(A LIBRARY M Q Michigan State N7 University This is to certify that the thesis entitled MATERIAL PROPERTY DOCUMENTATION AND FRACTURE ANALYSES OF THE DEVELOPING SKULL presented by TIMOTHY G. BAUMER has been accepted towards fulfillment of the requirements for the MS. degree in Engineering Mechanics MWCAWL/ fl Mgflar Professor's Signature M Date MSU is an Alfirmatlvo ActiorVEquaI Opportunity Employer 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 5108 K:IProj/Acc&Pres/CIRCIDateDue.indd MATERIAL PROPERTY DOCUMENTATION AND FRACTURE ANALYSES OF THE DEVELOPING SKULL By Timothy G. Baumer A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Engineering Mechanics 2009 ABSTRACT MATERIAL PROPERTY DOCUMENTATION AND FRACTURE ANALYSES OF THE DEVELOPING SKULL By Timothy G. Baumer Skull fracture interpretation is a necessary component of forensic analysis and contributes to the determination of the cause and manner of injury or death. The research presented in this thesis makes use of an in situ porcine model and finite element analysis to examine the responses of the developing skull to blunt force impulse and quasi-static crushing trauma. Chapter 2 documents age effects on the mechanical properties of parietal bone and coronal suture in porcine infants and correlates the bending properties of the bone to existing human infant data. Chapter 3 documents skull fiacture on infant porcine specimens with respect to age and interface in an effort to identify fi'acture characteristics for a developing skull under known conditions. Chapter 4 uses a simplified cranial model and the finite element method to generate representative patterns of principal stress and strain directions developed during quasi-static crushing injuries to the developing skull. The theoretical results were compared to four clinical cases of known childhood fatalities from crushing head injuries inflicted by vehicle tires. Chapter 5 uses a simplified porcine cranial model and the finite element method to predict experimentally documented fracture patterns developed during impact loading to the parietal bone. The material properties of the cranial model were taken from the previously obtained results from Chapter 2. The experimentally inflicted fractures to the cranium fi'om Chapter 3 were used for comparison with the theoretically developed principal stress and strain directions. ACKNOWLEDGMENTS I would like to thank my advisor and mentor, Dr. Roger Haut, for all his support and guidance over the past 4 years. I am also grateful to Dr. Todd Fenton for all his support and willingness to teach anthropology to an engineer. I would also like to thank all the members of the 4th floor of East Fee Hall for their fiiendship and support. I am grateful to Jean Atkinson, Jane Walsh, and Cliff Beckett for their continual technical support. Finally, I would like to acknowledge the contributions of Mr. Ed Reed and Ms. Star Lewis for supplying and collecting the porcine specimens. iii TABLE OF CONTENTS LIST OF TABLES .......................................................................................................... vi LIST OF FIGURES ............................................................................................... i ......... vii RESEARCH PUBLICATIONS ....................................................................................... x CHAPTER 1: Characteristics of the Developing Skull - An Introduction ...................... 1 CHAPTER 2: Age-Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human ......................................................... 10 Abstract .............................................................................................................. 10 Introduction ........................................................................................................ 12 Materials and Methods ....................................................................................... 14 Results ................................................................................................................ 18 Discussion .......................................................................................................... 23 References .......................................................................................................... 28 CHAPTER 3: Age-Dependent Fracture Characteristics of Rigid and Compliant Surface Impacts on the Infant Skull — A Porcine Model .......................................... 30 Abstract .............................................................................................................. 30 Intrdduction ........................................................................................................ 32 Materials and Methods ....................................................................................... 34 Results ................................................................................................................ 38 Discussion .......................................................................................................... 42 References .......................................................................................................... 47 CHAPTER 4: Characteristics and Prediction of Cranial Crush Injuries in Children ..... 50 Abstract .............................................................................................................. 50 Introduction ........................................................................................................ 52 Materials and Methods ....................................................................................... 54 Results ................................................................................................................ 56 Discussion .......................................................................................................... 61 References .......................................................................................................... 65 CHAPTER 5: Finite Element Modeling of Rigid and Compliant Surface Impacts to the Pediatric Porcine Parietal Bone ...................................................................... 66 Abstract .................................................................. 66 Introduction ........................................................................................................ 68 Materials and Methods ....................................................................................... 70 Results ................................................................................................................ 72 Discussion .......................................................................................................... 74 References .......................................................................................................... 78 CHAPTER 6: Conclusions and Recommendations for Future Work ............................ 80 iv APPENDD( A ................................................................................................................ 83 APPENDIX B ................................................................................................................. 94 LIST OF TABLES Table 5.1. (Average) properties used in development of the FE model and impact conditions. .................................................................................................... 71 Table A. 1. Raw data collected from 4-point bending tests of bone-suture-bone specimens ..................................................................................................... 84 Table A2. Raw data collected fi'om 4-point bending tests of perpendicular bone specimens ..................................................................................................... 86 Table A3. Raw data collected from 4-point bending tests of parallel bone specimens ..................................................................................................... 88 Table A4 Raw data collected from 4-point bending tests of bone-suture-bone specimens NOT resulting in suture failure ................................................... 90 Table A5. Raw data collected fi'om 4-point bending tests for multiple bone- suture-bone specimens taken from the same cranium .................................. 91 Table A6. Porosity data from stained cross-sections of perpendicular bone specimens ..................................................................................................... 92 Table A.7. Porosity data fi'om stained cross-sections of parallel bone specimens ......... 93 Table B.1. Raw data collected fiom rigid interface impacts .......................................... 