This, however, does not explain the seemingly protective effect of an increase in fluid volume on cells in the deep zone of equilibrated explants. Unfortunately, the mechanisms responsible for cell damage resulting from a blunt impact are yet unclear. The current data, in combination with previous studies that document zonal dependencies on the amount of free swelling in articular cartilage (Narmoneva et al., 1999), infer that susceptibility to cell death throughout the thickness of chondral explants during blunt impact may, in part, depend on local fluid content (Morel and Quinn, 2004). However, more research in this area is required in order to validate this hypothesis. The current study did suggest that changes in fluid content arising fiom 24 hours of equilibration in culture media had an effect on the levels of cell death throughout the thickness of chondral explants, but additional studies on mechanisms of cell death via blunt impact loading are warranted. On average, the total length of surface fissuring on equilibrated explants was greater than that on non-equilibrated explants. Previous studies correlate the onset of surface fissures with the development of high tensile strains in the collagen fibril network 23 (Morel and Quinn, 2004; Askew and Mow, 1978; Eberhardt et al., 1991). Studies that have modeled in situ blunt impact loading have documented high tensile strains near the surface of the cartilage (Li et al., 1995; Atkinson et al., 1998). The compressive strain at a given stress (Figure 1), and consequently the in-plane tensile strain, based on the Poisson effect, was less for non-equilibrated cartilage possibly explaining the occurrence of more surface fissuring in the equilibrated explants. This, however, does not explain why equilibrated explants exhibited a less stiff response to high rates of mechanical loading. The precise reasons as to why an increased fluid content would result in softening of the tissue is not clear, however, it is possible that swelling created geometric differences in the collagenous microstructure. A study by Wilson et al. (2004) used a poroviscoelastic fibril reinforced model to show that local stress and strain fields in cartilage are highly influenced by collagen orientation. Interestingly the softening of the equilibrated explants was observed to occur in the initial response to loading. This was exhibited by a significant decrease in the instantaneous shear modulus (u) with no significant difference in the stiffening factor (at) for equilibrated versus non-equilibrated explants. A possible explanation may be that fluid taken on by the explant during equilibration was more easily expelled during the initial phase of loading. Although theoretical models show that under impact loading cartilage responds as an incompressible elastic material (Armstrong et al., 1984; Mak et al., 1987), i.e., fluid cannot flow out, fluid taken on by chondral explants during equilibration may have substantially increased the initial permeability of the cartilage. Future studies may wish to observe swelling in the thickness direction and how it correlates to deformation during the initial loading phase. The current study suggests that the increased length of surface fissures documented in the equilibrated 24 versus non-equilibrated explants may have resulted from greater tensile stresses generated near the surface of equilibrated explants as a result of fluid gained during equilibration. A high fluid content, as a result of a damaged collagen network, is one of the early alterations documented in osteoarthritis (Buckwalter, 1995; Bank et al., 2000). Equilibrated explants displayed a reduction in dynamic stiffiress exhibited by a greater peak strain and a lower shear modulus parameter p. Both of these quantities were correlated with the amount of fluid gained during tissue equilibration. Osteoarthritic cartilage has also been shown to exhibit a softening effect (Ateshian et al., 1994; Mow et al., 1992; Setton et al., 1993) associated with tissue “edema” (Maroudas et al., 1986; Mankin and Thrasher 1975). The present results may suggest that damage to the collagenous network and associated tissue matrix swelling that occurs in early stages of OA make diseased tissue more vulnerable to a blunt mechanical injury. Blunt mechanical loads on a joint exhibiting early stages of disease may help accelerate the disease process in these individuals. While additional studies are needed to validate this hypothesis, the result may have important implications in forensic biomechanics. Additionally, tissue swelling could make joint tissues relatively more susceptible to blunt impact injury following a surgical procedure. In the controlled displacement tests significant correlations were documented between the peak stress generated during unconfined compression and various parameters. In all cases the correlations were stronger for non-equilibrated versus equilibrated explants. In the case of explant weight, a significant correlation was lost in the equilibrated group of explants. Previous studies that have determined the mechanical 25 properties of cartilage across the surface of an intact joint document changes in stiffness parameters that correlate with site-specific collagen and proteoglycan content (Appleyard et al., 2003; Brama et al., 2000). In the current study the decrease in the strength of correlations in the equilibrated explants may indicate that relationships between solid matrix constituents and the mechanical response of articular cartilage are shrouded by an artificially high amount of fluid acquired during ‘tissue equilibration’. Furthermore, the peak stress was found to significantly decrease as fluid gained during equilibration increased suggesting that increased fluid leads to a less stiff response. These data agree with the data obtained from the load controlled explant group. In conclusion, cartilage explant studies that allow for a period of tissue equilibration in standard culture media should consider the possible effects this may have on the tissue’s mechanical characteristics. Severe levels of matrix damage and associated cell death occurring in chondral explants exposed to injurious levels of compression may be in part due to an artificially high level of interstitial fluid due to the experimental protocol. Changes in the explant stiffness, resulting from tissue swelling, have an effect on chondrocyte viability throughout the thickness, as well as an increased susceptibility for surface fissuring. In order to help understand the large amount of variation in fluid gain between explants from the same joint surface noted in this study, additional studies are needed. Acknowledgements: This study was supported by grants from The National Center for Injury Control and Prevention, The Centers for Disease Control and Prevention (R49/CCR503607) and The TRW Automotive Fund. Its contents are solely the responsibility of the authors and do not necessarily represent the official views of the 26 CDC or The TRW Automotive Fund. The authors wish to gratefirlly acknowledge the help of Clifford Beckett, Austin McPhillamy, and Aaron Stewart (AS) for technical assistance during this study. 27 References Abramowitch, S.D., Woo, S.L.Y., 2004. An improved method to analyze the stress relaxation of ligaments following a finite ramp time based on the quasi-linear viscoelastic theory. Journal of Biomechanical Engineering 126, 92-97. Appleyard, R.C., Burkhardt, D., Ghosh, P., Read, R., Cake, M., Swain, M.V., Murrell, GA. 2003. Topographical analysis of the structural, biochemical and dynamic biomechanical properties of cartilage in an ovine model of osteoarthritis. Osteoarthritis and Cartilage 11, 65-77. Armstrong, C.G., Lai, W.M., Mow, V.C., 1984. An analysis of the unconfined compression of articular cartilage. Journal of Biomechanical Engineering 106, 165-173. Askew, M., Mow, V.C., 1978. The biomechanical function of the collagen fibril ultrastructure of articular cartilage. Journal of Biomechanics 100, 105-1 15. Ateshian, G.A., Lai, W.M., Zhu, W.B., Mow, V.C., 1994. An asymptotic solution for the contact of two biphasic cartilage layers. Journal of Biomechanics 27, 1347-1360. Ateshian, G.A., Wang, H., 1995. A theoretical solution for the fiictionless rolling contact of cylindrical biphasic articular cartilage layers. Journal of Biomechanics 28, 1341-1355. Atkinson, P.J., Haut, RC, 2001. Injuries produced by blunt trauma to the human patellofemoral joint vary with flexion angle of the knee. Journal of Orthopaedic Research 19, 827-833. Atkinson, P.J., Haut, RC, 1995. Subfracture insult to the human cadaver patellofemoral joint produces occult injury. Journal of Orthopaedic Research 13, 936-944. Atkinson, T.S., Haut, R.C., Altiero, N.J., 1998. Impact-induced fissuring of articular cartilage: an investigation of failure criteria. Journal of Biomechanical Engineering 120, 181-187. Bank, R.A., Soudry, M., Maroudas, A., Mizrahi, J., TeKoppele, J.M., 2000. The increased swelling and instantaneous deformation of osteoarthritic cartilage is highly correlated with collagen degradation. Arthritis and Rheumatism 43, 2202- 2210. 28 Blanco, F .J ., Guitian R., Vazquez-Martul E, de Toro F.J., Galdo F., 1998. Osteoarthritis chondrocytes die by apoptosis: A possible pathway for osteoarthritis pathology. Arthritis and Rheumatology 41, 284-289. Borrelli, J., Torzilli, P.A., Grigiene, R., Helfet, DL, 1997. Effect of impact load on articular cartilage: development of an intra-articular fracture model. Journal of Orthopaedic Trauma 11, 319-326 Brama, P.A., Tekoppele, J.M., Bank, R.A., Barneveld, A., van Weeren, RR, 2000. Functional adaptation of equine articular cartilage: the formation of regional biochemical characteristics up to age one year. Equine Veterinary Journal 2000 May;32(3):217-21. Brandt, KD., Mankin H.J., Shulman LE, 1986. Workshop on etiopathogenesis of osteoarthiritis. Journal of Rheumatology 13, 1126-1160. Buckwater, J.A., 1995. Osteoarthritis and articular cartilage use, disuse, and abuse: Experimental studies. Journal of Rheumatology Supplement 43,13-15. Bush, P.G., Hall, AC, 2003. The volume and morphology of chondrocytes within non- degenerate and degenerate human articular cartilage. Osteoarthritis and Cartilage. 11, 242-251. Clark, J .M., 1990. The organisation of collagen fibrils in the superficial zones of articular cartilage. Journal of Anatomy. 171, 117-130. Clark, J .M., 1991. Variation of collagen fiber alignment in a joint surface: a scanning electron microscope study of the tibial plateau in dog, rabbit, and man. Journal of Orthopaedic Research 9, 246-257. Clements, KM., Bee, Z.C., Crossingham, G.V., Adams, M.A., Sarif, M., 2001. How severe must repetitive loading be to kill chondrocytes in articular cartilage? Osteoarthritis and Cartilage 9, 499-507. DiMicco, M.A., Patwari, P., Siparsky, P.N., Kumar, S., Pratta, M.A., Lark, M.W., Kim, Y.J., Grodzinsky, A.J., 2004. Mechanisms and kinetics of glycosaminoglycan release following in vitro cartilage injury. Arthritis and Rheumatology. 50, 840- 848. D’Lima, D.D., Hashimoto, S., Chen, P.C., Colwell, C.W. Jr., Lotz, MK, 2001. Impactof mechanical trauma on matrix and cells. Clinical Orthopaedics and Related Research 391, 890-894 Eberhardt, AW., Lewis, J.L., Keer, LM, 1991. Normal contact of elastic spheres with two elastic layers as a model of joint articulation. Journal of Biomechanical Engineering 113, 410-417. 29 Ewers, B.J., Dvoracek-Driksna, D., Orth, M.W., Haut, RC, 2001. The extent of matrix damage and chondrocyte death in mechanically traumatized articular cartilage explants depends on rate of loading. 19(5), 779-784. Ewers, B.J., Jayaraman, V.M., Banglamaier, R.F., Haut, RC, 2002. Rate of blunt impact loading affects changes in retropatellar cartilage and underlying bone in the rabbit patella. Journal of Biomechanics 35, 747-755. Krueger, J.A., Thisse, P., Ewers, B.J., Dvoracek-Driksna, D., Orth, M.W., Haut, RC, 2003. The extent and distribution of cell death and matrix damage in impacted chondral explants varies with the presence of underlying bone. Journal of Biomechanical Engineering 125, 114-119. Lee, R.C., Frank, E.H., Grodzinsky, A.J., Roylance, D.K, 1981. Oscillatory compressional behaviour of articular cartilage and the associated electromechanical properties. Journal of Biomechanical Engineering 103, 280- 292. Lewis, J.L., Deloria, L.B., Oyen-Tiesma, M., Thompson, R.C., Jr., Ericson, M., Oegema, T.R., Jr., 2003. Cell death after cartilage impact occurs around matrix cracks. Journal of Orthopaedic Research 21, 881-887. Li, X., Haut, R.C., Altiero, N.J., 1995 . An analytical model to study blunt impact response of the rabbit P-F joint. Journal of Biomechanical Engineering 117, 485- 491. Macirowski, T., Tepic, 8., Mann, R.W., 1994. Cartilage stresses in the human hip joint. Journal of Biomechanical Engineering 116, 10-18. Mak, A.F., Lai, W.M., Mow, V.C., 1987. Biphasic indentation of articular cartilage-I. Theoretical analysis. Journal of Biomechanics 20, 703-714. Mankin, H.J., Thrasher, AZ, 1975. Water content and binding in normal and osteoarthritic human cartilage. Journal of Bone and Joint Surgery 57A, 76-80. Maroudas, A., 1975. Biophysical chemistry of cartilaginous tissues with special reference to solute and fluid transport. Biorheology 12, 233-248. Maroudas, A., Mizrahi, J., Katz, E.P., Wachtel, E.J., Soudry, M., 1986. Physiochemical properties and functional behavior of normal and osteoarthritic human cartilage. In: Kuettner, K., et al. (Eds), Articular Cartilage Biochemistry. Raven Press, New York, pp. 311-329. Maroudas, A., Muir, H., Wingham, J., 1969. The correlation of fixed negative charge with glycosaminoglycan content of human articular cartilage.Biochimica et Biophysica Acta 177, 492-500. 30 Mansour, J .M., Mow, V.C., 1976. The permeability of articular cartilage under compressive strain and at high pressures. Journal of Bone and Joint Surgery 58, 509-16. Milentijevic, D., Helfet, D.L., Torzilli, RA, 2003. Influence of stress magnitude on water loss and chondrocyte viability in impacted articular cartilage. Journal of Biomechanical Engineering 125, 594-601. Morel, V., Quinn, T.M., 2004. Cartilage injury by ramp compression near the gel diffusion rate. Journal of Orthopaedic Research 22, 145-151. Mow, V.C., Kuei, S.C., Lai, W.M., Armstrong, CG, 1980. Biphasic creep and stress relaxation of articular cartilage in compression? Theory and experiments. Journal of Biomechanical Engineering 102, 73-84. Mow, V.C., Ratcliffe, A., Poole, AR, 1992. Cartilage and diarthrodial joints as paradigms for hierarchical materials and structures. Biomaterials 13, 67-97. Myers, E.R., Lai, W.M., Mow, V.C., 1984. A continuum theory and an experiment for the ion-induced swelling behavior of articular cartilage. Journal of Biomechanical Engineering 106, 151-158. Narmoneva, D.A., Wang, J.Y., Setton, LA, 1999. Nonuniform swelling-induced residual strains in articular cartilage. Journal of Biomechanics 32, 401-408. Newberry, W.N., Garcia, J.J., Mackenzie, C.D., Decamp, C.E., Haut, RC, 1998. Analysis of acute mechanical insult in an animal model of post-traumatic osteoarthrosis. Journal of Biomechanical Engineering 120, 704-709. Newberry, W.N., Mackenzie C.D., Haut RC, 1998. Blunt impact causes changes in bone and cartilage in a regularly exercised animal model. Journal of Orthopaedic Research 16, 348-354. Park, S., Krishnan, R., Nicol], S.B., Ateshian, GA, 2003. Cartilage interstitial fluid load support in unconfined compression. Journal of Biomechanics 36, 1785-1796. Parsons, J .R., Black, J ., 1979. Mechanical behavior of articular cartilage: quantitative changes with alteration of ionic environment. Journal of Biomechanics 12, 765- 773. Sah, R.L., Kim, Y.J., Doong, J.Y., Grodzinsky, A.J., Plaas, A.H., Sandy, JD, 1989. Biosynthetic response of cartilage explants to dynamic compression. Journal of Orthopaedic Research 7, 619-636. Sauerland, K, Raiss, R.X., Steinmeyer, J ., 2003. Proteoglycan metabolism and viability of articular cartilage explants as modulated by the frequency of intermittent loading. Osteoarthritis and Cartilage. 11, 343-350. 31 Setton, L.A., Zhu, W.B., Mow, V.C., 1993. The biphasic poroviscoelastic behavior of articular cartilage: role of the surface zone in governing the compressive behavior. Journal of Biomechanics 26, 581-592. Torzilli, P.A., Grigiene, R., Huang, C., Friedman, S.M., Doty, S.B., Boskey, A.L., Lust, G., 1997. Characterization of cartilage metabolic response to static and dynamic stress using a mechanical explant test system. Journal of Biomechanics 30, 1-9. Wilson, W., van Donkelaar, C.C., van Rietbergen, B., Ito, K., Huiskes, R., 2004. Stresses in the local collagen network of articular cartilage: a poroviscoelastic fibril- reinforced finite element study. Journal of Biomechanics 37, 357-366. Wright, V. 1990. Post-traumatic osteoarthritis—A medico-legal minefield. British Journal of Rheumatology 29, 474-478. Zarek, J .M., Edward, J ., 1963. The stress-structure relationship in articular cartilage. Medical Electroncs and Biological Engineering 1, 497-507. 32 Figgre Captions Figure l (a) Photograph of servo-hydraulic testing machine set up. Explant specimens were placed between the actuator arm and the lower platen which is attached to a load cell. (b) A two dimensional drawing showing the loading scenario created by the servo-hydraulic testing machine. Figure 2 Photographs taken at 25X magnification were analyzed using digital imaging software. Manual digital pixel overlays were drawn along the length of all visible surface cracks. The software counted the total number of pixels. Pixel amounts were converted into units of millimeters by calibrating the software with a photograph of a ruler at the same magnification. Figure 3 Fluorescence images taken at 100X were separated into superficial, middle, and deep zones. The dead (red) and viable (green) cells were manually counted in digital imaging software. Dead cell counts were divided by total cell counts in each zone in order to yield a percentage of dead cells. Figure 4 Non-linear stress versus strain curves were generated for all of the impacted explants. Values of strain were interpolated from stress values of l, 5, 10, 15, and 20 mega Pascals. The curves were fit with a hyper-elastic Ogden fimction. 