95 Table 3.2. Skull damage measurements for rigid interface impacts .............................. 97 Table 83. Raw data collected from compliant (1.1 MPa Hexcel) interface impacts 98 Table 3.4. Skull damage measurements for compliant interface impacts ................... 100 LIST OF FIGURES Figure 1.1. Diagram of a newborn skull ........................................................................... 2 Figure 2.1. Top view of cranium with locations of the beam specimens ....................... 15 Figure 2.2. Diagram of testing fixture ............................................................................ 16 Figure 2.3. Representative stained cross-section of a beam from a 14 day specimen showing separation of compact and porous layers. Voids in the bone matrix were colored black. The box represents the analyzed area of a test specimen .................................................................................. 18 Figure 2.4. Recorded values of stiffness for (a) both orientations of bone and (b) bone-suture-bone specimens flour 3 to 21 days. *Significantly different means ............................................................................................. 19 Figure 2.5. Linear regressions of ultimate stress for bone and bone-suture—bone specimens against age ...... . .......... 20 Figure 2.6. Ultimate strain for bone and bone-suture-bone specimens. *Significantly different means ..................................................................... 20 Figure 2.7. Bending modulus for bone and bone-suture-bone specimens versus age. *Significantly different means .............................................................. 21 Figure 2.8. Strain energy of bone and bone-suture bone specimens versus age. *Significantly different means ..................................................................... 22 Figure 2.9. Linear regression analyses of both orientations of infant porcine bone specimens against age. Both orientations displayed significant dependencies on age ..................................................................................... 22 Figure 2.10. Bending rigidity of human and porcine specimens versus age (Human - months; porcine — days). The human data was based on data fiom Coats and Margulies (2006) and unpublished data from Li et al (2007) ........................................................................................................... 24 Figure 3.1. Skull positioning and set-up showing rigid impact head ............................. 35 Figure 3.2. Drop test fixture set-up ................................................................................ 36 Figure 3.3. Overlay of force-time plots showing typical characteristics of an impact that caused fracture and one that did not .......................................... 38 Figure 3.4. Peakiforce during impact for rigid and compliant interfaces ....................... 39 vii Figure 3.5. Contact area recorded on pressure sensitive fihn during impacts for rigid and compliant interfaces ...................................................................... 39 Figure 3.6. Representative fracture sites for (a) rigid and (b) compliant interface impacts (five-day-old specimens shown). Fracture sites initiated at bone-suture boundaries. Compliant interface impact showed increased fracture lengths at similar initiation sites as the rigid interface impacts as well as diastatic fracture of coronal suture ............................................... 40 Figure 3.7. Frequency of diastatic fractures caused by both rigid and compliant interfaces versus age ..................................................................................... 41 Figure 3.8. (Average) Fracture length generated by rigid and compliant impacts at the same impact energy ................................................................................ 42 Figure 4.1. Simplified cranial model showing modeled landmarks of the basicranium. View of endocranial surface with vault removed ................... 55 Figure 4.2. Diagrams of fractures due to quasi-static crushing forces to skulls aged (a) 5 years -— Case 1, (b) 6 years — Case 2, (c) 3 years - Case 3, and (d) 1.5 years — Case 4. White cranial margins represent contact of applied forces (ground and tire) ................................................................................ 58 Figure 4.3. (Case 1) Crushing injury to the skull of a 5-year-old. Diagrams of (a) clinically observed fiactures and (b) an overlay of the highest 20% of principal tensile stresses produced from a FE model. White cranial margins represent contact of applied forces (ground and tire) ..................... 5 9 Figure 4.4. (Case 2) Crushing injury to the skull of a 6-year-old. Diagrams of (a) clinically observed fractures and (b) an overlay of the highest 20% of principal tensile stresses produced fi'om a FE model. White cranial margins represent contact of applied forces (ground and tire) ..................... 60 Figure 4.5. (Case 3) Crushing injury to the skull of a 3-year-old. Diagrams of (a) clinically observed fractures and (b) an overlay of the highest 20% of principal tensile stresses produced from a FE model. White cranial margins represent contact of applied forces (ground and tire) ..................... 60 Figure 4.6. (Case 4) Crushing injury to the skull of a 1.5 year old. Diagrams of (a) clinically observed fractures and (b) an overlay of the highest 20% of principal tensile stresses produced from a FE model. White cranial margins represent contact of applied forces (ground and tire) ..................... 61 Figure 5.1. Simplified cranial model with overlays of modeled structures on true structures. Coronal cut of CT used for thickness comparison. *Feature viii not modeled. TFeature simplified. #Suture width altered for consistency with material pr0perty data obtained previously. Dotted line represents location of frontal bone removal .................................................................. 71 Figure 5.2. Overlay of maximum principal tensile stresses from FE models on representative experimental cranial fractures for (a) rigid interface impact to a young specimen, (b) rigid interface impact to an old specimen, (c) compliant interface impact to a young specimen, and (d) compliant interface impact to an old specimen. Longer lines represent higher stresses ............................................................................................... 73 Figure 5.3. Overlay of maximum principal tensile strains from FE models on representative experimental cranial fiactures for (a) rigid interface impact to a young specimen, (b) rigid interface impact to an old specimen, (c) compliant interface impact to a young specimen, and (d) compliant interface impact to an old specimen. Longer lines represent higher strains ................................................................................................ 74 ix RESEARCH PUBLICATIONS PEER-REVIEWED MANUSCRIPTS Baumer T, Nashelsky M, Hurst C, Passalacqua N, Fenton T, and Haut R. Characteristics and Prediction of Cranial Crush Injuries in Children (in review) Baumer T, Passalacqua N, Powell J, Newben'y W, Smith W, Fenton T, and Haut R. Age-Dependent Fracture Characteristics of Rigid and Compliant Surface Impacts on the Infant Skull - A Porcine Model. J. Forensic Sciences. 2009 (in press) Baumer T, Powell B, Fenton T, and Haut R. Age Dependent Mechanical Properties of the Infant Porcine Skull and a Correlation to the Human. J Biomechanical Eng. 2009 (in press) Meyer E, Baumer T, Slade J, Smith W, Haut R. Tibiofemoral contact pressures and osteochondral microtrauma during ACL mpture due to excessive compressive loading and internal torque of the human knee. Am J Sports Med. 2008; doi: 10.1 177/0363546508318046 PEER-REVIEWED ABSTRACTS Baumer TG, Powell BJ, Fenton TW, Haut RC. Age Dependent Mechanical Properties of the Infant Porcine Skull and a Correlation to the Human. Am. Society of Mechanical Engineering, Lake Tahoe, CA. 2009. Fenton TW, Passalacqua NV, Baumer TG, Powell BJ, Haut RC. A Forensic Pathology Tool to Predict Pediatric Skull Fracture Patterns - Part 1: Investigations on Infant Cranial Bone Fracture Initiation and Interface Dependent Fracture Patterns. Am. Academy of Forensic Sciences (AAFS), Denver, Colorado, 2009. [Winner of the Ellis R. Kerley Award] Nashelsky MB, Fenton TW, Passalacqua NV, Hurst CV, Baumer TG, Haut RC. Cranial Fracture Patterns in Pediatric “Crushing” Injuries and Preliminary Biomechanical Modeling Using a Simple Finite Element Model. AAFS, Denver CO, 2009. Meyer EG, Baumer TG, Haut RC. Knee joint relative motion during ACL rupture by internal tibial torsion