33 Figure 5 Average stress-strain plots for equilibrated and non-equilibrated chondral explants were constructed following blunt impact mechanical loading. Non-equilibrated explants exhibited a stiffer response resulting in a shift to the left on the graph. Figure 6 A plot of the shear modulus parameter p vs. water gain for the equilibrated explants fit with a least squares linear regression line showing a statistically significant negative correlation between this mechanical parameter and fluid intake during equilibration using a Pearson correlation coefficient test. Figure 7 A plot of peak strain vs. water gain for the equilibrated explants fit with a least squares linear regression line showing a statistically significant positive correlation between explant strain and fluid intake during equilibration using a Pearson correlation coefficient test. Figure 8 Gross photographs were used to determine the average total fissure length for cartilage explants following blunt impacts. Less fissuring was documented in non-equilibrated explants (a) compared to equilibrated specimens (b). Figure 9 A plot of the length of surface fissuring vs. water gain during 24 hours of equilibration fit with a least squares linear regression line showing a statistically significant positive correlation between the total length of surface cracks and fluid intake during equilibration using a Pearson correlation coefficient test. 34 Figure 10 Gross inspection of flouresence cell viability pictures taken at 100X show a higher occurrence of deep zone cell death in the non-equilibrated (b) versus equilibrated group (a). Figure 11 Comparison bar graph showing percentages of cell death in the various zones of equilibrated and non-equilibrated chondral explants after 30 MPa of unconfined compression. Significant differences (*) in cell death were noted in both deep and superficial zones between goups (p<0.05). Figure 12 A plot of the percentage of dead cells in the deep zone vs. water gained in the equilibrated explants fit with a least squares linear regression line showing a statistically significant negative correlation between cell death in the deep zone and fluid intake during equilibration using a Pearson correlation coefficient test. Figure 13 Pearson correlation tests revealed significant correlations between the peak compressive stress generated during unconfined compression and thickness (a), explant weight (b), dry weight (c), and the percentage of water content ((1) in the non-equilibrated explants. 35 Figure 14 Pearson correlation tests revealed significant correlations between the peak compressive stress generated during unconfined compression and thickness (a), dry weight (c), the percentage of water content ((1), and the percentage of fluid gained during equilibration. A significant correlation was not documented between explant weight and peak compressive stress (b). 36 Figures .2” Hydraulic Actuator ,4. Lower Platen a» Load Cell Figure l Drawn on _, digital ., '. m 1' ment Explant ' easu e _ lmes Specrmen (25X) \ Figure 2 37 559?“ 95’s” / '1’». .tefimfivmfigflusémasm Me Cross section slice divided into sections . 1‘.\..§'=.l§.%§.‘15. 2' Dead Cells (red) —|~ )N N 01 O 01 Stress (MPa 8 OUT 1 o 0.2 0.4 0.6 0.8 Strain Figure 4 38 Stress (MPa) 25 20 ~ + Non-Equil 15 - + Equil 10 a O T l l O 0.1 0.2 0.3 0.4 0.5 0.6 Strain (mm/mm) Figures 4.5 o 1’ 3.5 3 1\ R 3 E a 2.5 c o. . . 3 2 . x . 3 1: 1.5 A o E o o\ h 1 . a a 2 0.5 o_ in o I l I l 0.00% 5.00% 10.00% 15.00% 20.00% 25.00% Fluid Gain (°/o) Figure6 39 0.7 9 ~1 01 Peak Strain (mm/mm) 9 N .0 A 01 _o A 0.00% 5.00% 10.00% 15.00% Fluid Gain (%) 20.00% 25.00% Figure 7 Figure 8 40 1 00 90 8O 70 60 50 40 30 20 1 0 0 0.00% 5.00% 10.00% 15.00% 20.00% 25.00% Fluid Gain (%) Length of fissuring (mm) ~,___l Fiure9 Figure 10 41 g 40% 3 [ISupiedicial '6 30% .DMiddle g llDeep g 20% - 10% - 0% - Non-Equilibrated Equilibrated Figure 11 l 0.35 A7 I . l 0.3 —_ ‘I—I , _e ‘ ii 0.25 ° ’74—- O a 3 0'2 \ . :0 0.15 0 o a o g 0.1 - . . O \ 0.05 0 0 g n O l o \ 0 T I I c I 0.00% 5.00% 10.00% 15.00% 20.00% 25.00% L Fluld Gain (%) Figure 12 42 1 O 30 O E 0.0 . a 25 1 E - ~07 0.6 5 2° in E 15 j ‘” 0.4 i or E 0.2 ~ 3 5 a p— o T a e 0 f e t 0 20 40 60 8o 0 20 40 60 80 3) Peak Stress (MPa) b) Peak Stress (MP3) 0 75% . -. . 3, 0 , ° .. 70% — ° ' E 0 . g 65% - E , g 60% - g 0 g 55% - ED ~ g 50% ~ r=-0.719 5 0- “- 45% l p<0.001 0 TIT I l 40% U T f 0 20 40 60 80 O 20 40 60 80 C) Peak Stress (M Pa) d) Peak Stress (MPa) Figure 13 43 1.2 35 T . A 1 q . . 3o . . .. , E00- . 225] M V ‘ 0 O . . v _ o. . DO 0 . g 0.6 T .0 00 9 .. H 20 o o .0 .. . Q) o . a 15 - o g 0.4 — ’ r=0.313 '6 10 . r= 0.207 E 0.2 . p = 0.034 3 5 , p = 0.167 0 T T I o I I T 0 20 40 60 80 0 20 40 60 80 8) Peak Stress (MPa) b) Peak Stress (M Pa) 12 80% ’5 10 ~ . ,_. E 8 _ . o: . 3:3 700‘ ‘1 E ° 8 .g 6 '1 o .0 ° 0 60% 4 . 3 4 i . ° r= 0.539 3 50% 4 r--0.576 E 2 - p < 0.001 H- p < 0.001 0 r r r 40% r r t 0 20 40 60 80 0 20 40 60 80 C) Peak Stress (M Pa) d) Peak Stress (MPa) 50% 40% i ' .E 8 30% ~ g zoo/ti “- 10% - 0% 0 20 40 60 80 6) Peak Stress (MPa) Figure 14 44 CHAPTER 2: THE LIMITATION OF ACUTE NECROSIS IN RETRO- PATELLAR CARTILAGE AFTER A SEVERE BLUNT IMPACT TO THE IN VIVO RABBIT PATELLO-FEMORAL JOINT Abstract: Our laboratory has previously shown that severe levels of blunt mechanical load, generating contact pressures greater than 25 MPa, on chondral and osteochondral explants produce surface lesions and acute necrosis of chondrocytes. In vivo studies by our laboratory have also found surface lesions and chronic degradation of retro-patellar cartilage within 3 years following a 6 Joule impact intensity with an associated average pressure of 25 MPa in the rabbit patella-femoral joint. A hypothesis of the current study was that cellular necrosis is produced acutely in the retro-patellar cartilage of the aforementioned rabbit model as a result of a 6 Joule impact. Another hypothesis of the study was that an early injection of the non-ionic surfactant, poloxamer 188 (P188) would significantly reduce the percentage of necrotic cells in the traumatized retro-patellar cartilage. In the current study eighteen rabbits were equally divided into 3 groups. One group was termed ‘time zero’, and the other groups were carried out for 4 days. One ‘4 day’ group was administered a 1.5 m1 injection of P188 into the impacted joint immediately after trauma, while the other was injected with a placebo solution. Impact trauma produced surface lesions on retropatellar cartilage in all groups. Approximately 15% of retro-patellar chondrocytes suffered acute necrosis in the ‘time zero’ and ‘4 day no poloxamer’ groups. In contrast, significantly fewer (7%) cells suffered necrosis in the ‘4 day poloxamer’ group. The effect of this treatment was most significant in the superficial layer of the cartilage. 45 The study indicated the potential use of P188 surfactant in the acute repair of mechanically damaged cell membranes in the in vivo setting. Its use early after severe trauma to articular cartilage may allow sufficient time for damaged cells to heal, which may in turn mitigate the potential for a post-traumatic osteoarthritis in the joint. Additional studies are needed to improve the efficacy of this surfactant and to determine the long-term health of joint cartilage after P188 intervention. Introduction: Osteoarthritis (OA) affects over 21 million Americans and is the leading cause of disability in the United States (Wright, 1990). While the mechanisms responsible for this disease are unknown, the risk of CA is increased significantly in joints suffering a major injury (Felson, 2004). However, establishing a cause and effect relationship between joint injury and OA can be difficult as after an osteochondral injury joint disease may not be diagnosed for 2 to 5 years, while less severe joint injuries may not be diagnosed for 10 or more years after the trauma (U .5. Census Bureau, 2000). Clinically this disease is characterized by joint pain and narrowing of the joint, as diagnosed by radiological examination (Flores and Hochber, 1998). Pathologically, the disease exhibits a loss of cartilage and sclerosis of underlying bone. Experimental studies with animal models have been conducted in attempts to understand the potential association between acute joint injury and the chronic pathogenesis of OA. This laboratory has developed a model of post-traumatic OA using the Flemish Giant rabbit (Haut et al., 1995). In a recent study a rigid impact mass was dropped with 6 Joule of energy onto the flexed patello-femoral joint, using this model (Ewers et al., 2002). The study found surface fissures acutely with a progressive increase 46 in the degradation of retro-patellar surface cartilage and the thickening of underlying subchondral bone after 3 years. Another study found that during this impact event patella-femoral contact pressures reach approximately 25 MPa (Newberry et al., 1998). A study by another laboratory, using the New Zealand white rabbit subjected to a 10 Joule impact intensity, found significantly advanced OA-like changes in the patello- femoral joint, with fibrillation, ulceration and erosion of retro-patellar cartilage within 6 months post-contusion (Mazieres et al., 1987). Early OA-like changes have also been described using the flexed canine patello-femoral joint subjected to approximately 2.2 kN of impact force delivered with a gravity-dropped rigid mass (Thompson et al., 1991). The study acutely found surface fissures on retro-patellar cartilage with clefts in the underlying subchondral bone and calcified cartilage. After 6 months the surface fissures had loss of safi'anin-O staining and new bone formation in the underlying subchondral bone. While the studies described above found an association between blunt impact trauma to a joint and the subsequent development of OA-like changes in the traumatized cartilage, the mechanisms of acute damage leading to these pathological changes are yet unclear. Excessive mechanical loads resulting from a single blunt impact have been found to result in surface fissures and necrotic cell death in chondral (Ewers et al., 2001; Torzilli et al., 1999) and osteochondral explants (Krueger et al., 2003; Morel and Quinn, 2004). Forty megapascals (MPa) of unconfined compression on bovine chondral explants results in significant surface fissuring of the tissue with approximately 50% of cells acutely necrotic (Ewers et al., 2001). Necrotic cells have been found in the superficial zone (upper 20%) located near fissures and in the middle zone (next 50% of thickness) 47 zone, with a few dead cells found in the deepest zone of the cartilage, especially in osteochondral explants (Krueger et al., 2003). A critical threshold stress of 15-20 MPa was found for cell death and collagen matrix damage in unconfined compression using the bovine chondral explant (Torzilli et al., 1999)). A later study using osteochondral explants found that significant surface fissuring with adjacent cellular necrosis occurs when impact rates of loading exceed the gel diffusion rate of the cartilage (Morel and Quinn, 2004). This study also documented cell death largely in the superficial zone adjacent to fissures. Necrotic cell death has also been found in situ using the bovine patella subjected to a single impact of 53 MPa (Lewis et al., 2003)). In that study cell death occurred around surface cracks, but not in impacted areas away from these cracks. The authors suggested “that early stabilization of damaged areas of the cartilage may prevent late sequelae that lead to OA.” Other studies have also associated the physiopathology of articular cartilage in CA with the death of chondrocytes (Hashimoto et al., 1998). Few in vivo experimental studies, however, have been conducted that specifically associate cell death with the development of OA. In one study significant necrosis of chondrocytes was induced in the patella-femoral joint of rabbits by localized freezing of the tissue in vivo (Simon et al., 1976). After 1 year the affected cartilage had changes that were consistent with early signs of OA. These tissue changes were similar to those documented by our laboratory using the impacted joint model (Ewers et al., 2001). A hypothesis of the current study was that a 6 Joule blunt impact onto the rabbit patello- femoral joint would induce acute necrosis of chondrocytes in the retro-patellar cartilage. 48 A defining feature of cell death by necrosis is damage to the plasma membrane, and the inability of the cell to maintain ionic gradients across its membrane resulting in swelling with subsequent rupture of the cell (Duke et al., 1996). Due to the nature of necrotic cell death, mild surfactants have been used to restore integrity to cells after physical and chemical stresses (Clarke and McNeil, 1992; Papoutsakis, 1991). Specifically, poloxamer 188 (P188) has been found to ‘save’ neurons from early necrotic death after severe mechanical loading (Marks et al., 2001; Borgens et al., 2004). A recent study by our laboratory also found that P188 surfactant was able to ‘save’ chondrocytes from acute necrotic death in bovine chondral explants subjected to 25 MPa of unconfined compression (Phillips and Haut, 2004). A second hypothesis of the current study was that injection of P188 surfactant into the in vivo patella-femoral joint capsule shortly after a 6 Joule blunt impact would significantly reduce the percentage of necrotic cells in retro-patellar cartilage in the acute setting (4 days post trauma). Methods: Eighteen skeletally mature, Flemish Giant rabbits (aged 6-8 months) were used in this study. The project was approved by the All-University Committee on Animal Use and Care. The blunt impact experiments have been described previously (Haut et al., 1995). Briefly, a 1.33 kg mass with a flat 25 mm diameter aluminum impact interface was dropped from 0.46 m (6 Joules of impact energy) onto the right patellofemoral joint of anesthetized animals (2% Isoflurane and oxygen) (Figure l). The opposite limb was not impacted and used as a paired, un-impacted control. A load transducer (2.225-kN capacity) (model 31/432; Sensotec, Columbus, OH, USA.) was attached behind the rigid 49 interface to record impact loads. Peak contact load, time to peak, and total contact duration were collected at 10 kHz. The mass was arrested electronically after the first impact, preventing multiple impacts. Six rabbits were randomly selected as ‘time zero’ animals. These were sacrificed immediately after impact. The remaining 12 animals were sacrificed 4 days post impact. During these 4 days the animals were exercised 10 minutes a day at 0.3 mph on a treadmill (Oyen-Tiesma et al., 1998). The exercise protocol was initiated on the first day approximately 4 hours after impact. Animals were housed in individual cages (122 cm x 61 cm x 49 cm) when not exercising. Six of the animals (4 day poloxamer group) received a single 1.5 m1 injection of an 8mg/ml concentration of P1 88 surfactant in sterile phosphate-buffered saline (PBS) injected into the traumatized patella-femoral joint capsule shortly (within approximately 2 minutes) after impact. The remaining six animals (4 day no poloxamer group) received a sham injection of 1.5 mL sterile PBS into the impacted joint shortly after trauma The combination P188 in PBS and the PBS sham solutions were filter sterilized prior to injection in the joints using a 0.2 mm vacuum filter (Nalgene, Nalge Nunc Int., Rochester, New York, USA). After injection, the animal’s limb was exercised manually to help distribute the P188 surfactant and PBS solutions in the joint. Patellae from the impacted and the opposite un-impacted limbs were excised immediately after sacrifice from each animal. The retro-patellar surface was wiped with India ink to highlight surface defects and was photographed using a digital camera (Polaroid DMC2, Polaroid Corporation, Waltham, MA, USA) under a dissecting 50 microscope (Wild TYP 374590, Heerbrugg, Switzerland). Each patella was then wrapped in PBS soaked gauze and prepared for cell viability analyses. Osteochondral sections fi'om the impacted and opposite un-impacted patella from each animal were prepared for cell viability analyses. A low speed bone saw (model# 11- 1180-170, Buehler, Inc., Lake Bluff, IL) was used to remove trabecular bone from the overlying retro-patellar cartilage, leaving approximately 0.5 mm of subchondral bone (Figure 2). Full depth sections (0.5 mm thick) of retropatellar cartilage and subchondral bone were cut using a specialized cutting device (Ewers et al., 2001) from the area of known patella-femoral contact during blunt impact to the joint (Haut et al., 1995). The sections were stained with Calcein AM and Ethidium Homodimer, according to the manufacturer’s specifications (Live/Dead Cytotoxicity Kit, Molecular Probes, Eugene, OR). Approximately 8 sections from each patella were viewed near the midline of the patella using a fluorescence microscope (Leitz Dialux 20 (frequency: 50-60 Hz), Leitz Mikroskopie und Systeme GmgH, Wetlzar, Germany). The sections were photographed with a digital camera (Polaroid DMC2, Polaroid Corporation, Waltham, MA, USA). Each photographic section was divided into two zones: superficial (top 20% of the cartilage depth) and a deep zone (bottom 80%), using an established protocol (Phillips and Haut, 2004). A blinded observer (A.S.) manually counted live and dead cells of in each section using image analysis software (Sigma Scan, SPSS INC., Chicago, IL, USA). The percentage cell death in each zone and the total percentage cell death were determined for each section. The individual section data was averaged to yield the percentage of cell death in each zone and the total percentage of cell death for each limb. The length of surface fissuring was measured on 4 randomly selected impacted and un-impacted pairs 51 of patellae from each group using the same image analysis software, according to an established protocol (Ewers et al., 2001). Two factor repeated measures AN OVA and Student-Newman-Keuls post-hoe tests were used to detect statistical differences in the percentage of cell death, percentage of cell death in each zone, and the average total fissure length between impacted and un- impacted limbs, as well as between treatment groups. A two factor (time and limb) ANOVA was performed to detect differences between and within the ‘time zero’ and ‘4- day no poloxamer’ groups, while a separate two factor (P188 treatment and time) ANOVA was performed to detect differences between and within the ‘4-day no poloxamer’ and ‘4-day poloxamer’ groups. A single factor ANOVA was used to detect differences in peak impact loads and time to peak loads between treatment groups. Statistical significance in these tests was set at p<0.05. All experimental data are presented as mean 1- one standard deviation. Results: There were no statistical differences in the times to peak impact load or the magnitudes of peak load between the treatment groups. The peak impact load developed on the patella-femoral joint was 6757:1344 N, and the time to peak load was 3.53:1.0 ms for this study. The animals did not appear to favor one limb over the other during the 4 days of treadmill exercise, per observations by the licensed veterinary animal technician‘(J.A.). 52 Gross inspections of the impacted and un-impacted patellae indicated surface fissures on the retro-patellar surfaces (Figure 3). On the impacted patellae there were more fissures with a proximal to distal orientation that were located near the mid-line of the patella. This resulted in a statistically significant difference in the length of surface frssuring on the impacted versus the un-impacted patellae for each group (Figure 4). No significant differences in the length of surface fissures were detected between treatment groups for either impacted or un-impacted patellae, however with the small sample size of the current study the statistical power between groups for this analysis was below the desired level of 0.8. Magnification of the stained chondral sections showed the presence of necrotic cells in both the superficial and deep zones of impacted and the opposite, un-impacted retro-patellar cartilage (Figure 5). Analyses of the cell viability across the depth of the retro-patellar cartilage showed statistically significant differences in the percentages of dead cells between the impacted and un-impacted patellae for the ‘time 0’ (p<0.001) and for the ‘4 day no poloxamer’ treatment groups (p=0.002)(Figure 6). In contrast, there was not a statistically significant difference (p=0.072, power = 0.947) in the percentage of total cell death between the impacted and un-impacted patellae for the ‘4 day poloxamer’ group. The effect of P188 injections into the joint shortly after impact trauma was also evident in a significant difference (p=0.027) in the percentage of total cell death between the impacted limbs of the ‘4 day no poloxamer’ group and the ‘4 day poloxamer’ group. In contrast, there was not a significant difference in the percentage of total cell death in the Mpacted (p=0.195) or the un-impacted (p=0.899) retro-patellar cartilage between the ‘time 0’ and the ‘4 day no poloxamer’ groups. 53 Gross inspection of the cartilage sections indicated a large amount of cell death in the superficial zone (Figure 5b). While there were statistically significant differences between the percentages of cell death in the superficial zones of impacted versus un- impacted retropatellar cartilage for the ‘time zero’ (p=0.002) and the ‘4 day no poloxamer’ (p=0.008) groups, there was no significant difference (power = 0.994) recorded in the percentage of cell death in this zone for the ‘4 day poloxamer’ (p=0.095) group (Figure 7). In contrast, however, there were significant differences in the . percentages of cell death in the deep zones between impacted and un-impacted patellae in all treatment groups. Statistical differences were not detected for the superficial or deep zones between groups in this study, but with the small sample size the statistical power between groups was again less than 0.8. Discussion An objective of this study was to test the hypothesis that 6 Joules of impact energy delivered to the flexed rabbit patello-femoral joint, via a rigid impact interface, would cause a significant amount of cell death in the retro-patellar cartilage. Impact to the joint produced surface lesions, or fissures, on the retro-patellar cartilage. The proximal to distal orientation of the impact-induced surface fissures have previously been associated with the impact event (Haut et al., 1995). In the current study impact trauma resulted in the acute death of chondrocytes throughout the depth of the retro-patellar cartilage by approximately 15% more in the impacted versus the opposite un-impacted patella. A second objective of the study was to test the hypothesis that an immediate post- trauma injection of P188 surfactant into the joint would significantly reduce the percentage of dead cells in the retro-patellar cartilage. The current study found that the 54 percentage of total dead cells in the impacted versus the un-impacted retro-patellar cartilage was not statistically different for the ‘4 day poloxamer’ group. While the zonal layer data showed statistically significant differences in the percentage of dead cells in the superficial and deep zones of impacted versus un-impacted patella for the ‘time zero’ and ‘4-day no poloxamer’ groups, in the ‘4-day poloxamer’ group the effect was only statistically significant in the deep zone. These data suggest that P188 was most effective in ‘saving’ cells from necrotic death in the superficial zone in the current study. The current study verified that a blunt impact to the rabbit patella-femoral joint with a 6 Joule intensity produced statistically significant cell death in the retro-patellar cartilage. A defining feature of cell death by necrosis is swelling, due to the injured cell being unable to maintain ionic gradients across a damaged plasma membrane (Duke et al., 1996). In the current study the percentage of dead cells was measured by membrane disruption, as membrane damage, was documented by the ability of EthD-l (ethidium homodimer) to only pass through [a damaged cell membrane. This particular mechanism of damage was also supported by the efficacy of P188 surfactant to repair these cells in this study. A previous study showed that this surfactant specifically inserts into only damaged areas of a cell membrane (Marks et al., 2001). In a similar, in vivo study by D’Lima et al. (2001) which used New Zealand White rabbits, a 3 Joule intensity blunt impact to the patello-femoral joint resulted in a 14% increase in apoptotic (TUNEL positive) cells in retro-patellar cartilage versus the un-impacted limb at 4 days post trauma (D’Lima et al., 2001). The study verified cellular apoptosis by examining attributes of nuclear morphology indicative of apoptosis. Chondrocyte apoptosis has also been shown in human biopsy tissue near sites of chondral 55 fracture (Kim et al., 2002). Canine cartilage explants subjected to 5 MPa of cyclic compression at 0.3 Hz exhibited the development of both cellular necrosis and apoptosis progressively increasing with load duration and time. Necrosis was observed 2 hours after cessation of loading, whereas apoptosis (TUNEL-positive cells) was not significant until 48 or more hours after loading stopped. Apoptosis was verified in some cells using transmission electron microscopy (Chen et al., 2001). A study that induced osteochondral wounding of a joint also found significant percentages of necrotic and apoptotic cells in the tissue (Tew et al., 2000). These data suggest that mechanical injury to a joint may result in both necrotic and apoptotic cell death. The D’Lima et al. study (2001) also found 34% of chondrocytes in human chondral explants die via apoptosis when exposed 1014 MPa of unconfined compression. Administration of z-VAD.fmk, a pan-caspase inhibitor, reduced cell death to 25% in these explants. A limitation of the current study was that cell death by other mechanisms, for example by apoptosis, was not examined following P188 intervention. An influx of Ca2+, for example, into the chondrocyte prior to membrane rescaling by P188 may trigger an early programmed cell death (Aigner and Kim, 2002). A previous study on neuronal cells has also found P188 to be effective in limiting apoptosis as detected by TUNEL staining (Serbest, 2003). It is likely that after trauma, interventions like z-VAD.fmk could be used in combination with P188 to more effectively reduce chondrocyte death or dysfunction in the longer term. The current study found that early administration of P188 surfactant into the joint resulted in a decrease in the percentage of total dead cells in retro-patellar cartilage from approximately 18% to 7%. In a previous study by our laboratory using bovine chondral explants, 25 MPa of unconfined compression applied in 1 second resulted in 56 approximately 34% of the cells necrotic 24 hours after impact (Phillips and Haut, 2004). Immediate treatment of these explants after impact with an 8 mg/ml concentration of P188 surfactant reduced cell death to approximately 14 % at 24 hours post trauma. Differences in total cell death between the previous and current studies may have been due to the presence of underlying bone in the in vivo model that stiffened the articular cartilage and prevented excessive deformation during impact to the joint (Krueger et al., 2003). Furthermore, in contrast to the current study in which P188 surfactant appeared more effective in ‘saving’ cells in the superficial zone of the retro-patellar cartilage, the treatment of chondral explants in the previous study with this same concentration of surfactant significantly reduced cell death in all layers of the tissue. One explanation for the more limited efficacy of the treatment in the current study may relate to the penetration of the P188 surfactant. In the previous study the chondral explants were ‘pumped’ immediately after administration of the surfactant and approximately 22 hours after impact with 10 cycles of unconfined compression at 1 MPa pressure and a frequency of 1 Hertz. In the current study the animal joint was flexed approximately 10 times immediately after injection of the surfactant, and the animals were exercised daily beginning on the day of blunt impact to the joint. It was assumed that post trauma exercise would help ‘pump’ the surfactant into the cartilage. The ‘pumping’ of surfactant into the tissue, while not verified in the previous or current studies, may have been less effective in the in vivo joint due to differences in the intensity of the pressure and the unknown loading of the joint during treadmill exercise. Penetration of the surfactant into the cartilage may have also been limited by the underlying subchondral bone in the in vivo joint. The efficacy of this treatment, on the other hand, may be enhanced by a 57 reduction in the concentration of the surfactant solution, after the relationship between efficacy and solution concentration is established in future in vitro studies with both chondral and osteochondral explants. Another limitation of the current study was that the quantity of surfactant solution was limited to 1.5 ml. There was no basis for choosing this quantity other than to limit the amount of fluid in the small rabbit knee, and that in a previous study this quantity of polysulfated glycosaminoglycan solution was injected into the rabbit joint (Ewers and Haut, 2000). The relationships between optimization of the surfactant concentration, the quantity of fluid, and the timing of the intervention post impact should be determined in future studies using the established rabbit model, or possibly another larger, yet undeveloped in vivo animal model. The current study used a small sample size (n=6) and an in vivo model which yielded large variations in both surface fissure and chondrocyte viability data. This resulted in an insufficient amount of statistical power between treatment groups. The average total length of surface fissures in the ‘4-day poloxamer’ group was approximately 4 times greater in the un-impacted and approximately 2 times greater in the impacted patellae than the ‘4-day no poloxamer’ group on average. A statistically significant difference was not detectable due to the low power. The greater length of surface fissuring in the ‘4-day poloxamer group’ resulted from two animals that exhibited large amounts of base-line surface fissuring on their un-impacted retropatellar surface. These animals were not disregarded due to the already small sample size. Furthermore, previous studies have documented chondrocyte death around surface cracks (Ewers et al., 2001; Lewis et al., 2002), but the ‘4-day poloxamer’ group exhibited significantly less percentage of dead cells on average in the impacted patellae versus the ‘4-day no 58 poloxamer’ group. These data suggest that the increased, baseline surface fissuring documented in the ‘4-day poloxamer’ group did not adversely affect the results of the study. In summary, 6 Joules of blunt impact to the flexed rabbit patello-femoral joint was found to result in a significant percentage of necrotic cells in the retro-patellar cartilage immediately after insult. It is hypothesized that the chronic degradation of this cartilage, which has been documented by our laboratory in a 3 year post trauma study (Ewers et al., 2002), may be due, at least in part, to the death of these chondrocytes. The current study also found, in concert with earlier studies on chondral explants (Phillips and Haut, 2004), that immediate administration of the traumatized joint with P188 surfactant resulted in a statistically significant decrease in the percentage of necrotic cells. The above hypothesis on a mechanism of post traumatic osteoarthritis may be tested in the future with injection of P188 surfactant into the joint immediately after blunt insult. The long term consequences of ‘saving’ these cells fi'om early necrotic death, in terms of them ultimately becoming apoptotic and producing degradation enzymes into the joint tissue (Pickvance et al., 1993) must be investigated in future studies. This intervention with P188 surfactant, however, may also allow sufficient time to evaluate the biological condition of these and other cells in the traumatized tissue and utilize additional pharmacological treatments, such as the caspase inhibitor mentioned earlier, if needed, for the long term survival of the joint cartilage. Acknowledgements: This study was supported by grants fi'om The Centers for Disease Control and Prevention, The National Center for Injury Control and Prevention 59 (R49/CCR503607) and The TRW Automotive Fund. Its contents are solely the responsibility of the authors and do not necessarily represent the official views of the CDC or the TRW Corporation. The authors wish to gratefully acknowledge the help of Clifford Beckett, Aaron Stewart (AS), and Jean Atkinson (J.A.) for technical assistance during this study. 60 References Aigner T, Kim HA. Apoptosis and cellular vitality: issues in osteoarthritic cartilage degeneration. Arthritis Rheum 2002 ;46:1986-96 Borgens RB, Bohnert D, Duerstock B, Spomar D, Lee RC. Subcutaneous tri-block copolymer produces recovery from spinal cord injury. J Neurosci Res 2004;76: 141 -54. Chen CT, Burton-Wurster N, Borden C, Huegger K, Bloom SE, Lust G. Chondrocyte necrosis and apoptosis in impact damaged articular cartilage. J Orthop Res. 2001;19:703-11 Clarke MSF, McNeil PL. Syringe loading introduces macromolecules into living mammalian cell cytosol. J Cell Sci 1992;102:533-41 Costouros JG, Dang AC, Kim HT. Inhibition of chondrocyte apoptosis in vivo following acute osteochondral injury. Osteoarthritis Cartilage 2003;11:756-9 D'Lima DD, Hashimoto S, Chen PC, Colwell CW, Lotz MK Impact of mechanical trauma on matrix and cells. Clin Orthop Rel Res 2001;39IS:90-9 Duke RC, Ojcius DM, Young JDE. Cell suicide in health and disease. Sci Am 1996;275:80-7 Ewers BJ, Dvoracek-Driksna D, Orth MW, Haut RC. The extent of matrix damage and chondrocyte death in mechanically traumatized articular cartilage explants depends on rate of loading. J Orthop Res 2001 ;19:779-84 Ewers BJ, Haut RC. Polysulphated glycosaminoglycan treatments can mitigate decreases in stiffness of articular cartilage in a traumatized animal joint. J Orthop Res 2000;18:756-61 Ewers BJ, Weaver BT, Sevensma ET, Haut RC. Chronic changes in rabbit retro-patellar cartilage and subchondral bone after blunt impact loading of the patella femoral joint. J Orthop Res 2002;20:545-50 Felson DT. An update on the pathogenesis and epidemiology of osteoarthritis. Radio] Clin North Am 2004 ;42:1-9 Flores R, Hochber M. Definition and classification of osteoarthritis. In: Brandt K, Doherty M, Lohmander S, editors. Osteoarthritis. Oxford: Oxford University Press: 1998. p. 1—12. Hashimoto S, Ochs RL, Komiya S, Lotz M. Linkage of chondrocyte apoptosis and cartilage degradation in human osteoarthritis. Arthritis Rheum 1998;41:1632-8 61 Haut RC, Ide TM, DeCamp CE. Mechanical responses of the rabbit patella-femoral joint to blunt impact. J Biomech Eng 1995;117:402-8 Kim HT, Lo MY, Pillarisetty R. Chondrocyte apoptosis following intraarticular fracture in humans. Osteoarthritis Cartilage 2002;10:747-9 Krueger JA, Thisse P, Ewers BJ, Dvoracek-Driksna D, Orth MW, Haut RC. The extent and distribution of cell death and matrix damage in impacted chondral explants varies with the presence of underlying bone. J Biomech Eng 2003;125:114-9 Lewis JL, Deloria LB, Oyen-Tiesma M, Thompson RC, Ericson M, Oegema TR. Cell death after cartilage impact occurs around matrix cracks. J Orthop Res 2003;21:881-7 Marks JD, Pan CY, Bushell T, Cromie W, Lee RC. Amphiphilic, tri-block copolymers provide potent membrane-targeted neuroprotection. FASEB 2001;15:1 107-9 Mazieres B, Blanckaert A, Thiechart M. Experimental post-contusive osteo-arthritis of the knee. Quantitative microscopic study of the patella and the femoral condyles. J Rheum 1987;14:119-21 Morel V, Quinn TM. Cartilage injury by ramp compression near the gel diffusion rate. J Orthop Res 2004;22:145-51 Morel V, Quinn TM. Short-term changes in cell and matrix damage following mechanical injury of articular cartilage explants and modelling of microphysical mediators. Biorheo logy 2004 ;4 1 2509-19 Newberry WN, Garcia JJ, Mackenzie CD, DeCamp CE, Haut RC. Analysis of acute mechanical insult in an animal model of post-traumatic osteoarthrosis. J Biomech Eng 1998;120:704-9 Oyen-Tiesma M, Atkinson J, Haut RC. A method for promoting regular rabbit exercise in orthopaedics research. Contempy Top Lab Anim Sci 1998;37:77-80 Papoutsakis ET. Media additives for protecting freely suspended animal cells against agitation and aeration damage. TIBTECH 1991 ;9:316-24 Phillips DM, Haut RC. The use of a non-ionic surfactant (P188) to save chondrocytes fiom necrosis following impact loading of chondral explants. J Orthop Res 2004;22:] 135-42 Pickvance EA, Oegema TR, Thompson RC. Immunolocalization of selected cytokines and proteases in canine articular cartilage after transarticular loading. J Orthop Res 1993;11:313-23 Serbest G. In vitro neuronal cell injury model: characterization and treatment strategies. Ph.D. dissertation, Drexel University, 2003;103-42 62 Simon WH, Richardson S, Herman W, Parsons JR, Lane J. Long-term effects of chondrocyte death on rabbit articular cartilage in vivo. J Bone Joint Surg Am 1976;58:517-26 Tew SR, Kwan APL, Hann A, Thomson BM, Archer CW. The reactions of articular cartilage to experimental wounding. Role of apoptosis. Arthritis Rheum 2000;43:215-25 Thompson RC, Oegema T, Lewis J, Wallace L. Osteoarthritic changes after acute transarticular load. J Bone Joint Surg Am 1991;73:990-1001 Torzilli PA, Grigiene R, Borrelli J, Jr., Helfet DL. Effect of impact load on articular cartilage. Cell metabolism and viability, and matrix water content. J Biomech Eng 1999;121:433-41 U.S.Census Bureau. State and County Quickfacts. 