of tibiofemoral compression. North Am. Congress on Biomechanics (NACOB), Ann Arbor, Michigan, 2008. Meyer EG, Baumer TG, Haut RC. Tibiofemoral contact pressures and osteochondral microtrauma from ACL mpture via hyperextension and joint compression. NACOB, Ann Arbor, MI, 2008. Meyer EG, Baumer TG, Haut RC. Tibiofemoral contact pressures and osteochondral microtrauma during ACL rupture due to excessive compressive loading and internal tibial torsion. NACOB, Ann Arbor, MI, 2008. CHAPTER ONE CHARACTERISTICS OF THE DEVELOPING SKULL - AN INTRODUCTION Skeletal fi’actures are diagnosed in nearly 1 of 3 children investigated for physical abuse with eighty percent of all fractures from abuse occurring in children less than 18 months of age (Belfer et al, 2001). Head injuries account for 80% of fatal child abuse (Case et al, 2001), while short falls rarely, if ever, cause serious injury or death (Reiber, 1993). However, non-fatal injury from short falls (less than 3 feet) can still result in skull fracture especially for infants less than 12 months of age (Gruskin and Schutzman, 1999). Distinguishing between abuse and accidental trauma can be very difficult since both types of trauma may produce similar types of injuries (Billrnire and Myers, 1985). Specifically, linear, complex, and depressed fiactures can be seen in both types of cases (Reece and Sege, 2000; Wheeler and Shope, 1997). Some attempts have been made to create a model of the developing skull, but data are limited. One modeling effort involves scaling of the adult skull, which has met with some success in predicting impact response of the pediatric skull (Prange et al, 2004). However, the head of an infant is smaller and geometrically unlike that of an adult (Schneider et al, 1986) and the validity of predicting skull fracture patterns in infants from adult data has not been investigated. Using adult data to predict skull fracture patterns in the pediatric skull may also be problematic due to the different structural (ease of deformation and decreased threshold to fracture) and mechanical properties of the infant skull (Thibault and Margulies, 1998; Prange et al, 2000; Prange and Margulies, 2002). The developing skull consists of several bones joined by soft tissue (sutures) that allow growth of the cranium during childhood (Figure 1.1). These bones are developed during fetal life and grow rapidly over the first two years after birth, with decreasing growth rates into adulthood (Mertz, 1984). The anterior fontanelle closes between 9 and 26 weeks after birth, but is not sealed until around 18 months (Knight, 1991). In adulthood, the sutures of the skull ossify creating one solid structure, whereas the sutures of infants result in more isolated regions. Frontal Bones Metopic Suture Anterior Fontanelle oronal Suture Sagittal Suture Parietal Bones Posterior Fontanell Occipital Bone Lambdoid Suture Figure 1.1. Diagram of a newborn skull. While several studies have documented important information regarding the response of the infant skull and brain to blunt force impact, these data are limited to one or two age groups. And because of ethical considerations, experimentation on human pediatric specimens is very limited. However, one animal model of the human infant has begun to emerge regularly in the current literature. A study by Dickerson and Dobbing (1967) shows a similarity in development of the central nervous system when correlating months of the human to weeks of the pig. Other studies have used porcine models to predict fracture loads for the infant human femur (Pierce et al., 2000) and strain in the braincase and sutures (Herring and Teng, 2000). An important finding in studies on the porcine infant cranial bone and suture is “that the elastic modulus, rupture modulus, and energy absorbed to failure of infant (2-3 day old) cranial (parietal) bone are similar to that of the human infant in three-point bending” (Margulies and Thibault, 2000). While Margulies and Thibault (2000) have attempted to correlate the mechanical behavior of human infant cranial bone to porcine infant cranial bone, the ages and number of specimens tested were limited. The data shows a linear increase in mechanical properties with age, but includes prenatal and postnatal aged specimens. Since questions of abuse in the pediatric patient typically involve subjects up to 18 months of age, which correlates to a time of rapid skull growth, the biomechanical properties of developing skull bones and sutures must be well documented over this entire period of developmental time if an accurate age-dependent model is to be created. Yet, the age-dependent properties of the developing skull are not the only conditions affecting fracture patterns. The risk of injury is also dependent on the contacting surface and fall height (Bertocci et al., 2003; Chalmers et al., 1996). Other variables, such as the area struck, thickness of the skull, thickness of scalp and hair, and impact direction can also affect the fracture pattern (Knight, 1991; Cooperrnan and Merten, 2001). Head injuries are often the result of dynamic forces in which a moving object strikes the cranium at a considerable speed or the cranium makes contact with a static object that causes the skull to be rapidly decelerated (Duhaime et al., 1995). However, cases of crushing injuries, where a static load is applied relatively slowly to the cranium, are also present in the literature. Unfortunately, due to the paucity of skull fracture tolerance data for infants and young children, pediatric trauma involving single-event head injuries with related cranial fractures represent one of the greatest challenges to forensic pathologists and anthropologists. Since it is often the job of the forensic professional to try to recognize characteristic patterns in these cases, a clear understanding of fracture propagation, and the effects of the aforementioned variables, becomes vital. Studies involving impact loading of the cranium have been presented, but there are also several case reports of basilar skull damage in the literature involving both crushing and impulse based loading scenarios mainly focusing on adult crania. A study involving 15 cases of crushing injuries to adult skulls aged 14-63 years and experimentally induced fiactures on 11 cadaver crania revealed similarities in basilar fracture patterns (Russell and Schiller 1949). However, a study by Harvey and Jones (1979) claimed that fractures through both petrous bones was not an indicator of impact site. Another study by Yasue (1981) then suggested that the overall path of fracture through the basicranium could, in fact, be related to the region of the skull impacted. While it is assumed that tinctures will generally follow the path of least resistance, this path may be different between the child and adult cranium. While all of these aforementioned studies focused on adult crania, some recent case reports on quasi-static loading to the crania of children have also presented observations of basilar fiactures (Takeshi, 2005; Duhaime et al., 1995). The trend of quasi-static loading producing basilar fractures in both adults and children may suggest that prediction of these fractures is possible. Theoretical techniques are widely used for prediction of material failure in structures. The Finite Element (FE) method is a powerful analytical modeling tool that can be used to determine internal stresses and strains within a complex solid. Several studies have shown that maximum principal stress (or strain) directions from a modeled structure correlate well with areas of experimentally or clinically observed bone fractures (Graham 2000; Silva 1998; Doorly 2006). While modeling of fiacture propagation generally involves high computational costs, one theory suggests that there is a close, although not exact, relationship between pre-failure stress trajectories and the ultimate path of propagation. In brief, Frank and Lawn (1967) proposed that crack growth occurs in such a way as to maximize the energy dissipation per unit length of growth, and that the greatest amount of dissipation occurs when the crack plane is perpendicular to the direction of greatest principal tensile stress. In other words, the directions of principal tensile stress developed on a solid may be a reasonable predictor of propagation direction. Skull fracture interpretation is a necessary component of forensic analysis and contributes to the determination of the cause and manner of injury or death. The research presented in this thesis makes use of an in situ porcine model and finite element analysis to examine the responses of the developing skull to blunt force impulse and quasi-static