2000. US. Public Information Office. Wright V. Post-traumatic osteoarthritis- A medico-legal minefield. Br J Rheum 1990;29:474-8 63 Figgge Captions Figure 1 (a) Photograph of the rabbit impact set-up (b) Sketch of the setup depicting the load alignment with the knee joint. Figure 2 Patellae were glued to cylindrical block (a,b) in order to align the surface parallel with the Isomet gravity saw blade (0). The saw was used to create osteochondral sections of the retro-patellar surface. Figure 3 Digital photographs (25X) of un-impacted (a) and impacted (b) retro-patellar surfaces were taken for each rabbit. Impact induced fissures were stained with India ink. The fissures were digitally measured and recorded with digital imaging software. Figure 4 Digitally measured and recorded surface fissure lengths were averaged and depicted in a bar chart. Brackets indicate significant differences between groups using a 2 factor repeated measures AN OVA . There was a statistically significant greater average length of surface fissuring in impacted versus un-impacted patellae in all of the groups. Figure 5 Impacted (a) and un-impacted (b) patellar cross sectional slices were stained for cell viability. Live cells were stained green and dead cells were stained red. The impacted patellae were subjected to a 6 Joule impact and stained with calcein AM and ethidium homodimer either immediately following impact or 4 days post-impact. Images (a) and 64 (b) show patellae that were untreated and viewed immediately after impact. The impacted patella (b) shows extensive areas of cell death in the surface zone after trauma. Figure 6 Statistically greater amounts of cell death were documented in the impacted versus un- impacted patellae of both the “time 0” and the “4-day no poloxamer” groups. A significantly greater amount of cell death was recorded in the impacted patellae of the ‘4 day no poloxamer’ group versus the ‘4 day poloxamer’ group. The brackets denote statistical significance by a 2 factor repeated measures ANOVA. Figure 7 There was a statistically greater percentage of cell death in the superficial and deep zones of the impacted versus un-impacted patellae in both the “time 0” and “4 day no poloxamer” groups. There was also a significant difference in the total percentage of cell death in the deep zones of impacted versus un-impacted patellae in the “4 day poloxamer” group. The brackets denote statistical differences using a 2 factor repeated measures AN OVA performed separately for the superficial and deep zones. 65 Fr ures l igure F ) 2 Figure 66 Average Total Length of Fissures (mm) r—I —a N N DJ w ah A M o u- o u: o u. o u. o u- o I Un-impacted El Impacted time zero Figure 3 4-Day No Poloxamer Figure 4 67 4-Day Poloxamer Figure 5 68 40.00 a I I l 'i transacted? * * ” ' * "'lqtmpacteg ;— time zero 4-Day No Poloxamer 4-Day Poloxamer Figure 6 70.00 fi , t — _ l I Superficral l——| . 60.00 T WW W lIEIDeep Average % Cell Death N w J: U: .0 .0 .0 .O o o o o o o o o | l I 10.00 - 0.00 - Empacted impacted Bunnacted impacted unimpacted impacted time 230 —-‘ 4-Day No Poloxama I '— 4-Day Poloxamer --] Figure 7 69 CHAPTER 3: GLUCOSAMINE SUPPLEMENTATION CAN HELP LIMIT MATRIX DAMAGE AND ADJACENT CELL DEATH IN TRAUMATIZED EXPLANTS Abstract Severe blunt mechanical trauma has been linked with the onset of osteoarthritis. Matrix damage in the form of surface fissuring and necrotic cell death has been documented in cartilage explants following blunt mechanical trauma. Surface fissuring and cell death are both believed to be pathways in the pathogenesis of osteoarthritis. Our laboratory has previously shown that bovine chondral explants treated with glucosamine (GlcN) will experience less compressive deformation when exposed to a high rate of impact. The current study hypothesized that cartilage explants exposed to GlcN for a period of 6 days will develop an extracellular matrix richer in glycosaminoglycans (GAG). These GAGs will reduce the permeability which will in turn result in hydrostatic pressure generated during loading to help limit distortional strains which have been suggested to be responsible for chondrocyte death. A finite element model, using a transversely isotropic biphasic material model for articular cartilage was developed in order to test the effects of permeability on the stress field in a chondral explant under unconfined compression. The results indicate that GlcN supplementation help reduce the percentage of dead cells in impacted cartilage explants at both a high (~50ms to peak pressure) and low (~1 s to peak pressure) rate of loading. The study also showed, using a finite element model, that a decrease in permeability reduced the area of high Von Mises stresses around a crack during unconfined compression. The current study suggests that pretreating chondral explants with GlcN can help limit impact induced cell death. 70 Introduction: Excessive mechanical loading to a joint has been linked to the long term chronic degradation of articular cartilage (Ewers et al., 2000). In situ studies that examine the acute response of a blunt impact to human knee joints document the appearance of surface lesions on the retro-patellar cartilage resulting fiom pressures of approximately 30 MPa (Atkinson and Haut, 2001). In viva studies using the giant Flemish rabbit also document the appearance of surface fissures on retro-patellar cartilage as the result of a 6 Joule blunt impact to the knee at pressures of approximately 25 MPa (Newberry et al., 1998). This same in viva rabbit model has been used to document significant degradation of articular cartilage 3 years post impact (Ewers et al., 2002). A similar study that looked at the acute response of a 6 Joule impact to the rabbit knee found surface lesions and associated chondrocyte death acutely in the retro-patellar cartilage (Baars et al., 2005). In vitro studies that subject bovine chondral explants to 30 MPa of unconfined compression find excessive surface fissuring and associated chondrocyte death (Ewers et al., 2000; Krueger et al., 2003; Phillips and Haut 2004). In an in viva study significant necrosis of chondrocytes was induced in the patella-femoral joint of rabbits by localized fieezing of the tissue in viva (Simon et el., 1976). After 1 year the affected cartilage shows changes that are consistent with early signs of OA. The tissue changes appear similar to those documented by our laboratory using the impacted joint model (Ewers et al., 2002). These data suggest that the association between blunt impact trauma to a joint and a subsequent degradation of articular cartilage may be the result of impact induced surface damage and associated chondrocyte death. 71 Previous studies correlate the onset of surface fissures with the development of high tensile strains in the collagen (Morel and Quinn 2004, Askew and Mow1978, Eberhardt et al., 1991). One study found that surface fissures in cartilage explants could not be induced when the loading rate was equal or less than the gel diffusion rate suggesting that fluid pressurization generated during loading causes collagen fiber failure (Morel and Quinn, 2004). While chondrocyte death has been documented to occur predominately around surface cracks at high rates (Ewers et al., 2000; Lewis et al., 2003), other studies have found chondrocyte death to occur throughout the thickness without the presence of surface cracks (Millentejivic and Torzilli, 2005; Morel and Quinn, 2004). A study by Ewers et a1. (2000) compared the effects of high and low rates of loading and found chondrocyte death to occur around surface cracks at both rates, but with a greater diffusion of cell death away from cracks in lower rate of loading experiments. It has been suggested that chondrocyte death may be limited in areas away from cracks if fluid flow is prevented, thus increasing the hydrostatic pressure and limit cell deformations during loading (Millentejivic and Torzilli, 2005). These data suggest that an increase in glycosaminoglycan content, and thus an increase in interstitial fluid pressurization and decrease in tissue permeability could potentially limit chondrocyte deformation during loading. Some studies have documented an upregulation of glycosaminoglycan production both in viva (Oegema et a1, 2002) and in cultured human chondrocytes (Dodge and Jimenez 2003) as a result of treatment with the nutraceutical glucosamine (GlcN). A study by this laboratory found that treating bovine chondral explants with GlcN prior to a high rate of blunt impact loading significantly reduced cartilage deformation, matrix 72 damage, and nitric oxide production post impact (Kreuger et al., 2000). A study by Lippiello (2003) found bovine cartilage explants exposed to GlcN experienced a 1000% increase in cell metabolism. In clinical trials GlcN has been shown to significantly reduce the progression of joint space narrowing (Reginster et al., 1999), decrease pain, and improve joint mobility in patients suffering from GA (Camarada and Dowless, 1998; Rovat, 1992; Vaz, 1982; Pujalte et al., 1980). In vitro studies, which chemically induce cartilage degradation via exposure to interleukin-1 (IL-1) or lipopolysaccharide (LPS), have documented significant decreases in the tissue’s subsequent production of degradation enzymes when exposed to GlcN (Fenton et al., 2000; Fenton et al., 2002). These data suggest that GlcN can improve the quality of the cartilage matrix by reducing the presence of degradation enzymes, as a well as promoting the synthesis of tissue glycosaminoglycans. The aim of the current study was to further investigate the effects of preconditioning chondral explants with GlcN prior to a severe blunt impact. The hypothesis was that chondral explants pretreated with GlcN and exposed to injurious unconfined compression will possess a less permeable matrix richer in glycosaminoglycans which will in turn resist fluid movement and result in less chondrocyte death around surface cracks. To test the hypothesis a transversely isotropic biphasic model of cartilage was developed in order to determine whether tissue permeability affects stress fields around the area of a crack in a chondral explant exposed to unconfined compression. Also, bovine chondral explants were exposed to GlcN for a period of six days prior to an unconfined compression of 30 MPa at both at high and low rates. These explants were then studied in regards to their mechanical response under 73 compression, the presence of surface cracks, and chondrocyte death both throughout the thickness as well as number of dead cells adjacent to surface lesions. Methods Experimental Four pairs of bovine forelegs were obtained from a local abattoir within 6 hours of slaughter. The legs were cut just above the knee joint leaving the metacarpal joints intact. The forelegs were cleaned and skinned prior to opening the joint under a laminar flow hood. A 6.35 mm punch with a smooth edge was used to make 120 plugs. Fifteen plugs were taken from each limb. The cartilage plugs were removed from the underlying bone with a scalpel. Sixty explants were placed in individual wells containing Dulbecco’s Modified Eagle’s Medium (DMEM):F12 (Gibco, USA #12500-039) supplemented with 10% fetal bovine serum and antibiotics (penicillin 100 units/ml, streptomycin lug/ml, amphotericin B 0.25 pig/ml), while the remaining sixty specimens were placed in the same media supplemented with 1 mg/ml glucosamine (GlcN) (FCHG49®). After 48 hours of equilibration, all explants were subjected to manual cyclic loads of 0.5 MPa at 1 Hz. Ten consecutive cycles were applied once each day for four days. The media was replaced every 48 hours. Manual cyclic loading was applied with a custom solid stainless steel cylinder placed in series with a 50 pound load cell (model 31/1432: Sensotec, Columbus, OH). The load cell was connected to a computer with an analog to digital converter. LabView (National Instruments Corp, Austin, TX) software was used to monitor the amplitude and frequency of the intermittent loading. After four days of intermittent loading, 56 explants from each group were subjected to unconfined compression. Twenty-eight low- and 28 high-rate of loading 74 experiments were conducted to a peak of 30 MPa in ls or 50ms, respectively, using a haversine impact load function. Each explant was placed between two highly polished stainless steel plates. Prior to the test, the plates were pressed together at 100 N and the location of the machine actuator was recorded. The thickness of each specimen was determined by finding the difference in the actuator’s location at 5 N of load during the test and when the plates were pressed together prior to loading. Each specimen received a 0.5 N preload before being compressed. This protocol has been established in previous experiments from this laboratory (Ewers et al., 2001, Krueger et al., 2003). The load, time, and displacement were recorded during each experiment with an accuracy of 0.1 N, 0.001 s, and 0.01 mm, respectively. Four explants from each media group were not impacted and used for positive and negative controls of cell viability. After the load cycled to approx. 30 MPa each explant was washed three times in media before being returned to pro-assigned wells with one ml of supplemented media and incubated for another 24 hours. Approximately 24 hours after each mechanical compression experiment, each explant was examined for matrix damage and cell viability. The surface of each explant was wiped with India ink. The explants were immediately photographed with a digital camera (Polaroid DMC2, Polaroid Corporation, Waltham, MA, USA) at 25X under a dissection microscope (Wild M5A, Wild Heerbrugg Ltd., Switzerland) to determine the total length of the surface fissures. The total fissure length was calculated with digital imaging software (Sigma Scan, SPSS Inc., Chicago, IL). One observer (AM) digitally measured the length of the surface fissures in each of the photographs. 75 For cell viability, two 0.5-mm slices were taken through the thickness at the center of each explant using a customized cutting tool (Ewers et al., 2001). The sections were cut perpendicular to the preferred orientation of surface fissures. The sections were stained with a kit containing calcein and ethidium bromide homodimer (Live/Dead &Viabi1ity/Cytotoxity, Molecular Probes, Oregon). Positive controls were put through three freeze-thaw cycles in order to kill all of the cells. Each section was viewed under a fluorescence microscope at 100X (Lecia DM LB (frequency: 50-60 Hz), Lecia Mikroskopie and Systeme GmgH, Germany). Full thickness, digital images were taken of a 2.5-mm length at the center of each explant. These images were partitioned into the superficial zone (top 20%), intermediate zone (middle 50%), and deep zone (bottom 30%). The viable (green) and dead (red) cells were manually counted by one observer (SR) using image software (Sigma Scan, SPSS Inc., Chicago, IL). The percent of cell death was computed for each layer. Also, in order to test whether or not potential differences in cell death were a result of differences in surface fissuring, a blinded observer (AM) manually measured the depth of each surface fissure on each section and counted the number of dead cells in the zone adjacent to each fissure. The ratios of dead cells per fissure and per unit fissure depth were documented. Unpaired t-tests were used to compare the total length of surface fissures, peak strain, peak stress, time to peak, cartilage thickness, and strains at given stresses between “with GlcN” and “no GlcN” groups for both high and low rates of loading tests. Two factor ANOVAs were performed to evaluate effects of glucosamine treatment and rates of loading on the percentage of cell death for each cartilage tissue layer (superficial, middle, and deep) as well as for the total overall percentage of cell death. Two factor 76 ANOVAs were also used to evaluate significant differences in the amount of dead cells recorded per fissure as well as per unit depth of fissure. The two factors or independent variables were glucosamine treatment and rate of loading while the dependent variable was percentage of cell death or number of dead cells. All data were reported as mean i one standard deviation. Statistical significance was indicated at p<0.05. Theoretical A transversely isotropic biphasic model of articular cartilage was developed using a commercial finite element software package (Abaqus v 6.1, Abaqus Inc., Providence, Rhode Island). The cartilage material model consisted of six independent material constants including Young’s modulus in the one and three directions (E1 and E3), shear modulus in the one-three plane (613), Poisson’s ratio for the one-three plane (013), and hydraulic permeability in both the one and three directions (kt and k3) (Figure la). For transverse isotropy, material properties in the two direction were set equal to properties in the one direction. Material properties were assigned based on data from indentation testing of rabbit retro-patellar cartilage (Newberry et al., 1998) (Table 1). An axisymmetric model of unconfined compression of a chondral explant was constructed (Figure lb). An arbitrary load of 10 Newtons was applied to a rigid platen in contact with the top surface of the explant. In order to ensure a uniform boundary pressure the platen was not allowed to rotate during loading. Boundary conditions for explant displacement were set to be zero in the axial direction along the bottom side of the explant. Transverse displacement was set at zero along the symmetric axis. Pore pressure was set to zero along the unrestricted edge of the explant in order to allow water flow at the edges. 