crushing trauma. Chapter 2 documents age effects on the mechanical properties of parietal bone and coronal suture in porcine infants and correlates the bending properties of the bone to existing human infant data. Chapter 3 documents skull fi-acture on infant porcine specimens with respect to age and interface in an effort to identify fracture characteristics for a developing skull under known conditions. This study hypothesized -A.. that impact energy required to initiate a skull fracture would depend on specimen age, and the extent of fracture damage for a given energy impact would depend on impact interface. Chapter 4 uses a simplified cranial model and the finite element method to develop representative patterns of principal stress and strain directions developed during quasi-static crushing injuries to the developing skull. The theoretical results were compared to four clinical cases of known childhood fatalities from crushing head injuries inflicted by vehicle tires. Chapter 5 uses a simplified porcine cranial model and the finite element method to predict experimentally documented fracture patterns developed dming impact loading to the parietal bone. The material properties of the cranial model were taken from the previously obtained results from Chapter 2. The experimentally inflicted fractures to the cranium from Chapter 3 were used for comparison with the theoretically developed principal stress and strain directions. The research presented in this thesis provides useful data in regards to the material properties of the developing porcine skull, and has suggested a correlation with the developing human skull. This research also documents fractures generated on the developing skull due to blunt force trauma with respect to age and interface. The experimentally and clincally observed fi'actures were predicted using a simplified theoretical model of impact and quasi-static loading of the cranium. This information may prove relevant in determining perimortem events in forensic cases of pediatric cranial injuries with unknown or questionable circumstances. These data may have ultimate utility in the study of potential child abuse cases. REFERENCES Belfer R, Klein B, and Orr L., 2001, “Use of the Skeletal Survey in the Evaluation of Child Maltreatment,” American Journal of Emergency Medicine, 19, pp. 122—4. Bertocci G. and Pierce M., 2006, “Applications of Biomechanics Aiding in the Diagnosis of Child Abuse,” Clinical Pediatric Emergency Medicine, 7, pp. 194-9. Billmire M. and Myers P., 1985, “Serious head injury in infants: accident or abuse?,” The American Academy of Pediatrics, 75, pp. 340-2. Case M, Graham M, Handy T, Jentzen J, Monteleone J ., 2001, “Position Paper on Fatal Abusive Head Injuries in Infants and Young Children,” The American Journal of Forensic Medicine and Pathology, 22, pp. 112-22. Chalmers D, Marshall S, Langley J, Evans M, Brunton C, Kelly A, et al., 1996, “Height and surfacing as risk factors for injury in falls from playground equipment: a case- control study,” Injury Prevention, 2, pp. 98-104 Cooperman D. and Merten D., 2001, “Skeletal Manifestation of Child Abuse,” In: Child Abuse: Medical Diagnosis and Management. Philadelphia: Lippincott, Williams, and Wilkins, pp. 123-56. Dickerson J. and Dobbing J ., 1967, “Prenatal and postnatal growth and development of the central nervous system of the pig,” Proceedings of the Royal Society of London, 166(1005), pp. 384-95. Doorly M, Gilchrist M., 2006, “The use of accident reconstruction for the analysis of traumatic brain injury due to head impacts arising from falls,” Computer Methods in Biomechanics and Biomedical Engineering, 9(6), pp. 371-7. Duhaime A-C, Eppley M, Margulies S, Heher K, Bartlett S., 1995, “Crush Injuries to the Head in Children,” Neurosurgery, 37(3), pp. 401-7. Frank F, Lawn B., 1967, “On the Theory of Hertzian Fracture,” Proceedings of the Royal Society of London, 299(1458), pp. 291-306. Graham R, Oberlander E, Stewart J, Griffiths D., 2000, “Validation and use of a finite element model of C-2 for determination of stress and fracture patterns of anterior odontoid loads,” Journal of Neurosurgery (Spine 1), 93, pp. 117-25. Gruskin K. and Schutzrnan S., 1999, “Head Trauma in Children Younger Than 2 Years,” Archives of Pediatric Adolescent Medicine, 153, pp.15-21. Harvey P, Jones A., 1980, “‘Typical’ Basal Skull Fracture of Both Petrous Bones: An Unreliable Indicator of Head Impact Site,” Journal of Forensic Sciences, 25(2), pp. 280- 6. Herring S. and Teng S., 2000, “Strain in the Braincase and Its Sutures During Function,” American Journal of Physical Anthropology, 112, pp. 575-593. Knight 3., 1991, “Fatal child abuse,” In: Forensic Pathology. London: Edward Arnold, pp. 457-73. Mertz H., 1984, “A Procedure for Normalizing Impact Response Data,” SAE #840884, 29th Stapp Car Crash Conference, Chicago, IL. Pierce M, Valdevit A, Anderson L, Inoue N, and Hauser D., 2000, “Biomechanical Evaluation of Dual-Energy X-Ray Absorptiometry for Predicting Fracture Loads of the Infant Femur for Injury Investigation: An In Vitro Porcine Model,” Journal of Orthopaedic Trauma, 14(8), pp. 571-6. Prange M, Luck J, Dibb A, Van Ee C, Nightingale R, Myers B., 2004, “Mechanical Properties and Anthropometry of the Human Infant Head,” Stapp Car Crash Journal, 48, pp. 279-99. - Prange M, Meaney D, Margulies S., 2000, “Defining brain mechanical properties: Effects of region, direction, and species,” Stapp Car Crash Journal, 44, pp. 205-213. Prange M. and Margulies S., 2002, “Regional, Directional, and Age-Dependant Properties of the Brain Undergoing Large Deformations,” Journal of Biomechanical Engineering, 124(2), pp. 244-52. Reece R and Sege R., 2000, “Childhood Head Injuries: Accidental or Inflicted,” Archives of Pediatric Adolescent Medicine, 154, pp. 11-15. Reiber G., 1993, “Fatal falls in childhood: How Far Must Children Fall to Sustain Fatal Head Injury?,” American Journal of Forensic Medicine and Pathology, 14, pp. 201-7. Russell W, Schiller F ., 1949, “Crushing Injuries to the Skull: Clinical and Experimental Observations,” Journal of Neurology, Neurosurgery, and Psychiatry, 12, pp. 52-60. Schneider L, Lehman R, Pflug M, Owings C., 1986, “Size and Shape of the Head and Neck From Birth to Four Years,” Washington, DC. The Consumer Product Safety Commission, Report No.: UMTRI—86-2. Silva M, Keaveny T, Hayes W., 1998, “Computed Tomography-Based Finite Element Analysis Predicts Failure Loads and Fracture Patterns for Vertebral Sections,” Journal of Orthopaedic Research, 16, pp. 300-8. Takeshi M, Okuchi K, Nishiguchi T, Seki T, Watanahe T, Ito S, et al., 2006, “Clinical Analysis of Seven Patients of Crushing Head Injury,” The Journal of Trauma-Injury, Infection, and Critical Care, 60(6), pp. 1245-9. Thibault K and Margulies S., 1998, “Age-Dependant Material Properties of the Porcine Cerebrum: Effect on Pediatric Head Injury Criteria,” Journal of Biomechanics, 31(12), pp. 11 19-26. Wheeler D. and Shope T., 1997, “Depressed skull fracture in a 7-month old who fell from bed,” Pediatrics, 100, pp. 1033-4. Yasue M., 1981, “Clinical and experimental investigation of mediobasal skull fracture,” Neurol Med Chir (Tokyo), 21, pp. 1041-9. CHAPTER TWO AGE DEPENDENT MECHANICAL PROPERTIES OF THE INFANT PORCINE PARIETAL BONE AND A CORRELATION TO THE HUMAN ABSTRACT An infant less than 18 months of age with a skull fracture has a 1 in 3 chance of abuse. Injury biomechanics are often used in the investigation of these cases. In addition to case-based investigations, computer modeling, and test dummies, animal model studies can aid in these investigations. This study documents age effects on the mechanical properties of parietal bone and coronal suture in porcine infants and correlates the bending properties of the bone to existing human infant data. Three beam specimens were cut from porcine specimens aged 3, 7, 10, 14, 18, and 21 days: one across the coronal suture and two from the parietal bone, one parallel to and one perpendicular to the coronal suture. An actuator-mounted probe applied 4-point bending in displacement control at 25nmr/sec until failure. Bending stiffness of bone specimens increased with age; bone-suture—bone specimens showed no change up to 14 days, but increased from 14 to 18 days. All three specimen types showed decreases in ultimate stress with age. Ultimate strain for the bone-suture-bone specimens was significantly higher than that for the bone specimens up to 14 days, with no differences thereafter. There was no change in the bending modulus with age for any specimen type. Bone-suture-bone bending modulus was lower than that of the bone specimens up to 14 days, with no differences thereafter. There was no change in strain energy to failure with age for the bone specimens; bone-suturebone specimens showed no change up to 14 days, but decreased from 14 to 18 days. There was an increase in specimen porOsity with age. Correlation 10 analysis revealed a weak (-0.39), but significant, negative correlation between ultimate stress and porosity. While the mechanical properties of parietal bone and coronal suture did not change significantly with age, bone specimens showed an increase in bending stiffness with age. Bone-suture—bone specimens showed an increase in bending stiffiress only between 14 and 18 days of age. Correlation analyses using existing and new data to compute the bending rigidity of infant parietal bone specimens suggested that days of pig age may correlate with months of human age during the most common time frame of childhood abuse cases. Such a correlation may support the developing porcine model as an important biomechanical tool for the study of pediatric skull fiacture in efforts to investigate cases of childhood abuse. 11 INTRODUCTION Skeletal fractures are diagnosed in nearly 1 of 3 children investigated for physical abuse (Belfer et al., 2001). Often it is the job of the forensic anthropologist to try to recognize characteristic patterns in these cases, but in reality it is difficult to diagnose abuse simply based on a fracture. In fact, age together with the type and site of tincture are important features to consider when assessing cases of potential abuse (Kemp et al., 2008). Eighty percent of all fiactures from abuse are seen in children less than 18 months of age. For these children the probability of abuse given a skull fracture, for example, is 0.30 (range 0.19 to 0.46). The most common site of skull fracture in abuse and non-abuse cases is the parietal bone (Meservy et al., 1987), with the most common fracture type being linear (Reece and Sege, 2000). Injury biomechanics are often used in the investigation of cases suspected to involve child abuse (Bertocci and Pierce, 2006). In addition to case-based investigations, computer modeling, and the use of test dummies, animal model studies can aid in these investigations. While the relationship between animal studies and the human pediatric patient is yet unclear in most circumstances, one animal model seems to have regularly emerged in the current literature. For example, a study has been conducted on prediction of fracture loads for the infant femur using porcine models aged 3-12 months (Pierce et al., 2000). In another study on strain in the braincase and its sutures, juvenile pigs have again been used as a model of the human infant (Herring and Teng, 2007). Dickerson and Dobbing (1966) have previously compared the development of the porcine central nervous system to the human and found that months of life in the human are roughly comparable to weeks in a pig. An important finding in studies on the porcine infant 12 cranial bone and suture is “that the elastic modulus, rupture modulus, and energy absorbed to failure of infant (2-3 day old) cranial (parietal) bone are similar to that of the human infant in three-point bending” (Margulies and Thibault, 2000). While Margulies and Thibault (2000) have attempted to correlate the mechanical behavior of human infant cranial bone to porcine infant cranial bone, the ages and number of specimens tested were limited. The data shows a linear increase in mechanical prOperties with age, but the data includes prenatal and postnatal aged specimens. Other studies have shown that growth of the human skull involves large increases in size over the first two years of life, with decreasing growth rates into adulthood (Mertz, 1984). Since questions of abuse in the pediatric patient typically involve subjects up to 18 months of age, an objective of the current study was to document the biomechanical properties of developing porcine skulls and sutures over a longer period of developmental time than previous studies. Additionally, Margulies and colleagues have documented the biomechanical properties of the infant porcine parietal bone using beam specimens oriented perpendicular to and across the coronal suture, but case-based investigations show linear skull fractures in infants typically initiate at an ossification center and run parallel to and perpendicular to spiculae in the parietal bone (Holck, 2005). It has also been shown that the mechanical properties of prenatal human infant parietal bone is non-isotropic (McPherson and Kriewall, 1980). Therefore in the current study, beam specimens were taken from the developing porcine parietal bone both parallel and perpendicular to the coronal suture, as well as across it. The data fiom the current study will provide 13 information on the age factor for developing parietal bone and coronal suture in the porcine infant model, as it may relate to the human pediatric patient. MATERIALS AND METHODS Porcine specimens aged 3 days (n=5), 7 days (n=8), 10 days (n=7), 14 days (n=5), 18 days (n=5), and 21 days (n=4) were obtained from a local supplier and stored at —20°C until testing. All animals died of natural causes and the specimens were frozen within 12 hrs of death. None of the specimens had outward signs of head injury, which was later confirmed during specimen preparation. Each specimen was thawed at room temperature for 24 hours prior to testing. The scalp was removed, but the periosteum was lefi intact to help prevent drying of the bone and sutures. The top palette of the skull was removed using a surgical saw (Autopsy Saw, Stryker, Kalamazoo, MI). Cuts were made at the larnbdoid suture, along the lateral aspect of the parietal bone, and diagonally across the home] bone from the lateral aspect of the coronal suture to an anterior location along the sagittal suture. Three beam specimens were cut from each skull: one across the coronal suture and two from the parietal bone, one parallel to and one perpendicular to the coronal suture (Figure 2.1). The specimens were cut with a bone saw (Isomet 1000, Buehler, Lake Bluff, IL), while bathed in phosphate buffered saline (PBS). The specimen width varied from 4.5 to 5.5 mm wide. Due to the non-uniform thickness of the beam specimens, three thickness measurements were taken from each specimen at equally spaced intervals along the beam. These measurements were used to calculate an average thickness of each specimen. The length of the beam specimens was fixed at 15 mm to 14 yield a length to thickness ratio of approximately 8:1. The cut specimens were stored in room-temperature PBS for less than 1 hour before testing. Sagittal Suture Bone-Suture- Bone Coronal Suture Parallel Perpendicular Occipitall Figure 2.1. TOp view of cranium with locations of the beam specimens. The beam specimens were loaded to failure in 4-point bending. Each specimen was mounted between two aluminum sleeves using air-hardening cement (Technovit, J orgensen laboratories Inc., Loveland, CO). The fixed testing length of 15 mm was maintained with a thin aluminum bar fastened between the sleeves. This bar also served to prevent damage to the specimen during preparation and mounting in the test fixture. The potted specimens were mounted in a custom-designed 4-point bending fixture (Figure 2.2) attached to a servo-hydraulic testing machine (Model 1331, Instron Corp., Canton, OH). One end of the fixture was mounted on linear bearings to allow displacement along the specimen axis. An actuator-mounted probe was located on each of the two sleeves to apply 4-point bending. A 111.2-N load-cell (Model #JP25, Data Instruments Inc., Acton, MA) was mounted between the probe and actuator. The experiments were run with the machine in displacement control at 25mm/sec to generate the bending moment on the beam specimens. 