77 Two separate models of the unconfined compression experiment were created. One model used an intact cartilage explant and the other had a pre-existing crack, in order to evaluate stress fields around a crack during loading (Figure 1c). Validation of the model was performed by compressing an intact explant at two different rates of loading. The high rate of loading experiment required peak load to be reached in 50ms, and a low rate required peak load to be reached at 1 second. Loading was applied using a ramp. A previous study that applied these rates of loading to bovine chondral explants documented a higher overall strain in the low rate of loading experiments (Ewers et al., 2001). In the current study total distance traveled by the platen was documented and regarded as axial deformation and compared between the high and low rate experiments. In order to simulate the effects of GlcN treatment the material properties of the cartilage were altered. The permeability for GlcN treated cartilage was decreased an amount that the permeability of rabbit retro-patellar articular cartilage decreased in a separate study where the rabbits were orally administered GlcN (unpublished data) (Table 1). In this study rabbits were fed GlcN for a period of two months prior to indentation testing of the retro-patellar cartilage. Unconfmed compression experiments were ran at both high and low rates for GlcN treated cartilage with a pre-existing crack. Von Mises stress patterns were documented around the crack for an explant with normal and reduced permeability. Von Mises stress, is used to estimate yield criteria for ductile materials. It is calculated by combining stresses in two or three dimensions, with the result compared to the tensile strength of the material loaded in one dimension. The formula for the Von Mises stress is given below (http://en.wikipedia. org/wiki/Von_Mis es_stress). 78 a. = (on —a.)'2 +(a. —a.)2 +(o-3 —a.)2 2 Where 0‘1, 0'2, and o; are the principal stresses at a point in the material. The current study evaluated Von Mises stresses since it is a measure of the distortional strain energy and distortional strains have been suggested as a potential mechanism for impact induced chondrocyte death (Millentejivic and Torzilli, 2005). Results: Experimental In high rate of loading experiments there was no difference in the peak pressures between the “no GlcN” group (28.5 i 0.5MPa, n=28) and the “with GlcN” group (28.4 i 0.6MPa, n=28). The time to reach peak load for the “with GlcN” group (44 :h 4ms, n=28) was not significantly different from that of the “no GlcN” group (43 i 4ms, n=28) in the high rate experiments. There was, however, a significant difference in peak strain for the “no GlcN” group (55 i 6%, =28) versus the “with GlcN” group (51 :1: 6%, n=28) (p=0.043). There were no significant differences in cartilage explant thickness between the “with GlcN” group (0.68 i 0.1mm, n=28) versus the “no GlcN” group (0.67 n: 0.1mm, n=28). From the data recorded during the high rate tests, nonlinear stress-strain curves were generated for explants from both groups. The “with GlcN” group experienced an increase in explant stiffness by producing a shift in the response curves to the left. There were significant differences in the amount of strain seen by the explant at 20 (p = 0.022), 15 (p = 0.025), 10 (p = 0.029), and 5 (p = 0.033) MPa between the “with GlcN” and “no Glc ” groups (Figure 2). 79 In the low rate of loading experiments there was also no difference in peak pressures between the “no GlcN” group (28.9 :t 0.5MPa, n=28) and the “with GlcN” group (28.8 i 0.5MPa, n=28). The time to reach peak load for the “with GlcN” group (0.89 :1: 0.053, n=28) was not significantly different from that of the “no GlcN” group (0.89 i 0.045, n=28) in the low rate experiments. There were no significant differences seen in the peak strain between the “with GlcN” group (61 i 6%, n=28) versus the “no GlcN” group (59 :1: 6%, n=28). There were no significant differences in cartilage explant thickness between the “with GlcN” group (0.65 :t 0.1mm, n=28) versus the “no GlcN” group (0.66 3: 0.1mm, n=28). The non-linear stress-strain curves were very similar for both groups during low rate loading, and showed no significant differences in strain at 5, 10, 15, or 20 MPa. Supplementation of the culture medium with GlcN significantly reduced the amount of surface fissuring in the “with GlcN” group for explants exposed to high rates of loading (Figure 3). Digital analysis of stained explant photographs indicated that for explants in the “no GlcN” group, the total length of fissuring in the high rate tests was 42.8 :1: 15.3m (n=28). In contrast, for “with GlcN” explants the total length of fissures was 31.5 :t 13.3mm (n=28) (p =3 0.003) (Figure 4). The low rate explants showed no significant differences in total surface fissuring between “with GlcN” (48.3 i 19mm, n=28) versus “no GlcN” (46.4 3: 11mm, n=28). For high rates of loading there was a significant difference (p < 0.001) in the percentage of total cell death in the “with GlcN” (20 i ll %, n=28) group versus “no GlcN” (35 i 17 %, n=28) group of explants (Figure 5a,b). There was also a significant difference (p = 0.017) in the total percentage of cell death in the low rate of loading 80 experiments between the “with GlcN” (25 i 10 %, n=28) and “no GlcN” (34 i 12 %, n=28) groups (Figure 5c,d). Zonal analysis revealed that significantly less cell death was recorded in both the superficial and middle zones in the “with GlcN” groups versus the “no GlcN” groups in both the high and low rates of loading (Figure 6). There were no significant differences found in the percentage of cell death of the deep zone as a result of treatment or rate of loading. The percentage of cell death in the superficial zone for the high rate of loading “with GlcN” group (32 i 16 %, n=28) was significantly less (p < 0.001) than that for the low rate of loading “with GlcN” group (49 i 15 %, =28). Cell death typically appeared adjacent to surface lesions in a butterfly wing-like pattern (Figure 7). There was a significantly greater amount of dead cells per fissure in the “no GlcN” group versus the “with GlcN” group at both high and low rates of loading (Figure 8). There was also a significantly greater amount of dead cells per millimeter fissure depth in the “no GlcN” group versus the “with GlcN” group at both rates of loading (Figure 9). The positive and negative controls showed virtually all dead and all live cells, respectively (Figure 10). Theoretical The total axial deformation of the explant in the finite- element, unconfined compression model was greater for a low rate versus a high rate of loading. The peak axial deformation for the low rate experiment was 6.354 x 10'5 m and 6.554 x 10'5 m for the high rate. These data display a rate effect on tissue stiffness that is consistent with previous studies (Ewers et al., 2001; Krueger et al., 2003). The pattern of Von Mises stress gradients around the crack tip were similar to the patterns of chondrocyte death observed in explants around surface lesions (Figure 11). For the high rate of loading 81 simulation the explant with a decreased permeability, or rather the ‘with GlcN’ explant, experienced a higher range of Von Mises stresses (0.259 — 0.766 MPa) versus the explant with normal permeability (0.257 — 0.741 Wa). However, the stress field was different between explants with different permeabilities. Thee explant with with a smaller permeability experienced higher Von Misses stresses concentrated locally around the crack as to where the normal permeability explant experienced a more diffuse pattern of high Von Mises stress around the crack (Figure 11). Similar trends were observed for the low rate of loading simulation, however, the differences in both stress range and pattern were not as strong (Figure 12). Discussion: This study confirmed some of the results of a previous study (Krueger et al., 2000), showing that supplementation of the culture medium with glucosamine reduced peak explant strain and the degree of matrix damage during a high rate of unconfined compression. The current study was able to show these results with a concentration of 1 mg/mL versus 2.5 mg/mL, which was used in the previous study (Krueger et al., 2000). The current study documented stiffening in the high rate of loading ‘With GlcN” explants as observed by a shifting to the left of the stress versus strain response curve (Figure 2). A separate study found the compressive stiffness of articular cartilage to correlate with glycosaminoglycan (GAG) content as determined by magnetic resonance imaging (MRI) techniques (Kurkijarvi et al., 2004). A previous study, which measured the dynamic shear modulus of ovine articular cartilage found it to correlate with cartilage thickness, collagen organization, and GAG content (Oakley et al., 2004). Previous studies have documented an upregulation of GAGs in cartilage exposed to GlcN both in viva and in 82 vitro (Lippiello et al., 2000; Lippiello 2003).. These data, as well as data from the previously mentioned studies, suggest that GlcN treatment may have had an affect on proteoglycan content and/or collagen organization that in turn affected the mechanical response of explants during a high rate of loading. A limitation of the current study was that collagen organization and proteoglycan content was not quantified. Further studies are needed to examine the affects of GlcN treatment on collagen organization and proteoglycan content. The current study reported a significant reduction in the average total length of surface fissures in GlcN treated explants exposed to a high rate of loading. This phenomenon was not, however, documented in explants exposed to a lower rate of loading. Previous studies correlate the onset of surface fissures with the development of high tensile strains in the collagen fibril network (Morel and Quinn 2004; Askew and Maw, 1978; Eberhardt et al., 1991). Theoretical models of in situ cartilage loading scenarios suggest that a potential mechanism for surface fissuring could be the development of high tensile strains near the surface of cartilage during loading (Li et al., 1995; Atkinson et al., 1998). It has been documented that the development of surface fissures occurs at fast rates of loading thatdon’t allow for fluid movement through the semi-permeable solid matrix (Morel and Quinn, 2004). This inability of fluid to escape fiom the matrix results in interstitial fluid pressurization generated during loading that exceeds the restraining capacity of the collagen network (Pins et al., 1995). The maximum compressive strain was less, on average, in the high rate of loading “with Glc ” group of explants suggesting that the in-plane strain, based on the Poisson affect, was also less. There were no significant differences in maximum compressive strain 83 documented in the low rate of loading explants between the “with GlcN” versus the “no GlcN” explants, possibly helping to explain no significant differences in average surface fissure length between groups. The data from the current study suggest that the reduction in surface fissuring in the high rate of loading “with GlcN” group may have resulted from the stiffened response as documented by a shift to the left of the stress versus strain curve (Figure 2). A significantly less amount of total cell death was documented in explants pretreated with GlcN for both high and low rates of loading tests. These findings are in contrast with those from a previous study which did not document any significant differences in cell death between GlcN treated and untreated explants exposed to a blunt impact (Krueger et al., 2000). Potential causes for the contrast in data may have been fi'om differences in the GlcN pretreatment incubation period, GlcN concentration, and the introduction of intermittent cyclic loading prior to impact. The previous study exposed explants to a concentration of 2.5 mg/ml of GlcN, as opposed to the current study which used 1 mg/mL GlcN. Recent studies have found that high concentrations of GlcN have cytotoxic effects at high concentrations (Mello et al., 2004). Furthermore, the previous study pretreated explants for a period of 48 hours, as opposed to the current study which used a 6 day incubation period. Other in vitro studies that have shown an increase in cartilage matrix synthesis as a result of GlcN treatment have used incubation periods as long as 10 days (Lippiello, 2003; Fenton et al., 2004). Also, the introduction of intermittent low level cyclic loading throughout the incubation period may have also contributed to the differences in cell viability in the current versus the previous study. In a separate study that examined the effects of a mild surfactant on chondrocyte viability 84 low level cyclic loading after treatment was necessary in order for the treatment to be effective. The protocol was suggested to have allowed for the surfactant to be ‘pumped’ into the tissue (Phillips and Haut, 2004). A separate study found a 1000% increase in cartilage metabolism in GlcN treated explants exposed to 24 hours of static stress, as opposed to a 40% increase in those that were not stressed. The data from the current study suggests that GlcN dosage concentration, pretreatment time duration, and mechanical stress may play a role in GlcN’s ability to prevent chondrocyte death as a result of a blunt impact in vitro. The increased benefits of GlcN on traumatized cartilage in explants that were exposed to low level cyclic loading prior to injury may also suggest co-benefits to the cartilage that may arise fiom exercise in combination with GlcN in an in viva situation. A reduction in impact induced surface cracks, as a result of less compressive strain in the GlcN treated, high rate of loading group of explants may have been responsible for the reduction in chondrocyte death in the superficial zone. This agrees with previous studies that document the presence of cell death primarily around surface cracks in explant studies (Lewis et al., 2003; Ewers et al., 2000). However, there was also a significant decrease in chondrocyte death in the “with GlcN” versus “no GlcN” explants for low rates of loading. In order to eliminate the effects of surface fissuring on chondrocyte viability the number of dead cells per surface fissure, and dead cells per unit depth of surface fissure were quantified. Results from this data showed a significant reduction in dead cells per fissure and dead cells per unit fissure depth in GlcN supplemented groups for both rates of loading versus their respective “no GlcN” groups. While the mechanism for the action of GlcN on cell death near surface cracks is yet 85 unclear, upregulation of tissue PGs in the pericellular matrix (Quinn et al., 1998) may help limit shearing strains in chondrocytes near the cracks. The effect could, as well, be that GlcN increases the yield stress and reduces post-yield strains in articular cartilage. Both of these alterations can reduce the zone of plastic strain adjacent to surface fissures (Mendelson 1968). The zone of plastic strain around a crack tip resembles the zone of dead cells in an impacted cartilage explant (Figure 13). The data from the current study indicate that pro-treatment with GlcN can help reduce the zone of plastic strain perhaps by increasing the yield strength of the tissue. Previous studies have documented cell death to occur predominately around surface cracks when loaded at a high rate (Lewis et al., 2003; Ewers et al., 2001) and away fi'om cracks and further into the depth of the cartilage when loaded at slower rates (Morel and Quinn 2004). It has been suggested that large distortional strains developed in the cartilage matrix during loading may contribute to acute cell death (Milentejivic and Torzilli, 2005). The same study suggested that hydrostatic pressure developed during loading could help limit cell deformations. A decrease in permeability, to model GlcN treatment, was shown to reduce the zone of high Von Mises stresses around a crack in the theoretical model of unconfined compression. These data may suggest a mechanism whereby GlcN treatment may help limit the potential for developing a post-traumatic osteoarthritis by decreasing cartilage permeability, perhaps as a result of an up-regulation of tissue GAGs, to increase a hydrostatic pressure effect in the cartilage during impact loading and limit chondrocyte death. More studies need to be performed in order to validate this hypothesis. Further experimental and theoretical analyses are needed to better explain the chondroprotective effect of glucosamine for blunt impact situations. 86 In conclusion, in the current study bathing cartilage explants in GlcN supplemented culture media prior to a blunt impact had some significant beneficial effects. Impact induced chondrocyte death was found to be significantly reduced in explants that were pretreated with GlcN. This reduction in cell death was found to occur predominately around surface lesions with fewer dead cells per lesion in the GlcN supplemented explants. However, the mechanism by which GlcN was able to reduce cell death around these traumatized areas cannot be deduced based on the collected data from the current study. Future investigations should explore potential mechanisms of this phenomenon. Acknowledgement: This research was supported by a grant from the Cendter for Disease Control (R49/ CCR503607). Nutramax Laboratories, Inc., Edgewood, MD provided the supplement. The Authors would like to thank Clifford Beckett for his technical assistance. 87 References Askew M, Mow V: The biomechanical function of the collagen fibril ultrastructure of . articular cartilage. Journal of Biomechanical Engineering 100:105-115, 1978 Atkinson TS, Haut RC, Altiero NJ: An investigation of biphasic failure criteria for impact-induced fissuring of articular cartilage. Journal of Biomechanical Engineering 120:536-537, 1998 Atkinson PJ, Haut RC: Injuries produced by blunt trauma to the human patello femoral joint vary with flexion angle of the knee. Journal of Orthopaedic Research 19, 827-833, 2001 Baars D, Phillips DM, Haut RC: Repair of damaged chondrocytes in the in viva traumatized joint. Transactions of the Slst Annual meeting of Orthopaedic Research Society 2005 Camara, CC, Dowless GV: Glucosamine sulfate for osteoarthritis. Annals of Pharmacotherapeutics 32(5), 580-587, 1998 Dodge GR, Jimenez SA: Glucosamine sulfate modulates the levels of aggrecan and matrix metalloproteinase-3 synthesized by cultured human osteoarthritis articular chondrocytes. Osteoarthritis and Cartilage 11(6): 424-432, 2003 Eberhardt AW, Lewis JL, Keer LM: Normal contact of elastic spheres with two elastic layers as a model of joint articulation. Journal of Biomechanical Engineering 113, 410-417, 1991 Ewers B, Dvoracek-Driksna D, Orth M, Altiero N, Haut R: Matrix damage and chondrocyte death in articular cartilage depends upon loading rate. Transactions of the 46th Annual meeting of Orthopaedic Research Society. 2000 Ewers BJ, Dvoracek-Driksna D, Orth MW, Haut RC: The extent of matrix damage and chondrocyte death in mechanically traumatized articular cartilage explants depends on rate of loading. Journal of Orthopaedic Research 19:779-784, 2001 Ewers BJ, Weaver BT, Sevensma ET, Haut RC: Chronic changes in rabbit retro-patellar cartilage and subchondral bone after blunt impact loading of the patello femoral joint. Journal of Orthopaedic Research 20:545-550, 2002 http://en.wikipedia. org/wiki/Von_Mises_stress Krueger J, Dvoracek-Driksna D, Ewers B, Haut R, Orth M. Effects of pretreatment with glucosamine on mechanically traumatized cartilage explants. 47, 257. 2001. Transactions of the 47th Annual Meeting of the Orthopaedic Research Society. 2001. 88 Krueger JA, Thisse P, Ewers BJ, Dvoracek-Driksna D, Orth MW, Haut RC: The extent and distribution of cell death and matrix damage in impacted chondral explants varies with the presence of underlying bone. Journal of Biomechanical Engineering 125(1):]14-119, 2003 Kurkijarvi JE, Nissi MJ, Kiviranta I, Jurvelin JS, Nieminen MT: Delayed gadolinium- . enhanced MRI of cartilage (dGEMRIC) and T2 characteristics of human knee articular cartilage: topographical variation and relationships to mechanical properties. Magnetic Resonance in Medicine 52(1):41-6, 2004 Lewis JL, Deloria LB, Oyen-Tiesma M, Thompson RC, Ericson M, Oegema TR: Cell death after cartilage impact occurs around matrix cracks. Journal of Orthopaedic Research 21 :881-887, 2003 Li X, Haut RC, Altiero NJ: An analytic model to study blunt impact response of the rabbit P-F joint. Journal of Biomechanical Engineering 117:485-491, 1995 Lippiello L, Han MS. Dose and stress response characteristics of articular cartilage chondrocytes to glucosamine and chondroitin sulfate. 49th Annual Meeting of the Orthopaedic Research Society, 2003 Lippiello L, Woodward J, Karpman R, Hammad T: In vivo chondroprotection and metabolic synergy of glucosamine and chondroitin sulfate. Clinical Orthopaedics and Related Research 381:2292240, 2000 Mello DM, Nielsen BD, Peters TL, Caron JP, Orth MW: Comparison of inhibitory effects of glucosamine and mannosamine on bovine articular cartilage degradation in vitro. American Journal of Veterinary Research 65(10): 1440-5, 2004 Milentijevic, D., Helfet, D.L., Torzilli, RA, 2005. Influence of stress rate on water loss, matrix deformation and chondrocyte viability in impacted articular cartilage. Journal of Biomechanics 38, 493-502. Morel V, Quinn TM: Short-term changes in cell and matrix damage following mechanical injury of articular cartilage explants and modeling of microphysical mediators. Biorheology 41 (3-4):509-5 l 9, 2004 Newberry, W.N., Garcia, J.J., Mackenzie, C.D., Decamp, C.E., Haut, RC, 1998. Analysis of acute mechanical insult in an animal model of post-traumatic osteoarthrosis. Journal of Biomechanical Engineering 120, 704-709. Oakley SP, Lassere MN, Portek I, Szomor Z, Ghosh P, Kirkham BW, Murrell GA, Wulf S, Appleyard RC: Biomechanical, histologic and macroscopic assessment of articular cartilage in a sheep model of osteoarthritis. Osteoarthritis and Cartilage. 12(8):667-79, 2004 89 Oegema TR, Deloria LB, Sandy JD, Hart DA: Effect of oral glucosamine on cartilage and meniscus in normal and chymopapain-injected knees of young rabbits. Arthritis and Rheumatism 46(9):2495-2503, 2002 Phillips DM, Haut RC: The use of a non-ionic surfactant (P188) to save chondrocytes from necrosis following impact loading of chondral explants. Journal of Orthopaedic Research 22:1135-42, 2004 Pins GD, Huang EK, Christiansen DL, Silver FH: Effects of static axial strain on the tensile properties and failure mechanisms of self-assembled collagen fibers. Journal of Applied Polymer Science 63: 1429—40, 1995 Pujalte JM, Llavore EP, Ylescupidez FR: Double-blind clinical evaluation of oral glucaosamine sulphate in the basic treatment of osteoarthrosis. Current Medical Research and Opinion 7:110-114, 1980 Quinn TM, Grodzinsky AJ, Hunziker EB, Sandy JD: Effects of injurious compression on matrix turnover around individual cells in calf articular cartilage explants. Journal of Orthopaedic Research 16:490-499, 1998 Reginster J, Dacre J, Rovati L, Gosset C: Glucosamine sulfate significantly reduces progression of knee osteoarthritis over 3 years: A large randomized, placebo- controlled, double-blind, prospective trial. Osteoarthritis and Cartilage 8(2): 151- 152, 2001 Rovati LC: Clinical research in osteoarthritis: Design and results of short-term and long- term trials with disease-modifying drugs. International Journal of Tissue Reaction 14:243-251, 1992 Simon WH, Richardson S, Herman W, Parsons JR, Lane J: Long-term effects of chondrocyte death on rabbit articular cartilage in vivo. Journal of Bone and Joint Surgery 58(4):517-526, 1976 Vaz AL: Double-blind clinical evaluation of the relative efficacy of ibuprofen and glucosamine sulphate in the management of osteoarthrosis of the knee in out- patients. Crurent Medical Research and Opinion 8: 145-149, 1982 90 Tables Table 1. Transversely isotropic and biphasic material properties values for the theoretical cartilage material model “no GlcN” “with GlcN” E1(MPa) 7.835 7.835 133 (MPa) 1.53 1.53 R] (*10'13 ml) 5.95 3.149 R3 (*10'13 m2) 1.222 0.858 01 0.5 0.5 U3 0 O 6,, (MPa) 0 0 91 Figure Captions Figure 1. An axisymmetric finite element model of an unconfined compression of a chondral explant was developed. Cartilage was modeled as being transversely isotropic biphasic (a). Two separate geometries were used; one was an intact cartilage explant (b) while the other was cartilage with a pre-existing crack(c) Figure 2. Stress vs. Strain response curves for high rate of loading tests for both the “with GlcN” and “no Glc ” groups. Stress and strain data was collected during unconfined compression tests. Stars indicate statistically significant different amounts of strain at the same pressure (p<0.05 Figure 3. Gross photographs were used to determine the average total fissure length for cartilage explants following blunt impacts. In the high rate of loading groups, less fissuring was documented in GlcN treated explants (3) compared to non-treated specimens (b). Figure 4. The total length of surface fissuring was quantified using digital imaging software. Bar graphs of average values of total fissure length both high rate and low rate tests were made. Error bars represent the standard deviation. “*” indicates statistically significant (p<0.05) difference from corresponding “no GlcN” group Figure 5. Explants cross sections were stained with calcein and ethidium homodimer. Photographs were taken at 100x with a digital camera attached to a flouresence microscope. Viable cells are indicated via a green stain and dead cells a red stain. (a) low 92 rate “with GlcN” (b) high rate “with GlcN” (c) low rate “no Glc ” ((1) high rate “no GlcN” Figure 6. Cross section chondrocyte viability photographs were divided into superficial, middle, and deep zones. Amount of dead and live cells were manually counted in each zone using digital imaging software. Percentages of chondrocyte death were placed in bar charts for both (a) high rate and (b) low rate of loading tests. A * indicates a statistical difference versus respective ‘No GlcN’ group. A ** indicates a statistical difference in % cell death between rates of loading. Figure 7. Fluorescence photographs of explant cross sections reveal a greater area of chondrocyte death around impact induced cracks for both rates of loading. (a) “No GlcN” high rate, (b) “With GlcN” high rate (c) “No GlcN” low rate ((1) “With G1cN”Low rate. Figure 8. The number of dead cells around individual fissures were manually counted, recorded, and averaged for each group. “*” indicates a statistical difference from corresponding “no GlcN” groups. Figure 9. The number of dead cells around individual fissures were manually counted and divided by the depth of the fissure in millimeters. “*” indicates a statistical difference from corresponding “no GlcN” groups. Figure 10. Half of the un-impacted explants were put through three freeze-thaw cycled and revealed virtually all dead cells (a). Other un-impacted explants showed virtually no cell death (b). 93 Figure 11. The Van Mises stress patems were examined after a high rate of load application to a cracked explant in a finite element model of unconfined compression. GlcN modeled cartilage was given a lower permeability based on a hypothesized increase in glycosaminoglycan content. The area of increased Von Mises stresses appeared to be greater in the “No GlcN” modeled cartilage versus the “with GlcN” modeled cartilage. Figure 12. The Von Mises stress patterns were examined after a low rate of load application to a cracked explant in a finite element model of unconfined compression. There did not appear to be any differences in Van Mises stress patterns between the ‘with GlcN’ and ‘no GlcN’ groups. Figure 13. (a) The area of plastic strain around a crack tip resembles the shape of a butterfly wing. (b) The pattern of plastic strain around a crack superimposed over a fluorescence image of a cartilage explant cross section exposed to 30 MPa of unconfined compression. The pattern of cell death is falls within the bounds of the plastic strain. 94 Figures «I. «<1. Figure l 95 Stress (MPa) 25 +"With GlcN" High 20 - Rate . * —I—"No GlcN" ngh Rate 15 — 10 - 5 _ O I l l l l T 0 0.1 0.2 0.3 0.4 0.5 0.6 Strain (mm/mm) Figure 2 96 Surface Fissure Length (mm) 80 7O 60 50 40 30 20 1O 7L5, Rate WW-W—thighRate With GlcN No GlcN Figure 5 97 80% 70% 60% 50% 40% “/o Cell Death 30% 20% 10% 0% 80% 70% 60% 50% % Cell Death on A O O o\° e\° 20% 1 0% 0% No GlcN No GlcN (b) Figure 6 98 Superficial I Middle ll Deep I Total With GlcN Superficial I Middle I Deep I Total GlcN Figure 7 99 i in ll High Rate it! Low Rate l_ 180 , 160 m=oo been no Lon—:32 No GlcN With GlcN Figure 8 :1 64 4‘ I High Rat LolRat i N l l l *1 _l E, i l * I .. i O O O O 0 I I c _ fl 0 O O O 0 0 O 0 O m 2 O 8 6 4 2 1 1 1 EE can m=oo been co con—:32 No GlcN With GlcN Figure 9 100 Figure 10 Figure 11 101 Figure 12 Crack Tip ‘t ‘ed (3.2 - 4:} ('3‘ Figure 13 102 Conclusions and Recommendations for Future Work The previous chapters describe the results of an injurious blunt impact load to articular cartilage in terms of matrix damage and chondrocyte viability using both in vitro and in viva models. Potential therapeutic treatments that can help limit cell death, either by membrane rescaling or matrix strengthening were studied. In Chapter 1 ‘tissue equilibration’ of chondral explants prior to an injurious unconfined compression experiment was found to result in a greater overall length of surface fissuring versus explants that were impacted immediately after removal from the joint. This corresponded to a greater percentage of cell death in the superficial zone for the equilibrated explants. In contrast, the percentage of chondrocyte death in the deep zone of non-equilibrated explants was significantly greater then for equilibrated explants. It was hypothesized that increased fluid present in the equilibrated explants resulted in higher deep zone hydrostatic pressures generated during loading, which in turn limited cell deformation. Future studies should consider the presence of underlying bone, and determine the effect on deep zone cell death. Additionally, it was observed that correlations between stiffness and various parameters, indicative of water and solid matrix content, were diminished when explants were allowed to equilibrate. Future studies should include biochemical assays in order to determine correlations between mechanical characteristics and various matrix constituents, and whether or not these correlations are also diminished as a result of equilibration. Future studies that wish to compare theoretical models with experimental data should be aware of the effects that ‘tissue equilibration’ might have on the mechanical properties of the tissue. 103 Chapter 2 described a study on rabbit patello-femoral joints subjected to a 6 Joule intensity impact. The major findings of this study were the presence of acutely necrotic cells in the retro-patellar cartilage and a reduction in the percentage of these dead cells with the administration of poloxamer 188 (P188) directly into the joint immediately after impact. The methods used in this chapter did not account for the long term viability of cells ‘saved’ by P188. The possibility that these cells, which were saved within the first 4 days after impact, may not function properly or die by apoptosis at a later time needs further investigation. Future studies should acknowledge this possibility and document chondrocyte viability over a longer time period. Methods for determining apoptosis, such as TUNEL staining, should also be included in future studies. Also, a small sample size (n=6) was used in this study, which resulted in a lack of statistical power in numerous analyses. Unfortunately, biologically meaningful differences may have existed, but they were unable to be detected due to the lack of statistical power in the current study. Furthermore, two rabbits in the ‘4-day poloxamer’ group had very high base-line levels of surface fissuring, resulting in average total fissure lengths approximately twice as great as that documented in the other treatment groups. This compromised the current database. The concentration of P188 and the amount of exercise of the rabbit were chosen arbitrarily and should be studied more specifically in firture studies. The ability of the P188 to penetrate into the deep zone of the tissue was not recorded in the current study. Future studies may wish to ‘tag’ the P188 and track the presence of it throughout the thickness of the cartilage. In Chapter 3 the effects of bathing chondral explants in culture media supplemented with the nutraceutical glucosamine (GlcN) were examined. This study 104 found that explants treated with GlcN experienced less compressive strain when impacted at a high rate. This limited the amount of surface fissuring. The study also found that the percentage of chondrocyte death was reduced in the group of explants treated with GlcN, in both the superficial and middle zones of the tissue. This affect appeared due to a smaller dispersion of dead cells in the vicinity of impact induced cracks with GlcN supplementation of the medium. A transversely isotropic, biphasic theoretical model of the cartilage explant exposed to unconfined compression found the area of high Von Mises stresses around a crack reduced if the permeability of the tissue was reduced. My study was limited by the fact that no independent measure of permeability was performed to determine if GlcN supplementation, indeed, decreased tissue permeability. Future studies should incorporate mechanical testing of explants both treated and not treated with GlcN to deduce hydraulic permeability. Furthermore, it was suggested that GAGs produced by chondrocytes as the result of GlcN treatment might have been deposited in the pericellular matrix, in turn having a protective effect on the cells. However, in a companion study no significant increase in GAG was documented in explants treated with the same concentration of GlcN used in the current study over a 6 day period. Future studies may wish to examine GAG contents in the immediate vicinity of the chondrocyte, as there may be a rather small, localized effect yielding these results. Also, the current study exposed all explants to manual cyclic loading during pre-treatment with GlcN. Future studies should firrther examine the potentially synergistic affects of cyclic mechanical loading and GlcN treatment on cartilage explants. While the current studies find effects of tissue equilibration, P188, and GlcN pre- treatment on the injury response of articular cartilage, they do not firlly explain their 105 mechanisms. Future studies should incorporate other experimental methods such as biochemical assays, molecular magnetic resonance imaging techniques, and apoptotic cell death detection methods to firrther investigate these findings. 106 Appendix A: Chapter 1 Raw Data Lopd Controlled Egperiments Table 1. Mechanical data for the non-equilbrated explants Time to Peak Peak Specimen Peak Stress Strain Thickness p. or [81 IMPal ImMI 1 0.041 29.55 0.5932 0.70 2.723 -2.408 2 0.033 29.13 ' 0.5798 0.59 2.258 -2.5 3 0.041 29.00 0.4294 0.67 5.333 -3.632 4 0.042 29.24 0.5203 0.70 3.422 -3.08 5 0.04 29.15 0.6268 0.66 2.078 -2.313 6 0.041 28.60 0.5427 0.51 3.336 -2.672 7 0.035 28.84 0.5346 0.56 3.813 -2.318 8 0.04 29.05 0.5475 0.73 2.403 -3.492 9 0.037 28.82 0.5532 0.73 1.816 -3.853 10 0.042 28.73 0.5376 0.66 3.57 -2.668 1 1 0.029 28.88 0.4994 0.50 3.149 -2.749 12 0.025 29.48 0.5595 0.51 2.831 -2.462 13 0.038 29.06 0.5755 0.40 3.579 -1.779 14 0.026 28.85 0.6334 0.42 1.794 -1.803 15 0.036 28.91 0.6703 0.45 1.815 -1.532 16 0.034 28.75 0.6622 0.50 1.963 -1.418 17 0.021 29.36 0.5386 0.40 3.336 -2.313 18 0.013 29.96 0.5931 0.46 2.002 -2.218 19 0.025 29.10 0.6654 0.40 1.614 -1.816 20 0.026 29.1 1 0.4423 0.64 4.274 -3.68 Average 0.03325 29.08 0.5652 0.56478597 2.85545 -2.5353 Std. Dev. 0.00824541 3 0.323 0.0665 0.1 1 829986 0.9870341 0.71 4273 107 Table 2. Mechanical data fro the equilibrated explants Time to Peak Specimen Peak Stress Peak Strain Thickness p a [S] [MPal me} 21 0.028 28.76 0.483135075 0.6860243? 3.696 -3.457 22 0.03 28.39 0.636599281 0.591 1 175 2.086 -1.746 23 0.024 29.1 1 0.551089931 0.63070628 4.124 -1.897 24 0.032 28.13 0.597155084 0.51655235 2.624 -1.874 25 0.028 28.38 0.573146548 0.74151165 2.013 -3.17 26 0.028 28.60 0.659859868 0.68541583 1 .289 -2.638 27 0.019 27.96 0.652667409 0.82900104 1 .572 -1.968 28 0.026 28.08 0.525510225 0.60141453 3.071 -2.957 29 0.029 27.86 0.602277073 0.61 1 17848 1.817 -3.04 30 0.018 29.04 0.6428061? 