15 Linear Actuator Load Cell Probe O _l:::|_ O [: Beam ] Specimen i c— Translation —> i | J Figure 2.2. Diagram of testing fixture. The force and actuator displacement were recorded to failure during each experiment. From these values, initial specimen stiffiress, ultimate stress (cult): ultimate strain (a uh), bending modulus (E), and strain energy to failure (U) were determined using the following information. Initial stiffiress was calculated from the initial linear portion of the force-actuator displacement plot. Ultimate stress at the endocranial surface was determined fiom simple beam theory using M* _ 3’ Cult _ 1 f where M. was the bending moment at failure, y was the distance from the inner surface (2.1) of the specimen to the centroidal axis, and If was the second moment of the rectangular area measured at the fiacture location. Ultimate strain at the fracture location was calculated from s = —- (2.2) where y was the distance from the inner surface of the specimen to the centroidal axis, or was the angular rotation of the beam under the probes at failure, and L was the length of the specimen. The elastic modulus was also determined fi‘orn simple beam theory using __rr L __ (23) a 21 av E g where War was the moment-angular rotation ratio during the initial linear region of the force-displacement trace, and Iavg was the second moment of area determined using the average thickness over the length of the beam specimen. The failure strain energy was then computed with the following integral equation 2 LENF) U: I2E1 0 avg where ‘x’ was the distance along the beam from the stationary aluminum cylinder. dx (24) The gross morphological features of the beams were documented by staining the length-wise cross-section of the failed specimens with India ink (Figure 2.3). Each test specimen was photographed with a digital camera mounted to a dissecting microscope at 50x magnification (Wild M5A, Wild Heerbrugg, Switzerland). A representative section of bone was cropped fi'om the exocranium to the endocranium that included an equal distance along the length of the beam. This image was always taken in an area between characteristic undulations in the specimen thickness, as this was the area of failure during the experiments. The image was imported to an image analysis program (Sigma Scan, J andel Scientific, San Rafael, CA) and converted to grayscale. The porosity was determined as the area of the stain (voids) per total area. 17 Compact Layer Porous Layer Figure 2.3. Representative stained cross-section of a beam from a 14 day specimen showing separation of compact and porous layers. Voids in the bone matrix were colored black. The box represents the analyzed area of a test specimen. Each material property data set was analyzed for age effects using linear regression analyses. Comparisons between the three specimen types was performed using a two-factor (age, type) AN OVA. A Pearson Product Moment correlation was used to analyze the effect of porosity on ultimate stress. Statistically significant effects were reported for p<0.05. RESULTS There was no difference in average thickness measurements between the bone and bone-suture-bone specimens for any age. Linear regression analyses also showed a significant increase in thickness with age (p<0.001 for each specimen type). For each orientation of parietal bone specimen, there was a significant increase in bending stiffiress with age (Figure 2.4a). The slopes were not statistically dependent on orientation. Interestingly, there was no significant change in bending stiffness in the bone-suture-bone specimens up to fourteen days, but between fourteen and eighteen days there was a significant increase (Figure 2.4b). The stiffness of the suture was lower than that of the bone up to fourteen days, but it became statistically greater than the parallel bone specimens at eighteen days (suture 1197i246 N/m; parallel 864d:367 N/m). 18 2500 — - Perpendicrlar a Parallel 2000 4 n 1500 - 1000 - Stiffness (N/m) UT :3 53 r O (a) O) (D 12 15 18 21 Age (days) A m v 2000 - _s .3 N O) C O O C r r 800 - 400 . Stiffness (N/m) 3 7 10 14 18 21 (b) Age (Days) Figure 2.4. Recorded values of stiffness for (a) both orientations of bone and (b) bone- suture-bone specimens from 3 to 21 days. *Significantly different means. All bone specimens failed at a thinned area and all bone-suture-bone specimens failed in the suture. Linear regression analyses showed significant decreases in ultimate stress for both orientations of bone (p<0.05) and for the bone-suturebone specimens (p<0.01) with age (Figure 2.5). The slopes were not different. There was no change with age in ultimate strain in the bone (0.0163:I:0.0053 mm/mm) and bone-suture-bone (0.0313i0.0072 mm/mm) specimens (Figure 2.6). The ultimate strain for the bone-suture-bone specimens was significantly higher than that for the bone specimens up to fourteen days. 19 1000 - o Suture E I I Perpendicular E 300 ‘ a Parallel g 600 - a g 400 - g 200 - :42 3 o 0 3 6 9 12 15 18 21 Age (days) Figure 2.5. Linear regressions of ultimate stress for bone and bone-suture-bone specimens against age. 0.06 - * * FL I Suture ’T‘ T ‘1" I Perpend. .g 0-05 ' re "" |'_'l Parallel 5 0.04 - 03 .. g 0.03 E 0.02 - 3 0.01 - 0 - 3 7 10 14 18 21 Age (Days) Figure 2.6. Ultimate strain for bone and bone-suture-bone specimens. *Significantly different means. There was no change with age in the bending modulus of the bone (6.01:1:1 .73 GPa) and bone-suture-bone (2.731084 GPa) specimens (Figure 2.7). The bending modulus of the bone-suture-bone specimens was significantly lower than that of the bone specimens up to fourteen days. 20 I Suhte ,, I Perpend. "H * * El Parallel Bending Modulus (GPa) 3 7 1 0 14 1 8 21 Age (Days) Figure 2.7. Bending modulus for bone and bone-suture-bone specimens versus age. *Significantly different means. There was also no change with age in the strain energy to failure for the bone specimens, being 0.03:l:0.03 J and 0.03:1:004 J for perpendicular and parallel orientations, respectively (Figure 2.8). The bone-suture-bone specimens showed no change up to fourteen days (02310.13 J). For ages up to fourteen days, the strain energy to failure of the bone-sutrrre-bone specimens was significantly higher than that of the bone specimens. From fourteen to eighteen days, however, there was a significant decrease in the strain energy of the bone-suture-bone specimens. Therefore, the strain energy to failure was no longer different for the bone (0.01:1:000 J) versus bone-suture- bone specimens (0.01:l:0.00 J). The morphological characteristics of the skull bone fiom the infant porcine specimens changed during the aging process. It was found that two distinct layers, one that was visually quite compact and another more porous layer, were present as early as three days. The porous layer appeared as pockets beneath the compact layer which 21 developed through ten days, and by fourteen days there was a clear separation of zones along the entire specimen (see Figure 2.3). Strain Energy (J) 0.6 - 0.5 d 0.4 - 0.3 - 0.2 - 0.1 - 0.. I Sutue * * I Perpend. T T t * El Parallel re A 3 7 1o 14 1a 21 Age(DaVS) Figure 2.8. Stain energy of bone and bone-suture bone specimens versus age. *Significantly different means. Analysis of both orientations of bone specimens revealed an increase in porosity with age (p<0.001) (Figure 2.9). Correlation analysis of the combined data revealed a weak (-0.391), but significant (p<0.01), negative correlation between ultimate stess and porosity. No other mechanical property had a significant correlation with porosity. ItY Poros 20% - 10% - 0% 40% - I Perpendicular :- Parallel 30% - ' 0 3 6 9 12 15 18 21 Age(Days) Figure 2.9. Linear regression analyses of both orientations of infant porcine bone specimens against age. Both orientations displayed significant dependencies on age. 22 DISCUSSION The current study was performed to document potential changes in the mechanical properties of infant porcine parietal bone as a function of age. Linear regression analyses on the bending modulus of porcine specimens showed no significant change with age. A few previous studies on human specimens have shown a linear increase in the bending modulus of prenatal human infant parietal bone and it has been suggested that this tend may continue up to six months (Margulies and Thibault, 2000), or even six years (McPherson and Kriewall, 1980) of age. These studies, however, assumed a linear increase from the prenatal data to the older specimens and did not collect data between the endpoints. A study by Coats and Margulies (2006) investigates the mechanical properties of both prenatal and postratal infants from twenty-six weeks of gestation to thirteen months of age. The authors claimed that there was a significant increase in bending modulus with age, but a closer inspection of their data indicated that the increase in these properties was significantly influenced by the prenatal data. When linear regression analysis was performed on the postnatal data alone there was no statistical change in bending modulus with age. These data are in agreement with the current study using infant porcine skull bone. The bending modulus data tom the human parietal bone of Coats and Margulies (2006) have also been compared to the bending modulus of the porcine parietal bone. It was clear that the porcine parietal bone bending modulus was significantly higher than that of the human parietal bone. However, this is a material rather than a stuctural parameter of comparison. Since the current thought was to use the porcine model to study issues of potential childhood abuse, 3 stuctural comparison may be more 23 appropriate. In this effort we were able to obtain unpublished human infant parietal bone thickness data for ages two to twenty-one months, n=314, fi‘om a recent study by Li et al. (2007). While the human skull was significantly thicker than the porcine skull over the ages studied here, a linear regression of parietal bone thickness versus age indicated growth rates of parietal bone thickness for humans to be 0.0487 mm/month and that of pigs to be 0.0492 mm/day. Using bending modulus data from Coats and Margulies (2006) and thickness values obtained during the Li et al. (2007) study, the bending rigidity of human infant parietal bone versus age in months was compared to that of the porcine specimens versus age in days (Figure 2.10). The slopes of parietal bone bending rigidity versus age were 0.505 GPa*m4/month for humans and 0.596 GPa*m4/day for the porcine specimens. The difference between human and porcine data may be due to differences in the stain rates of testing. The human specimens were tested at a high velocity (approximately 2 m/sec), while the porcine specimens were tested at a lower 25 - 0 Human o Porcine Bending Rigidity Figure 2.10. Bending rigidity of human and porcine specimens versus age (Human — months; porcine - days). The human data was based on data from Coats and Margulies (2006) and unpublished data from Li et a1 (2007). 24 velocity (0.025 m/sec). While a previous study by Wood (1971) shows that the bending modulus of adult human cranial specimens between twenty-five and ninety-five years of age increases with increases in stain rate, the previous study by Coats and Margulies (2006) showed no stain rate dependency in bending modulus with age for specimens less than one month of age. Funn‘e studies may be needed to more firlly examine rate dependencies in both the human and porcine infant cranial bone during early developmental ages. The current study was also performed to document potential changes in the mechanical properties of the infant porcine coronal suture as a function of age. Margulies and Thibault (2000) show a significant difference between the bone-suture- bone and bone specimen mechanical properties for two to three day old porcine specimens. The data tom the current study was in agreement with this finding up to fourteen days. Specifically, in the current study, the bone-suture-bone specimens were less stiff, absorbed more energy to failure, and withstood larger bending deformation before failure than the bone specimens. The differences between bone-suture-bone specimens and bone specimens changed significantly in the current study between fourteen and eighteen days. During this period of development, the bone-suture-bone specimens underwent a significant increase in stiffness, a significant decrease in the energy absorbed to failure, and a decrease in stain at failure. These changes resulted in mechanical properties of the bone-suture-bone specimens that were more similar to those of the bone specimens at eighteen and twenty-one days. The explanation for this change between fourteen and eighteen days is currently unclear tom visual analysis of the specimen cross-sections. This could also be related to changes in the biochemical make- 25 up of the suture itself during this period. Some gross changes in the morphology of the porcine bone did appear to occur over the time period of the current study. One important feature was the development of two distinct layers of bone. Examination of each cross-section revealed isolated pockets of porous bone beneath an outer, more compact layer of bone in specimens as young as three days. By fourteen days, a clear separation between the compact and porous layers could be seen over the entire cross- section of these specimens. Previous studies on human (McElhaney et al., 1970; Behrens et al., 1974) and bovine (Martin, 1984) samples have shown that small increases in porosity of bone specimens can result in large decreases in their ultimate stess. The Coats and Margulies (2006) study also documents an increase in ultimate stess with specimen age, but our analysis of postnatal data alone out to thirteen months again revealed no statistical change with age. These observations warrant additional studies on the morphological development of human and porcine infant parietal bone. The current study also showed no significant difference in orientation on the bending properties of parietal beam specimens. This finding was in contast to previously published data for human specimens up to 40 weeks of gestation (McPherson and Kriewall, 1980). Interestingly, the aforementioned study described a distinct fiber orientation in both pre—term and term human infant parietal bones, but visual observation of the developing infant porcine skull did not show any distinct spiculae that might be related to a fiber orientation effect. While this finding seemed to reveal an inconsistency between infant human and porcine skulls, we were not able to find definitive information on when this visible fiber orientation effect disappears in humans. 26 In summary, the current investigation documented alterations in the mechanical properties of beam specimens tom the porcine infant skull with age. While it remains difficult to correlate these changes with those that may occur in the human pediatric patient, due to limited human pediatric data, the current study has attempted to make a correlation by combining existing and new, unpublished data. A good correlation appeared to exist for days of pig age and months of human age in terms of flexural rigidity of the beam specimens. Such a correlation may lend support to the notion that studies on the developing porcine model may provide an important biomechanical tool for the study of pediatric skull tacture in efforts to investigate cases of childhood abuse. 27 REFERENCES Behrens J, Walker P, and Shoji H., 1974, “Variations in Stength and Stucture of Cancellous Bone at the Knee,” Journal of Biomechanics, 7, pp. 201-7. Belfer R, Klein B, and Orr L., 2001 , “Use of the Skeletal Survey in the Evaluation of Child Malteatnent,” American Journal of Emergency Medicine, 19, p. 122—4. Bertocci G. and Pierce M., 2006, “Applications of Biomechanics Aiding in the Diagnosis of Child Abuse,” Clinical Pediatric Emergency Medicine, 7, pp. 194-9. Coats B. and Margulies S., 2006, “Material Properties of Human Infant Skull and Suture at High Rates,” Journal of Neurotrauma, 23(8), pp. 1222-32. Dickerson J. and Dobbing J ., 1966, “Some Peculiarities of Cerebellar Growth in Pigs,” Procwdings of the Royal Society of Medicine, 59, pp. 1088. Herring S. and Teng S., 2000, “Stain in the Braincase and Its Sutures During Function,” American Journal of Physical Anthropology, 112, pp. 575-93. Holck P., 2005, “What Can a Baby’s Skull Withstand? Testing the Skull’s Resistance on an Anatomical Preparation,” Forensic Science International, 151, pp. 187-91. Kemp A, Dunstan F, Harrison S, Morris S, Mann M, Rolfe K, Datta S, Thomas D, Sibert J, and Maguire S., 2008, “Patterns of Skeletal Fractures in Child Abuse: Systematic Review,” B.M.J., 337, pp. 1-8. Li H, Ruan J, Xie Z, Wang H, and Liu W., 2007, “Investigation of the Critical Geometric Characteristics of Living Human Skulls Utilising Medical Image Analysis Techniques,” Int. J. Vehicle Safety, 2(4), pp. 345-61. Margulies S. and Thibault K., 2000, “Infant Skull and Suture Properties: Measurements and Implications for Mechanisms of Pediatric Brain Injury,” Journal of Biomechanical Eng'neering, 122, pp. 364-71. Martin R., 1984, “Porosity and Specific Surface of Bone,” Critical Reviews in Biomedical Engineering, 10, pp. 179-222. McElhaney J, Fogle J, Melvin J, Haynes R, Roberts V, and Alern N., 1970, “Mechanical Properties of Cranial Bone,” Journal of Biomechanics, 3, pp. 495-511. McPherson G. and Kriewall T., 1980, “The Elastic Modulus of Fetal Cranial Bone: A First Step Towards an Understanding of the Biomechanics of Fetal Head Molding,” Journal of Biomechanics, 13, pp. 9-16. 28 Mertz H., 1984, “A Procedure for Normalizing Impact Response Data,” SAE #840884, 29th Stapp Car Crash Conference, Chicago, IL. Meservy C, Towbin R, McLaurin R, Myers P, and Ball W., 1987, “Radiographic Characteristics of Skull Fractures Resulting From Child Abuse,” A.J.R., 149, pp. 173-5. Pierce M, Valdevit A, Anderson L, Inoue N, and Hauser D., 2000, “Biomechanical Evaluation of Dual-Energy X-Ray Absorptiomety for Predicting Fracture Loads of the Infant Femur for Injury Investigation: An In Vito Porcine Model,” Journal of Orthopaedic Trauma, 14(8), pp. 571-6. Reece R. and Sege R., 2000, “Childhood Head Injuries: Accidental or Inflicted,” Archives of Pediatric Adolescent Medicine, 154, pp. 11-15. Wood J ., 1971, “Dynamic Response of Human Cranial Bone,” Journal of Biomechanics, 4, pp. 1-12. 