0.6410789 2.53 -0.896 31 0.028 28.01 0.585485859 0.4954436 2.228 -2.517 32 0.028 28.83 0.6137 0.43598885 2.126 -1.775 33 0.022 29.76 0.6684 0.5751 1755 1.905 -0.866 34 0.017 29.46 0.6765 0.44708819 0.813 -2.015 35 0.026 29.63 0.731 1 0.6264876 1.329 -0.822 36 0.023 28.59 0.7135 0.62121262 0.863 -1.418 37 0.026 29.16 0.5268 0.61 10233 2.292 -2.946 38 0.016 30.43 0.6321 0.66077322 0.554 -2.745 39 0.027 29.35 0.6692 0.53320697 1 .927 -1 .032 40 0.025 29.52 0.6554 0.45653088 1.277 -1.916 Average 0.025 28.84 0.6198221 44 0.59984369 2.0068 -2.08475 Std. Dev. 0.004507304 0.701 0.06459351 1 0.09968538 0.91 38422 0.820396 108 Table 3. Equilibrated explant weights and fluid gain Time 0 Time 24 Specimen Weight Weight Fluid Gain Lg'1OA-5L [g‘10"-5] 21 2090 2253 7.80% 22 1420 1650 16.20% 23 1768 1834 3.73% 24 1418 1603 13.05% 25 2106 2401 14.01% 26 1504 1767 17.49% 27 2116 2374 12.19% 28 1850 1896 2.49% 29 1523 1623 6.57% 30 1762 2105 19.47% 31 1453 1620 11.49% 32 1519 1660 9.28% 33 2028 2479 22.24% 34 1445 1720 19.03% 35 1885 2294 21 .70% 36 1966 2266 1 5.26% 37 2021 2246 11.13% 38 2236 2647 18.38% 39 1749 1930 10.35% 40 1666 1921 15.31% Average 1776.25 2014.45 13.36% Std. Dev. 271 .215757 332.5921281 5.62% 109 Table 4. Surface fissure data in pixels and millimeters for non-equilibrated (specimens 1-20) and equilibrated (specimens 21-40). Matrix Matrix Matrix Matrix Specimen Damage Damagg_ Specimen Damage Damage Pixels mm Pixels mm 1 7071 75.35 21 4209 44.85612789 2 5573 59.39 22 5229 55.72646536 3 2604 27.75 23 3308 35.25399645 4 5383 57.36 24 5320 56.69626998 5 5528 58.91 25 7266 77.43516874 6 3190 33.99 26 5386 57.39964476 7 6983 74.41 27 7098 75.64476021 8 3707 39.50 28 41 1 8 43.88632327 9 5137 54.74 29 4179 44.53641208 10 6129 65.31 30 6290 67.03374778 11 1184 12.61 31 3432 36.57548845 12 5203 55.44 32 5360 57.12255773 13 3387 36.09 33 6496 69.22912966 14 3285 35.00 34 6351 67.68383659 15 3964 42.24 35 6005 63.9964476 16 3721 39.65 36 8105 86.37655417 17 2881 30.70 37 2786 29.69094139 18 3350 35.70 38 6609 70.43339254 19 5016 53.45 39 4759 50.71758437 20 1299 13.843 40 5711 60.86323268 Average 4229.75 45.077 Average 5400 57.55790409 Std. Dev. 1667.918964 17.775 Std. Dev. 1425 15.18769892 110 Table 5. Cell viability data for non-equilibrated (specimens 1-20) and equilibrated (specimens 21-40). The percentage of dead cells was recorded in each zone. % % Specimen % Dead % Dead Dead Dead Super Middle Deep ALL 1 12.24% 2.70% 5. 56% 5. 61 % 2 13.22% 7.20% 14.04% 10.70% 3 22.64% 4.40% 17.76% 12.03% 4 24.76% 8.31% 15.02% 14.37% 5 37.93% 30. 38% 52.73% 37.56% 6 30.34% 12.20% 6.76% 13. 74% 7 18.12% 10.33% 53.68% 20. 08% 8 23.47% 23.43% 24.30% 23.68% 9 23.0870 4.54% 24.04°/o 12.82% 10 26.57% 13.37% 31.17% 19.97% 1 1 34.95% 6.56% 44.86% 22.03% 12 16.00% 5.12% 5.93% 7.39% 13 23.01% 21.97% 52.94% 29.05% 14 39.08% 20.09% 37.37% 28.10% 15 21.80% 11.93% 14.81% 15.90% 1 6 43. 88% 1 6.74% 24. 19% 24.83% 17 16.48% 3.85% 11.61% 8.21% 18 44. 55% 29.21% 52 .94% 36.47% 19 47.92% 1 3. 86% 7.97% 18.44% 20 24.53% 3.13% 1 1 .88% 9.40% Average 27.23% 12.47% 25.48% 1 8.52% Std. Dev. 1 0.73% 8.63% 17.51% 9.27% 111 % % % Specimen Dead % Dead Dead Dead Super Middle Deep All 21 25.96% 1 1 .45% 24.00% 16.99% 22 35.06% 12.73% 12.00% 15.99% 23 46.58% 14.51% 30.87% 26.01% 24 22.34% 8.30% 7.76% 1 1 . 16% 25 30.61% 8.12% 13.73% 13.56% 26 29.90% 12.33% 19.05% 1 7.35% 27 31 .30% 5.60% 7.1 1% 10.55% 28 19.44% 2.28% 10.71% 8.66% 29 36.36% 18.75% 32.29% 26.80% 30 49.37% 1 .63% 2.26% 9.38% 31 61.25% 25.21% 5.71% 25.55% 32 48.72% 13.07% 4.26% 14.89% 33 36.36% 20.96% 5.41 % 16.33% 34 29.82% 3. 39% 0.00% 9.42% 35 36.07% 10.31% 5.41% 12.41% 36 26.92% 22.91% 5. 79% 16.87% 37 34.85% 21 .36% 30.51% 26.24% 38 67. 1 6% 1 6.03% 4.90% 21 .35% 39 27.50% 26.92% 26.09% 26.74% 40 29.10% 6.12% 5.93% 12.27% Ave rage 38.23% 1 3.1 0% 12.89% 1 8.93% Std. Dev. 1 2.47% 7.85% 1 0.49% 8.37% 112 Displacement Controlled Experiments Table 6. Mechanical and weight data for all un-equilibrated explants Time to Peak Peak Initial Dry Percent Specimen Peak Stress Strain Thickness Weight Weight Water Is] [MPa] [mm] {91 L9] 1 0.048 60.607 0.608 0.801 0.02581 0.00864 66.52% 2 0.046 30.466 0.618 0.549 0.01753 0.00507 71.08% 3 0.046 49.009 0.567 0.635 0.02017 0.00642 68.17% 4 0.046 40.033 0.626 0.684 0.01807 0.00574 68.23% 5 0.045 27.980 0.624 0.597 0.01803 0.00488 72.93% 6 0.044 41.043 0.621 0.704 0.02300 0.00647 71.87% 7 0.046 28.458 0.573 0.488 0.01556 0.00444 71.47% 8 0.046 31.563 0.540 0.485 0.01509 0.00405 73.16% 9 0.044 33.857 0.644 0.685 0.0221 0.00631 71.45% 10 0.046 34.799 0.626 0.612 0.0217 0.00599 72.40% 11 0.043 33.752 0.584 0.548 0.01892 0.0061 67.76% 12 0.044 33.062 0.600 0.587 0.01871 0.00544 70.92% 13 0.043 37.060 0.605 0.660 0.01952 0.00584 70.08% 14 0.048 36.487 0.645 0.742 0.02462 0.00728 70.43% 15 0.045 23.802 0.613 0.514 0.01275 0.00353 72.31% 16 0.045 23.391 0.552 0.422 0.01195 0.00326 72.72% 17 0.044 66.081 0.535 0.653 0.02118 0.00745 64.83% 18 0.045 72.917 0.629 0.923 0.02784 0.01085 61.03% 19 0.046 73.336 0.578 0.815 0.02370 0.00909 61.65% 20 0.044 66.630 0.590 0.813 0.02357 0.00945 59.91% 21 0.046 61.414 0.548 0.675 0.02118 0.00702 66.86% 22 0.044 50.341 0.602 0.724 0.02001 0.00799 60.07% 23 0.048 57.210 0.568 0.709 0.02328 0.00630 72.94% 24 0.047 42.209 0.589 0.620 0.01876 0.00631 66.36% 25 0.045 34.214 0.563 0.521 0.01529 0.00467 69.46% 26 0.045 35.504 0.566 0.555 0.01676 0.00561 66.53% 27 0.042 38.019 0.537 0.526 0.01561 0.00521 66.62% 28 0.044 33.159 0.571 0.567 0.01764 0.00615 65.14% 29 0.046 44.463 0.570 0.678 0.02060 0.00676 67.18% 30 0.045 45.197 0.568 0.678 0.02261 0.00681 69.88% 31 0.044 38.309 0.607 0.687 0.01809 0.00571 68.44% 32 0.044 33.879 0.556 0.543 0.01785 0.00585 67.23% 33 0.045 39.708 0.567 0.519 0.01539 0.00448 70.89% 34 0.046 47.223 0.594 0.616 0.01918 0.00589 69.29% 35 0.046 50.515 0.633 0.754 0.01930 0.00673 65.13% 36 0.046 41.078 0.624 0.641 0.01897 0.00542 71.43% 37 0.046 26.137 0.627 0.596 0.01586 0.00447 71.82% 38 0.045 23.639 0.574 0.476 0.01358 0.00381 71.94% 39 0.043 50.356 0.572 0.775 0.02290 0.00809 64.67% 40 0.044 33.198 0.572 0.565 0.01960 0.00567 71.07% 41 0.046 32.308 0.595 0.575 0.01845 0.00597 67.64% 42 0.045 44.453 0.579 0.691 0.02380 0.00756 68.24% 113 43 0.043 Table 6 (cont) 33.885 0.531 0.508 0.01520 0.00435 7 1 .38% Spechnen Thus to Peak Peak Stress Peak Strain Thickness Weigh; Dry VWNght Percent VWflbr 13L [MPa] [mm] 191 [91 44 45 46 47 43 Ahnuege Std. Dev. 0.046 0.045 0.043 0.042 0.045 0.045 0.001 36.975 39.134 32.184 37.827 38.276 40.940 12.403 0.570 0.575 0.531 0.569 0.539 0.823 0.031 0.592 0.627 0.505 0.629 0.593 0.828 0.103 0.01820 0.01722 0.01486 0.01743 0.01901 0.01909 0.00348 0.00509 0.0051 1 0.00402 0.00539 0.00607 0.00802 0.001 55 Table 7. Mechanical and weight data for all equilibrated explants 72 . 03% 70.33% 72.95% 69 .08% 68.07% 0.68783 0.03470 Sp. Tune to Peak Peak Shess Peak Shaun Thick ness lnflhfl ‘Hkflght 24hr “knght Dry VWflght Fflnal VWMnr “finer (ifln IS] [MPa] [mm] 19L [fl JHJ 101 102 103 104 105 106 107 108 109 110 111 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 0.04 0.043 0.038 0.037 0.034 0.042 0.032 0.04 0.045 0.032 0.039 0.032 0.035 0.038 0.027 0.04 0.045 0.044 0.044 0.037 0.045 0.04 0.046 0.039 0.044 0.043 0.046 0.038 29.6132 30.7522 31.3075 31.0925 28.1859 28.1466 26.7588 31.6889 41.6000 31.8526 23.8261 16.0949 32.6873 38.6346 16.2095 56.4879 57.2592 35.7709 41.2968 54.5093 32.7198 46.7933 46.8620 37.4273 44.5050 39.2148 50.2143 18.6666 0.6101 0.6939 0.6260 0.6375 0.6108 0.6342 0.6196 0.6471 0.6163 0.5927 0.7299 0.6218 0.6519 0.6843 0.5877 0.6219 0.6352 0.6195 0.6428 0.5977 0.6766 0.6526 0.6543 0.621 1 0.5816 0.5860 0.6261 0.7040 0.5705 0.8176 0.6205 0.6574 0.6242 0.6062 0.6465 0.6901 0.6585 0.6512 1 .0422 0.7497 0.8105 0.9684 0.5229 0.8006 0.7503 0.5680 0.6864 0.7331 0.7072 0.8123 0.7695 0.6361 0.6154 0.5839 0.7461 0.7376 0.01517 0.02511 0.01856 0.01916 0.01612 0.02028 0.01950 0.02116 0.01762 0.01894 0.02115 0.01876 0.02102 0.02283 0.01617 0.02385 0.02377 0.01570 0.02189 0.02516 0.02037 0.02665 0.02177 0.01787 0.01846 0.01732 0.02168 0.01437 114 0.01666 0.03199 0.02089 0.02135 0.01780 0.02575 0.02800 0.02580 0.01834 0.02166 0.02656 0.02332 0.02447 0.02600 0.02342 0.02556 0.02483 0.01815 0.02356 0.02530 0.02401 0.02895 0.02351 0.02050 0.02042 0.01925 0.02240 0.01661 0.00524 0.00749 0.00585 0.00630 0.00546 0.00665 0.00706 0.00681 0.00619 0.00625 0.00760 0.00525 0.00828 0.00868 0.00581 0.00800 0.00808 0.00492 0.00678 0.00782 0.00668 0.00846 0.00671 0.00563 0.00601 0.00631 0.00601 0.00456 68.55% 76.59% 72.00% 70.49% 69.33% 74.17% 74.79% 73.60% 66.25% 71 . 14% 71 .39% 77.49% 66.16% 66.62% 75.19% 68.70% 67.46% 72.89% 71 .22% 69.09% 72.18% 70.78% 71 .46% 72.54% 70.57% 67.22% 73. 17% 72.55% 9.82% 27.40% 12.55% 1 1.43% 10.42% 26.97% 43.59% 21.93% 4.09% 14.36% 25.58% 24.31% 16.41% 13.89% 44.84% 7.17% 4.46% 15.61% 7.63% 0.56% 17.87% 8.63% 7.99% 14.72% 10.62% 1 1.14% 3.32% 15.59% 130 0.045 46.3564 0.6417 0.7545 0.01967 0.02045 0.00664 67.53% 0.046 50.2062 0.5860 0.6815 0.01977 0.02042 0.00633 69.00% Table 7 (cont.) 131 132 133 134 135 136 137 138 139 140 141 142 143 144 146 147 148 Avg Std. Dev. 0.043 0.036 0.044 0.045 0.038 0.045 0.046 0.045 0.033 0.039 0.043 0.037 0.045 0.045 0.039 0.037 0.040 48.4188 41 .7288 58.6704 43.7454 40.9824 44.4755 49. 1 01 3 46.8620 37.6432 33.5299 22.9986 40.5880 24.5931 44.5139 42.4388 49.5821 38.4048 1 0.7872 0.6481 0.6038 0.5824 0.6563 0.6417 0.6555 0.6616 0.5993 0.5961 0.5875 0.5720 0.6197 0.5545 0.5908 0.6245 0.5701 0.8255 0.0371 0.8814 0.6522 0.6777 0.7392 0.7389 0.7573 0.8178 0.641 1 0.7826 0.5814 0.4491 0.7759 0.4342 0.7088 0.8032 0.7717 0.7050 0.1185 0.02499 0.02052 0.02126 0.02153 0.02536 0.02174 0.02147 0.02042 0.02544 0.01783 0.0141 1 0.02290 0.01554 0.02006 0.02509 0.02376 0.02047 0.00320 115 0.02632 0.02181 0.02208 0.02323 0.03001 0.02273 0.02272 0.02107 0.02990 0.01872 0.01505 0.02612 0.01654 0.02129 0.02880 0.02528 0.02299 0.00391 0.00793 0.00647 0.00750 0.00715 0.00847 0.00667 0.00775 0.00655 0.00858 0.00596 0.00427 0.00813 0.00504 0.00735 0.00910 0.00982 0.00883 0.001 25 69.87% 70. 33% 66. 03% 69 22% 71 .78% 70.66% 65. 89% 68.91 % 7 1 .30% 68. 1 6% 71 . 63% 68. 87% 69 . 53% 65.48% 68.40% 61 . 1 6% 70.1 6% 3.1 3% 3.97% 3.29% 5.32% 6.29% 3.86% 7.90% 1 8.34% 4.55% 5.82% 3. 1 8% 1 7.53% 4.99% 6.66% 14. 06% 6.44% 6. 1 3% 14. 79% 6.40% 1 2.44% 9.85% Appendix B: Chapter 2 Raw Data Table 1. Mechanical data for all rabbits time to Load peak Group Rabbit N msec time zero M54 621.16 3.2 M57 529.94 2.1 BU63 575.8 2.4 BU62 744.43 5.4 BB3 939.64 3.1 BU64 863.48 4.9 ave 712.41 3.52 stdev 164.78 1 .34 4 Days No Poloxamer M40 738.44 3.2 M72 912.19 4.1 481 BF 550.45 3.1 BU66 525.07 2.3 BU60 595.65 3.8 BU61 600.54 5.1 ave 653.72 3.60 stdev 146.56 0.96 4 Days Poloxamer M74 730 4 M78 641 3 M90 741 4.2 M82 579.1 2.6 48OBF 510.68 3.7 BU65 764.3 2.2 ave 661 .01 3.28 stdev 101 .52 0.81 Total Average 675.7 3.5 Total St Dev 134.4 1 .0 116 Table 2. Matrix Damage (total surface fissure length) Data for all rabbits 117 Matrix Damage (mm) Un-impacted Impacted time zero BU62 3.904468085 26.15521277 BU63 0 8.226702128 BU64 0.942659574 2.036914894 BB3 0 8.365744681 Average 1.21 1781915 1 1 .19614362 St Dev 1.84930766 1 0.40022374 4 Days No Poloxamer 481BF 0.796702128 6.747659574 BU66 0 4.297765957 BU60 0 2.558617021 BU61 5.112978723 16.67085106 Ave rage 1 .47742021 3 7.568723404 St Dev 2.452631496 6.306694917 4 Days Poloxamer M74 13.95170213 26.654361? M90 11.07319149 19.24531915 4808F 0.620212766 5.912234043 BU65 2.878085106 18.76308511 Ave rage 7.1 30797872 1 7.64375 St Dev 8.391 152389 8.814868203 Table 3. Time Zero group cell viability counts % °/o Controls Dead test Dead Super Middle Deep Total Super Middle Deep Total 6jc1a 7.67 9.63 8.20 8.54 6jt1a 23.71 33.82 44.32 31.70 6jc2a 19.77 21.71 11.00 17.70 6jt25 21.35 22.35 29.58 23.60 6jc33 8.37 12.54 0.00 9.45 6jt3a 58.79 41.52 16.67 41.29 6jo4a 8.66 6.21 14.94 8.78 6jt4a 22.99 29.91 32.37 28.97 6jc5a 2.94 5.47 9.34 5.51 6jt5a 32.96 23.40 31.32 28.01 6j06a 16.32 18.90 8.33 15.39 6jt6a 8.40 9.29 19.12 11.43 6jc7a 8.31 11.51 11.06 10.27 6jt7a 25.48 16.92 19.97 19.84 ' 6jt8a 7.72 13.10 25.00 13.29 avg 8.71 12.28 10.48 9.66 20.37 23.79 24.86 24.77 std dev 4.30 6.11 2.51 3.24 9.19 10.86 6.38 9.89 6jc1a 12.00 6.83 26.87 11.88 6jt1b 34.02 25.20 33.33 29.68 6jc3a 14.83 16.61 37.35 18.81 6jt2b 13.14 22.18 14.57 16.99 6jC48 16.78 15.50 13.99 15.36 6jt3a 41.10 34.64 22.02 31.01 6j05b 10.94 8.53 6.96 8.68 6jt4b 21.66 16.55 24.58 19.69 6jt5a 9.65 20.26 28.54 20.83 6jt6a 55.84 37.40 35.92 40.89 avg 13.64 11.87 15.94 13.68 29.24 26.04 28.88 23.64 std dev 2.66 4.91 10.10 4.37 17.73 8.27 5.81 6.30 BU62La 1.32 16.32 9.23 12.13 BU62Ra 41.10 48.65 28.10 38.63 BU62Lb 6.02 3.15 14.51 6.96 BU62Rb 36.29 58.79 37.75 46.20 BU62Lc 4.84 15.34 21.64 12.89 BU62Rc 4.70 14.25 20.22 13.84 BU62Ld 17.72 10.21 16.07 13.51 BU62Rd 22.42 2.54 20.39 11.53 BUSZLe 4.12 6.49 19.39 8.88 BU62Re 57.55 61.76 44.90 55.36 BU62Lf 7.09 4.03 34.72 11.81 BU62Rf 16.67 10.24 14.17 11.95 BU62Lg 3.80 3.98 12.40 5.63 BUGZRg 44.44 31.17 21.20 29.42 BU62Lb 13.33 4.88 7.39 7.66 Average 5.79 8.05 14.37 9.93 31.88 32.49 23.64 29.56 St Dev 3.79 5.28 5.16 3.01 18.17 24.28 8.21 17.83 BU63Le 14.47 18.90 8.81 15.45 BU63Ra 28.46 12.48 5.87 12.19 BU63Lb 2.53 9.66 5.71 7.62 BU63Rb 49.66 9.40 19.70 18.40 BU63Le 2.17 2.29 4.56 3.14 BU63Rc 59.47 39.85 25.45 41.66 BUG3Ld 3.87 0.88 3.57 2.33 BU63Rd 84.31 36.67 18.23 34.37 BU63Le 14.81 7.59 8.83 8.74 BU63Ra 75.29 35.49 33.91 45.41 BU63Lf 1.69 6.63 9.79 6.91 BU63Rf 89.16 46.12 47.95 56.35 BU63Lg 13.16 6.19 29.13 14.32 BU63Rg 58.82 15.29 27.66 29.62 118 BU63Lh 6.81 5.87 3.09 5.39 BU63Rb 55.94 25.59 11.30 30.60 Average 7.44 5.59 6.34 7.99 67.52 27.61 20.30 33.57 St Dev 5.79 3.03 2.77 4.78 15.30 13.90 9.66 14.33 BU64La 5.93 3.92 11.41 6.72 BU64Ra 37.86 41.87 27.59 38.10 BU64Lb 16.33 3.00 1.65 4.04 BU64Rb 27.35 36.60 19.58 31.71 BU64Lc 5.13 7.65 7.35 7.38 BU64Rc 58.33 25.52 15.91 30.96 BU64Ld 22.22 5.28 14.02 9.96 BU64Rd 65.27 38.10 23.36 41.02 BU64Le 17.65 7.01 7.28 8.43 BU64Re 48.98 35.61 17.74 31.48 BU64Lf 3.96 9.31 12.23 9.32 BU64Rf 4.17 27.06 16.38 21.65 BU64Lg 14.68 5.43 7.33 7.42 BU64Lb 11.57 11.79 12.20 11.83 Average 12.18 5.94 10.26 8.72 47.56 34.13 18.59 34.65 St Dev 6.66 2.19 2.86 1.79 15.28 6.45 3.02 4.60 BB3La 24.01 11.99 25.76 18.05 BB3Ra 58.18 33.05 26.92 40.00 BB3Lb 31.95 25.74 10.77 23.20 BB3Rb 96.94 79.41 21.43 75.51 BB3Lc 20.69 1 1.48 26.86 17.52 BB3Rc 70.65 35.94 14.61 33.33 BB3Ld 15.63 24.79 7.50 17.74 BB3Re 69.00 47.97 15.25 45.10 BB3Rf 76.50 42.97 6.83 38.38 BB3Rg 48.00 39.52 31.30 38.13 BBBRh 87.39 48.33 15.23 41.73 Average 20.11 18.50 17.72 17.77 72.38 41.30 18.80 39.45 St Dev 4.22 7.82 10.01 0.27 16.61 6.27 8.32 3.94 119 Table 4. 4 Day No Poloxamer cell viability counts % Controls % Dead test Dead Super Middle Deep Total Super Middle Deep Total M406jc1a 14.80 15.09 12.08 14.43 M406jt1e 15.15 21.50 14.29 18.02 M406jc2a 8.45 8.89 7.95 8.60 M406jt4a 27.88 29.29 20.43 27.23 M406j03a 12.28 20.00 13.66 15.51 M406jt5a 18.87 20.11 20.14 19.81 M406jo4a 36.67 22.08 8.68 18.37 M406jt63 34.53 27.97 17.34 25.63 M406'L05a 16.37 11.52 16.04 14.23 Average 12.98 15.52 11.68 15.63 20.63 24.72 19.30 22.67 St Dev 3.45 5.55 3.39 1.91 6.55 4.59 1.70 4.45 M72La2 58.06 40.61 27.82 43.68 M72Ra 13.04 5.24 11.97 9.00 M72Lb 16.33 11.89 25.24 16.15 M72Rb 25.89 27.01 22.52 25.34 M72Lc 18.37 10.48 28.22 17.68 M72Rc 33.03 51.69 49.67 46.05 M72Ld 7.94 11.53 46.89 20.07 M72Rd 84.73 64.17 31.97 60.68 M72Lc 2.47 2.77 7.04 3.93 M72Re 53.39 43.21 49.74 48.01 M72Lf 16.67 8.52 10.69 10.90 M72Rf 59.52 50.89 1.96 48.43 M72Lg 6.74 13.71 12.78 11.05 M72Rg 18.13 5.76 6.92 8.70 M72Lh 10.66 11.21 29.57 15.91 M72Rh 40.00 25.40 23.32 27.72 M72Li 16.51 4.51 6.32 7.77 M72Ri 6.03 15.04 41.07 16.77 Average 11.96 9.33 18.46 12.93 31.13 32.04 26.57 32.30 St Dev 5.83 3.82 10.16 5.45 19.04 21.40 17.82 19.08 481BFLa 5.29 6.01 2.72 5.00 481BFRa 94.00 47.33 5.80 41.77 481BFLb 12.93 0.99 6.40 4.79 481 BF Rb 42.16 59.00 37.93 49.86 481BFLc 6.45 8.00 11.34 8.73 481BFRc 61.45 14.51 8.21 22.69 481BFLd 7.92 14.04 13.77 12.50 4813FRd 64.13 25.09 29.30 37.05 481 BF Le 5.02 4.13 9.63 5.52 481BFRe 40.54 15.69 28.02 22.50 481BFLf 12.66 7.69 12.07 9.84 481BFRf 85.09 14.79 10.34 26.36 Average 8.38 5.36 10.64 6.78 64.56 23.48 19.93 33.37 St Dev 3.57 2.89 2.80 2.34 21.80 14.03 13.46 11.29 BU66La 17.13 11.11 7.63 12.01 BU66Ra 28.38 34.76 25.00 31.16 BU66Lb 5.00 1.55 6.32 3.65 BU66Rb 29.24 22.86 27.98 25.94 BU66Lc 3.74 3.39 2.71 3.32 BU66Rc 63.25 16.82 33.17 33.04 BU66Ld 3.80 2.47 8.20 3.89 BU66Rd 32.20 18.87 9.94 20.30 BU66La 9.20 2.83 6.02 4.83 BU66Re 52.30 28.27 1 1.99 28.08 BU66Lf 2.70 6.97 1.55 4.98 BU66Rf 42.11 29.97 46.75 36.35 BU66Lg 9.77 6.87 17.82 9.72 BU66Rg 48.47 27.05 19.62 32.37 BU66Lb 25.29 15.75 10.53 15.26 BU66Rh 43.51 15.21 14.84 21.27 Average 7.33 5.03 6.14 6.06 39.46 24.23 20.36 28.56 St Dev 5.13 3.42 3.12 3.40 9.57 6.92 8.69 5.74 BU60La 60.00 20.30 13.27 27.85 BU60Ra 57.24 29. 10 8.70 31.60 BUSOLb 21.90 9.72 9.68 13.74 BU60Rb 76.83 25.24 5.30 28.12 BU60Lc 6.25 4.17 15.73 7.96 BUGORc 40.37 16.90 14.38 19.84 BUGOLd 37.38 23.81 21.65 27.52 BU60Rd 66.15 27.32 9.09 32.82 BUBOLe 20.75 8.85 9.28 11.41 BU60Re 47.97 26.25 3.29 28.22 120 BU60Lf 14.29 3.66 4.55 5.54 BU60Rf 38.81 29.92 10.19 26.49 BU60Rg 88.74 30.80 27.68 44.93 BU60Rb 65.56 43.04 80.91 56.50 Average 20.1 1 1 1.75 10.50 15.67 60.21 26.50 1 1.23 30.29 St Dev 11.49 8.42 4.26 9.72 17.59 4.68 8.07 7.69 BU61La 3.85 9.19 12.02 9.35 BU61 Ra 8.57 18.09 3.74 11.87 BU61Lb 12.50 18.32 27.54 20.26 BU61 Rb 5.26 8.33 14.17 8.94 BU61Lc 0.00 10.47 15.89 11.51 BU61Rc 0.00 0.99 12.77 3.56 BU61Ld 5.41 11.39 29.67 15.66 BU61 Rd 18.53 10.63 21.05 14.86 BU61 Le 28.89 17.96 31.67 23.18 BU61Re 16.77 17.13 7.89 13.58 BU61Lf 2.78 7.66 6.11 6.72 BU61Rf 6.45 9.68 7.89 8.85 BU61Lg 4.65 7.64 5.33 6.62 BU61Rg 9.21 7.99 23.53 13.42 BU61 Rh 10.90 17.63 33.19 20.87 Average 4.86 11.80 18.32 13.33 9.46 12.78 13.01 10.73 St Dev 4.19 4.54 11.22 6.56 6.03 i 4.61 7.25 3.91 121 Table 5. 4 Day Poloxamer Group cell viability counts Controls % Dead test % Dead Super Middle Deep Total Super Middle Deep Total M74LFT1a 3.82 5.19 6.29 5.10 M74RT1a 25.97 14.63 9.22 15.54 M74LFT2a 14.63 30.15 6.08 21.93 M74RT2a 38.76 8.10 26.28 20.09 M74LFT3a 2.04 1.48 0.00 1.46 M74RT3a 9.06 9.76 38.00 15.61 R74LFT4a 9.22 12.70 9.38 11.08 Average 5.03 6.46 7.25 5.88 24.60 10.83 24.50 17.08 St Dev 3.74 5.72 1.84 4.86 14.90 3.39 14.47 2.61 M82L1 13.97 9.66 10.71 11.28 M82R1 7.27 3.43 8.65 5.51 M82L2 65.85 45.96 32.84 45.68 M82R2 13.66 8.74 20.71 13.33 M82L3 2.31 3.00 15.55 6.85 M82R3 7.77 5.41 13.16 7.54 M82L4 4.62 5.21 19.64 8.07 M82R4 32.57 20.05 19.05 23.29 M82L5 6.43 7.47 4.73 6.54 M82R5 30.52 11.20 5.50 17.41 M82L6 1.05 0.79 10.64 2.61 M82R6 70.18 31.17 6.41 35.59 M82L7 9.78 2.95 13.86 7.68 Average 6.36 4.85 12.52 7.17 18.36 9.77 12.25 13.42 St Dev 4.84 3.27 5.08 2.80 12.32 6.48 6.50 7.26 M78LFT1a 1.60 0.24 0.00 0.40 M78RT1a 2.34 0.77 1.69 1.47 M78LFT2a 2.78 3.07 7.54 4.26 M78RT4a 21.52 13.89 10.63 15.57 M78LFT4a 17.91 12.88 16.75 15.05 M78RT5a 27.61 7.86 4.17 10.89 M78LFT5a 2.41 4.99 1.38 3.57 M78RT6a 1.33 5.33 6.60 4.83 M78LFT6a 4.39 3.93 5.59 4.46 Average 2.79 3.06 3.63 3.17 13.20 4.65 4.15 8.19 St Dev 1.17 2.03 3.53 1.89 13.36 3.59 2.46 6.27 M90L1 31.15 15.73 ' 7.84 16.06 M90R1 55.38 21.48 25.56 27.15 M90L2 0.00 0.22 0.00 0.17 M90R2 89.33 95.80 77.27 91.29 M90L3 82.72 38.03 24.22 44.06 M90R3 35.85 32.25 21 .27 29.66 M90L5 80.79 54.98 28.22 55.23 M90R4 14.29 12.75 7.78 11.95 M90L6 11.68 4.14 3.77 6.70 M90R5 50.98 61.27 12.93 40.70 Average 41.27 14.53 12.81 24.44 39.12 31.94 16.88 27.36 St Dev 38.60 17.00 12.63 24.02 18.55 21.12 8.02 11.84 4808FLa 19.13 9.56 2.52 10.18 4808FRa 35.09 15.24 23.62 24.69 4808FLb 17.47 4.72 3.61 7.90 48OBF Rb 27.59 6.09 8.87 1 1.59 4808FLc 10.39 2.16 5.26 3.79 4808FRc 8.49 3.47 5.07 5.16 48OBFLd 29.79 13.03 7.10 12.71 4803FRd 15.38 11.68 9.68 11.62 4808FLe 31.46 2.47 3.53 6.27 4808FRe 16.59 9.45 4.72 9.99 4808FLf 16.07 13.66 7.53 13.09 4808FRf 4.55 2.19 5.88 3.99 4808FLg 13.04 6.79 3.49 6.41 4803FRg 27.19 3.98 9.45 10.77 122 4803FLh 3.67 1.18 4.43 2.61 480BFRh 15.79 9.72 7.58 11.22 Average 17.63 6.70 4.68 7.87 18.83 7.73 7.32 9.19 St Dev 9.35 4.92 1.81 3.87 10.34 4.54 2.10 3.22 BU65La 5.94 8.04 2.53 5.91 BU65Ra 40.28 34.47 5.47 27.05 BU65Lb 3.16 4.35 6.14 4.73 BU65Rb 44.67 24.54 11.41 24.08 BU65Lc 21.52 9.50 7.33 9.96 BU65RC 64.90 29.22 4.47 24.07 BU65Ld 25.00 7.35 15.18 13.19 BU65Rd 60.30 44.71 12.90 38.97 BU65Le 1.35 0.93 7.64 3.34 BU65Re 48.78 23.61 3.16 20.00 BU65Lf 13.51 4.55 8.66 7.06 BU65Rf 12.12 4.56 3.79 5.88 BU65Lg 7.98 2.78 10.45 6.86 BU65Rg 40.33 23.71 1.99 22.03 BU65Lh 2.35 3.50 26.82 10.70 BU65Rb 30.07 23.65 5.30 19.35 Average 10.10 5.13 8.28 7.72 47.05 29.13 5.08 22.76 St Dev 9.03 2.91 - 3.90 3.31 12.13 7.98 3.04 2.88 123 Appendix C: Chapter 3 Raw Data Table 1. Mechanical data for the high rate of loading “with GlcN” (1-28) and “no GlcN” (29- 56) explants Time to Peak Peak Specimen Peak Stress Strain Thickness [8] [MP8] {mm} 1 0.034 28.954 0.374 0.615 2 0.032 28.479 0.397 0.638 3 0.031 28.529 0.606 0.593 4 0.033 28.830 0.539 0.550 5 0.04 28.41 1 0.526 0.545 6 0.036 28.404 0.573 0.556 7 0.032 28.299 0.481 0.634 8 0.04 28.319 0.539 0.590 9 0.028 28.918 0.532 0.597 10 0.037 28.784 0.681 0.793 11 0.031 28.760 0.515 0.814 12 0.034 28.852 0.568 0.663 13 0.029 28.692 0.483 0.659 14 0.038 28.670 0.479 0.617 15 0.034 28.705 0.528 0.735 16 0.033 28.749 0.479 0.856 17 0.029 28.952 0.529 0.738 18 0.033 28.620 0.625 0.742 19 0.026 29.024 0.571 0.789 20 0.036 28.462 0.445 0.678 21 0.027 28.010 0.525 0.634 22 0.039 27.408 0.491 0.687 23 0.039 27.454 0.468 0.777 24 0.036 28.061 0.508 0.906 25 0.039 27.346 0.442 0.725 26 0.04 27.261 0.469 0.512 27 0.034 28.090 0.605 0.931 28 0.037 27.379 0.523 0.690 Avergge 0.034 28.31 5 0.517 0.681 Std. Dev. 0.004 0.575 0.065 0.1 1 2 124 Time to Peak Peak Specimen Peak Stress Strain Thickness [8} [MP8] {mm} 29 0.039 29.128 0.676 0.530 30 0.031 28.849 0.559 0.566 31 0.034 28.840 0.457 0.603 32 0.036 28.777 0.504 0.338 33 0.030 28.921 0.617 0.548 34 0.031 29.183 0.536 0.609 35 0.031 28.630 0.483 0.657 36 0.032 28.764 0.471 0.715 37 0.029 28.858 0.610 0.546 38 0.025 28.620 0.569 0.685 39 0.034 28.986 0.475 0.698 40 0.034 28.960 0.563 0.722 41 0.029 29.060 0.529 0.677 42 0.033 29.032 0.530 0.597 43 0.037 28.579 0.494 0.695 44 0.032 28.865 0.509 0.587 45 0.025 29.339 0.537 0.787 46 0.029 29.550 0.676 0.732 47 0.032 28.918 0.664 0.601 48 0.026 28.192 0.647 0.871 49 0.034 28.757 0.514 0.688 50 0.033 28.405 0.524 0.696 51 0.038 28.069 0.582 0.604 52 0.033 28.444 0.608 0.987 53 0.036 28.514 0.543 0.763 54 0.037 28.011 0.480 0.725 55 0.039 27.789 0.539 0.612 56 0.040 27.808 0.613 0.549 Average 0.033 28.641 0.551 0.661 Std. Dev. 0.004 0.473 0.062 0.11 6 125 Table 2. Mechanical data for the low rate of loading “with GlcN" (57-84) and “no GlcN” (85-112) explants Time to Peak Peak Specimen Peak Stress Strain Thickness [8} [MP8] {mm} 57 0.901 29.372 0.576 0.437 58 0.879 29.370 0.637 0.425 59 0.797 29.452 0.588 0.648 60 0.800 29.471 0.686 0.687 61 0.875 29.367 0.569 0.534 62 0.858 29.390 0.664 0.548 63 0.878 29.391 0.697 0.585 64 0.833 29.449 0.716 0.744 65 0.830 29.396 0.722 0.696 66 0.883 29.413 0.639 0.638 67 0.954 28.931 0.525 0.606 68 0.954 28.945 0.628 0.561 69 0.948 28.918 0.651 0.571 70 0.948 28.893 0.622 0.383 71 0.941 28.934 0.687 0.585 72 0.955 28.905 0.499 0.642 73 0.936 28.914 0.571 0.616 74 0.948 28.939 0.546 0.553 75 0.941 28.939 0.608 0.802 76 0.947 28.941 0.595 0.723 77 0.878 28.075 0.536 0.776 78 0.866 28.247 0.659 0.973 79 0.840 28.162 0.579 0.818 80 0.912 28.038 0.570 0.713 81 0.823 28.169 0.692 0.914 82 0.895 28.223 0.521 0.721 83 0.837 28.087 0.582 0.590 84 0.903 28.205 0.564 0.646 Avergge 0.890 28.833 0.61 5 0.653 Std. Dev. 0.049 0.521 0.062 0.135 126 Time to Peak Peak Specimen Peak Stress Strain Thickness [8] [MP8] 1mm} 85 0.823 29.467 0.619 0.574 86 0.828 29.529 0.715 0.747 87 0.874 29.432 0.625 0.583 88 0.886 29.452 0.578 0.552 89 0.880 29.409 0.556 0.523 90 0.861 29.452 0.636 0.681 91 0.886 29.411 0.540 0.452 92 0.840 29.452 0.605 0.708 93 0.826 29.524 0.655 0.949 94 0.852 29.426 0.645 0.649 95 0.941 28.882 0.413 0.569 96 0.945 28.916 0.536 0.700 97 0.945 28.868 0.612 0.550 98 0.946 28.882 0.550 0.468 99 0.943 28.880 0.481 0.411 100 0.983 28.927 0.569 0.592 101 0.946 28.895 0.491 0.627 102 0.942 28.939 0.584 0.653 103 0.934 28.921 0.631 0.675 104 0.910 28.973 0.712 0.889 105 0.870 28.306 0.572 0.839 106 0.867 28.260 0.644 0.908 107 0.857 28.156 0.622 0.720 108 0.925 28.208 0.600 0.713 109 0.866 28.188 0.685 0.875 110 0.854 28.151 0.628 0.597 111 0.916 28.113 0.485 0.642 112 0.912 28.139 0.556 0.623 Average 0.895 28.851 0.590 0.662 Std. Dev. 0.043 0.529 0.068 0.133 127 Table 3. Total length of surface fissures for the high rate of loading “with GlcN” (1-28) and “no GlcN” (29-56) explants Specimen Surface Fissure Lengtlu [ml 1 11.574 2 12.575 3 47.798 4 21.272 5 32.920 6 37.343 7 36.256 8 45.666 9 30.735 10 45.101 1 1 40.625 12 27.741 13 29.936 14 13.609 15 26.345 16 49.226 17 35.766 18 48.512 19 48.906 20 11.126 21 17.861 22 16.551 23 29.659 24 30.885 25 7.417 26 32.238 27 47.488 28 37.865 Average 31 .480 Std. Dev. 1 3.31 5 128 Specimen Surface F issure Lenfl {mm} 29 40.615 30 53.915 31 35.265 32 14.622 33 56.803 34 40.966 35 50.718 36 25.471 37 48.885 38 54.458 39 54.458 40 29.659 41 39.005 42 1 5.741 43 38.888 44 72. 107 45 46.870 46 59.243 47 56.750 48 13.460 49 17.542 50 63.176 51 37.716 52 47.989 53 53.339 54 53.670 55 55.247 56 48. 107 Average 42.787 Std. Dev. 1 5.273 129 Table 4. Number of dead cells per fissure, per millimeter of fissure depth, and percentage of dead cells in each zone for the high rate of loading “with GlcN” (1-28) and “no Glc ” (29-56) explants Dead Cells Dead Cells Spec per Depth per fissure Sugrficial Middle Deep % % Dead Dead % Dead cells - cells cells 1 271.244 45.5 47.57% 9.30% 9.23% 2 589.093 79 74.68% 14.90% 17.29% 3 395.119 69 25.61% 21.90% 4.76% 4 576. 166 54 38.33% 9.45% 12.50% 5 740.307 67 28.09% 1 3. 1 3% 3.88% 6 165.768 72 44.64% 23.41% 19.35% 7 203.675 23 28.99% 3.67% 7.84% 8 103.434 10.5 20.00% 14.47% 3.45% 9 344.611 59 37.50% 15.19% 9.03% 10 167.447 27 32.99% 8.66% 4.61% 11 492.672 61 57.07% 14.10% 14.77% 12 656.843 28 15.79% 21.22% 2.60% 13 763.862 121 52.19% 7.65% 5.56% 14 867.998 187 35.00% 76.66% 30.51% 15 621 .828 45 53.49% 49.02% 58.87% 16 252.752 32 14.25% 1.44% 0.64% 17 494.330 43 38.52% 30. 18% 5.59% 18 50.331 10 6.56% 8.39% 30.26% 19 52.011 11 8.21% 27.74% 31.71% 20 195.570 12 4.72% 4.39% 3.94% 21 485.298 76 23.08% 37.58% 18.37% 22 41 1 .184 50 28.28% 13.40% 94.95% 23 76.046 6.5 13.85% 9.18% 23.08% 24 364.015 70 42.86% 12.59% 7.74% 25 877.514 144 51.85% 15.64% 23.53% 26 295.421 18.5 34.88% 12.56% 8.96% 27 288.147 48 36.59% 16.04% 18.26% 28 92.807 25.5 31 .31 % 6.78% 3.64% Avergge 384.036 52.083 32.81% 17.57% 16.75% Std. Dev. 242.015 40.722 16.46% 15.34% 19.51% 130 Dead Cells Dead Cells Spec per Dean per fissure Superficial Middle Deep % Dead % Dead % Dead cells cells cells 29 1343.968 39.5 87.64% 30.51% 4.30% 30 1 1 11.769 30 76.83% 18.09% 1.89% 31 506.197 38 60.40% 42.13% 10.23% 32 715.533 62 41.67% 64.03% 16.83% 33 229.293 21 88.64% 23.23% 6.50% 34 778.702 81 56.90% 19.49% 0.82% 35 640.310 67 64.20% 19.65% 21.24% 36 849.974 89 36.36% 1 1.50% 21 .97% 37 625.101 43.5 53.70% 53.52% 38.10% 38 1294. 306 138 23.33% 27.37% 39.51% 39 1000.114 90 44.93% 49.46% 21.78% 40 403.560 98 87.64% 83.08% 20.00% 41 631.756 153 26.76% 17.33% 2.13% 42 384.203 58 77.08% 86.73% 32.26% 43 1075.168 120 74.53% 59.23% 8.00% 44 1451.708 254 63.30% 56.44% 11.71% 45 1036.651 199 35.65% 11.88% 0.00% 46 910.239 89 55.56% 72.88% 62.64% 47 850.355 77.50% 75.59% 43.1 7% 48 1965.854 205 48.59% 43.50% 17.50% 49 766.297 84 44.87% 7.54% 16.89% 50 243.619 17.5 48.62% 26.36% 14.04% 51 823.129 121 60.38% 29.33% 24.00% 52 431.722 92 32.89% 17.91% 22.81% 53 568.862 88.5 63.77% 14.93% 10.26% 54 615.385 36 75.31% 21.01% 26.62% 55 1919.585 82 49.69% 8.77% 17.46% 56 994.941 61 40.16% 25.00% 15.15% Average 831.779 90.150 56.30% 35.08% 18.31% Std. Dev. 438.445 55.038 18.50% 23.85% 14.09% 131 Table 5. Number of dead cells per fissure, per millimeter of fissure depth, and percentage of dead cells in each zone for the low rate of loading “with GlcN” (57-84) and “no Glc ” (85-112) explants Dead Cells Dead Cells Spec per Depth per fissure Superficial Middle Deep % Dead % Dead % Dead cells cells cells 57 729.990 44.5 59.76% 21.51% 12.61% 58 1285.474 98 74.32% 26.29% 16.42% 59 276.380 28.5 52.17% 18.11% 6.33% 60 502.623 47 68.09% 9.31% 9.68% 61 406.038 31 .5 30.83% 4.92% 4.27% 62 244.451 42 26.67% 10.34% 5.26% 63 338.911 47 50.65% 12.05% 2.86% 64 101.608 11 39.13% 3.80% 1.67% 65 250.448 29 46.91% 22.73% 13.95% 66 18.03% 13.28% 15.71% 67 712.474 158 58.46% 30.19% 6.73% 68 1849.571 127 47.44% 25.29% 97.70% 69 1238.857 102 43.30% 12.50% 5.13% 70 1602.702 214 66.09% 54.25% 8.64% 71 305.938 73 41.38% 21.70% 12.24% 72 917.397 72 37.69% 7.42% 2.78% 73 667.539 113 27.71% 63.56% 27.56% 74 1003.215 160 36.73% 41.48% 23.39% 75 729.990 44.5 57.80% 22 .03% 16.82% 76 1285.474 98 46.58% 23.74% 6. 77% 77 276. 380 28.5 64.00% 14.46% 17.86% 78 502.623 47 72.73% 12.82% 7.96% 79 406.038 31.5 58.76% 35.03% 17.91% 80 244.451 42 25.00% 6.25% 23.21% 81 1096.694 121 41.18% 11.01% 14.00% 82 338.911 47 57.35% 23.67% 8.18% 83 101.608 11 66.67% 14.58% 8.16% 84 250.448 29 53.45% 26.67% 22.62% Avegge 664.740 72.1 03 48.89% 21 .04% 1 4.87% Std. Dev. 467.848 50.385 15.20% 14.13% 17.68% 132 Dead Cells Dead Cells Spec per Depth per fissure Superficial Middle Deep ' % Dead % Dead % Dead cells cells cells 85 1164.815 160 61.76% 59.04% 19.80% 86 621.316 70 78.38% 64.00% 37.04% 87 1065.220 182 81 82% 56.68% 29.71% 88 1488.102 67.5 77.06% 25.96% 14.71% 89 708. 396 89 83.56% 32.38% 12.50% 90 809.758 126 64.08% 16.67% 17.48% 91 930.205 108 58.33% 33.74% 26.27% 92 990. 388 84 72.34% 49.22% 1 6.16% 93 1209.386 87 38.54% 20.91% 2.60% 94 1298.088 274 64.08% 46.41% 34.33% 95 652.543 94 43.48% 25.40% 4.23% 96 946.971 72.5 67.95% 1 1 .58% 10.62% 97 931.217 62.5 44.26% 19.67% 6.58% 98 896.670 142 66.00% 34.29% 21.71% 99 808. 595 109.5 64.84% 22.86% 12.33% 1 00 1286. 551 244 80.56% 54.01% 9.85% 101 703.538 149 44.23% 69.09% 23.31% 102 662.031 216 68.18% 53.42% 22.40% 103 398.012 77 50.00% 9.17% 13.75% 104 646.105 62 87.50% 28.24% 11.96% 105 522 .482 54 66.67% 29.02% 9.62% 106 1360.639 131 72.15% 30.57% 16.67% 107 611.612 72 41.94% 15.61% 8.70% 108 645.588 50.5 57.14% 6.09% 3.15% 109 907.029 80 51.30% 25.48% 28.21% 1 10 479.660 35 37.96% 2.27% 0.88% 111 1151.738 92 64.37% 34.66% 19.67% 112 925.558 162 61.36% 54.12% 16.85% Avme 898.920 1 10.794 62.49% 33.23% 1 6.1 1 % Std. Dev. 290.508 58.817 1 4.25% 18.41 % 9.41 % 133 Table 6. Material property values for the theoretical model “no GlcN” “with GlcN” E1 (MPa) 7.835 7.835 E3(MPa) 1.53 1.53 k1(*10'” m2) 5.95 3.149 k3 («10'13 mT) 1.222 0.858 134 Appendix D: Rabbit impact standard operating procedure RABBIT PATELLOFEMORAL IMPACT SOP What to bring with you: 1.) Portable computer w/ A2D board, 2.) LPS lubricant, ethanol, 3.) meter stick, 4.) rabbit data sheets (a copy is attached to the end of this SOP), 5.) blank IBM formatted disks Pre-test set up: (* Leave portable computer off until all cables are connected) 1. Plug in Valadyne strain gauge amplifier (SGA) to the wall. Turn on the SGA and insure that the trigger release switch (see figure 1) is turned off (down position) to protect from accidental triggering. The SGA will need to be on for 15 minutes in order to allow the electronics to stabilize. Assure that all electronic connections are in place. Figures 2 through 4 carefully detail where all connections should be made. Once you have made all the necessary connections, turn on and start up computer. Computer will give a list of possible configurations, choose “Ethernet configuration”. Spray some LPS greaseless lubrication on a rag and wipe down the steel rod and gray rails of the impact cart (see figure 5). Do this very sparingly. Use alchohol to clean the sides of the cart (see figure 5) i.e., the portion of the cart where the brakes act. Keep rabbit hair off the rubber brake pads. After the 15 minutes are up, run the program “rabinsur.vi” located on the portable computers desktop. If the program is running (hit small arrow in the top left of screen in LabView) you will see a readout for the load. Using the small screwdriver, zero the load cell on the SGA. Use the opening to the top left of the black dial (see figure 1). Use the readout in LabView. Calibrate the load cell by depressing and holding the shunt cal switch on the Valadyne strain gauge amplifier. If needed readjust the set point to 3349 N on LabView (or 7.53 Volts on the voltmeter) by using the Gain knob (See figure 1). **** The values in step 9 are susceptible to change whenever the load cell is changed. Always double check these values with the load cell specifications. 135 10. Double check to make sure the load cell is still zeroed, adjust if necessary. Rabbit preparation: 1. On the Data Sheet, record the Rabbit name, weight (kg), and sex. 2. Once the rabbit is fully sedated, pull the left hind foot through the very bottom hole of the leather strap of the holding chair and tuck the rest of the strap under the rabbit so it can be fixated to the underside of the chair. 3. Position the right leg so that the femur is pointing vertical and the impactor is directed to hit the middle of the patella. 4. Place the black strap around the right hind foot and pull tight to insure the foot is fully constrained. It is important that the femur remain vertical, but the tibia can be horizontal in order to flatten the patella. Fix the end of the strap, as well as the leather strap, to the Velcro pad on the underside of the chair. 5. Move the clamping bar into position and attach the free end to the distal clamp. 6. Slowly apply even pressure to both clamps until the clamps lock in place. 7. Slide the chair into position so the patella is directly under the head of the impacting cart. 8. Lower the cart so the head of the impacting cart is just above the patella, checking to insure the patella is centered under the head. 9. Raise the impact cart to the desired height. Measure from top of patella to the bottom of the impactor head using the meter stick (see figure 6) 10. The position of the beam holding the upper brake may need repositioning to accommodate the desired height. Do this by removing the screws of the beam and moving the beam to the desired height above the patella. 1 l. The height and weight of the sled should be: Energy Height Mass 6 J 0.46meters 1.33kg 10 J 1 meter 1 kg 0 figure 5 shows 6 J set-up I 0 6 J impacts should see between 500 and 600 N as to where 10 J impacts should see from 900 to 1200 N 136 Impacting the rabbit: 1. 2. 99:59“ 10. 11. Turn on the trigger enable switch on the Valadyne strain gauge amplifier. Click the arrow at the top left of the LabView interface (the “run” arrow) (see figure 7). This should enable the load readout to be constantly changing (if load readout is already constantly changing then the arrow has already been hit). Then click on the large green START button below the graph area on LabView (button that reads “disabled” in figure 7). WAIT AT LEAST 5 SECONDS BEFORE PROCEEDING. Press the red trigger button on the Valadyne SGA to drop the impact cart. Impact cart will drop and the A/D board will trigger and save the data LabView will prompt user for a filename and location. Switch the trigger enable switch to disable (down position) before removing the rabbit. Excel will automatically run and a macro will plot the data Note the peak load and time to peak on the data sheet. Also sketch the graph and note any comments. Choose “save as” from the file menu. Save the file as an Excel Worksheet i.e., “.xls” format. Back up all data files to a floppy disk. FOLLOWING IMPA CT ION OF ALL ANIMALS, REMO VE DISK FROM DRIVE ((1:12 AND COPY ALL FILES TO THE (g:luserlbimfladz DIRECTORY FOR PERMANENT DOCUMEN TA TION.’ When testing is done shut down computer before disconnecting cables. Turn off voltmeter and leave on SGA if more testing is going to be done in the next week. 137 Figures Load Cell Balance Screw Trigger Enable Switch Load Cell Gain Knob Trigger Button Load Out: to Interface Box see Fig 3 To Catch Release (yellow chord on drop fixture 120 VAC) Load In: from load cell .3 a Trigger Out: to Intterface Box see Fig 3 Power from Wall . 4 Figure 2. Rear view of Validyne SGA 138 Load Out from SGA (see figure 2) Voltmeter set to 20 VDC Trigger Out from SGA (see figure 2) Interface Box Out to A/D Board (see figure 4) Portable Computer with A2D Card Out from Interface Box n.»nnnn»» . ”nun/2” "Mme.” . Figure 4. Connection from interface box to A2D card 139 Removable Weights (nuts, screws included). Cart Sides Gray Plastic Rails Steel Rod Figure 5. Cart with 1.33 Kg (6 Joule) set up 140 Load Cell Figure 6. Lower view of impact drop fixture w/ load cell and impactor head 141 L ad Cell Reado t Figure 7. Screen shot of labview layout "Trigger Value has been changed to 50 in order to prevent accidental triggering 142 sl ‘1] I11 1. uljt‘jjjjjjjjj1111111