29 CHAPTER THREE AGE-DEPENDENT FRACTURE CHARACTERISTICS OF RIGID AND COMPLIANT SURFACE IMPACTS ON THE INFANT SKULL — A PORCINE MODEL ABSTRACT Head injuries account for 80% of fatal child abuse in younger children. However, non- fatal injury fi'om short falls (less than 3 feet) can still result in skull fracture especially for infants less than 12 months of age. In many cases the ability to determine child abuse from accidental injury may be difficult. This study documents skull fracture on infant porcine specimens with respect to age and interface in an effort to identify fiacture characteristics for a developing skull under known conditions. A single impact causing fiacture was conducted on the skull of porcine specimens aged 2 to 28 days (n=76). Impacts were conducted using a gravity-accelerated mass to deliver a contolled energy impact to the parietal bone. Paired rigid and compliant impacts at the same energy were conducted at each specimen age. Impact force, impact duration, and fracture length were recorded. Pressure sensitive film was used to record contact areas and pressures for each impact. Fracture length was measured to the nearest millimeter. Impact force increased with age for both rigid and compliant interfaces. Contact area for compliant interface impacts increased with age but no change with age was noted for rigid interface impacts. Bone fi'actures initiated at bone-suture interfaces and not at the point of impact. Diastatic fractures were observed on skulls as early as 4 days of age after impacts with a compliant interface, while rigid interface impacts did not show suture damage until 10 days of age. Diastatic fractures were common for both interfaces on skulls aged 10 to 17 days, but disappeared after 17 days of age. For a given energy, impact of the skull with a 30 compliant interface caused more fracture damage than with a rigid interface for specimens less than 17 days of age, but less damage for specimens aged 24 to 28 days. The age-dependent response of the developing skull to impact with rigid and compliant surfaces is important in forensic cases for determining potential differences between testimony describing the events leading to injury and the observed fracture pattern. 31 INTRODUCTION Head injuries account for 80% of fatal child abuse in younger children (Case et al., 2001), while short falls rarely, if ever, cause serious injury or death (Reiber, 1993). In one study of 89 children under 2 years of age, 19 of the 20 fatalities were due to abuse (Hobb, 1984), while a study of 1,881 falls tom bed height reported no deaths (Belechri et al., 2002). However, non-fatal injury fi'om short falls (less than 3 feet) can still result in skull fiacture especially for infants less than 12 months of age (Gruskin and Schutzrnan, 1999). Due to the paucity of skull fracture tolerance data for infants and young children, pediatric tauma involving single-event head injuries with related cranial fractures represent one of the greatest challenges to forensic pathologists and anthropologists. Distinguishing between abuse and accidental tauma can be difficult, as both may produce similar types of injuries (Billmire and Myers, 1985). Specifically, linear, complex, and depressed fractures can be seen in both types of cases (Reece and Sege, 2000; Wheeler and Shope, 1997). The most commonly fractured cranial bone in both accidental injury and abuse cases is the parietal (Hobb, 1984; Meservy et al., 1987; Leventhal et al., 1993). The risk of injury is also dependent on the contacting surface and fall height (Bertocci et al., 2003; Chalmers et al., 1996). Other variables, such as the area stuck, thickness of the skull, thickness of scalp and hair, and impact direction can also affect the fracture pattern (Knight, 1991; Cooperman and Merten, 2001). To develop reliable models for skull fi'acture prediction, the effects of these aforementioned variables must be better understood. 32 While several studies have documented important information regarding the response of the infant skull and brain to blunt force impact, these data are limited to one or two age groups. And because of ethical considerations, experimentation on human pediatric specimens is very lirrrited. While scaling of the adult skull has met with some success in predicting impact response of the pediatric skull (Prange et al., 2004), the head of an infant is smaller and geometrically unlike that of an adult (Schneider et al., 1986) and the validity of predicting skull fracture patterns in infants from adult data has not been investigated. Using adult data to predict skull fracture patterns in the pediatric skull may also be problematic due to the different stuctural (ease of deformation and decreased threshold to fiacture) and mechanical properties of the infant skull (Thibault and Margulies, 1998; Prange et al., 2000; Prange and Margulies, 2002). However, one animal model of the human infant has begun to emerge regularly in the current literature. A study by Dickerson and Dobbing (Dickerson and Dobbing, 1967) shows a similarity in development of the cental nervous system when correlating months of the human to weeks of the pig. Other studies have used porcine models to predict fiacture loads for the infant human femur (Pierce et al., 2000) and stain in the braincase and sutures (Herring and Teng, 2000). More recently, age-related rates of change in the bending rigidity of the infant human and infant porcine skulls has been found to be similar for a correlation based on months in humans to days in pigs (Baumer et al., in press). The hypotheses of the current study were two-fold. First, impact energy required to initiate a skull fiacture would depend on specimen age. Second, the extent of fracture damage for a given energy impact would depend on impact interface. These data using a developing porcine model may provide a baseline for the characterization of fiacture 33 patterns resulting from blunt force impacts onto different contact interfaces for the developing human pediatric skull. These data may have ultimate utility in the study of potential child abuse cases. MATERIALS AND METHODS Porcine specimens aged from 2 to 28 days (n=76) were obtained tom a local supplier and stored at —20°C. All animals died of natural causes. The specimens were fiozen within 12 hours of death. None of the specimens had outward signs of head injury, which was later confirmed during specimen preparation. Each specimen was thawed at room temperature for 24 hours prior to testing. All soft tissue was removed from the left side of the skull, leaving the periosteum to help prevent drying of the bones and sutures. This side of each skull was placed in a bed of air-hardening epoxy (Fibre Stand, Martin Senour Corp., Cleveland, OH). Capturing the skull in this manner allowed deformation of the skull bones and sutures, but eliminated gross motions (tanslations and rotations) of the skull as a whole. After the potting material had hardened, the skull was bathed in phosphate-buffered saline (PB S) solution in regular intervals until the experimental impact. Each skull, with the scalp in place, was impacted in the center of the right parietal bone. The impact location was contolled by a custom-made, 4 degree-of-fi'eedom, lockable fixture that was built to hold the potted skull (Figure 3.1). The skull was oriented so that the impact was normal to the bone surface. 34 Impact Head Pressure Film Potting Positioning Fixture Figure 3.1. Skull positioning and set-up showing rigid impact head. Each specimen was subjected to a single impact with a gravity-accelerated mass (GAM) (Figure 3.2). A load tansducer (4.45 kN capacity, model AL311CV, Sensotec, Columbus, OH) was mounted on the GAM behind the impact interface to record forces. A single impact was assured by using an operational amplifier comparator circuit to monitor the impact forces and energize an electomagnetic solenoid to catch the irnpactor immediately after the load returned to zero. Two interchangeable impact interfaces were created for the GAM to represent rigid and compliant impact interfaces. The rigid interface was a solid aluminum cylinder with approximately 16 cm2 of impact surface. The compliant interface was a deformable aluminum material (1.10 MPa crush strength Hexcel, Hexcel Corp., Stamford, CT), approximately 3 cm thick with a 16 cm2 impact surface, fastened to the rigid interface. Both interfaces had greater surface area than the resulting contact area on the skull. 35 Stainless Steel Shaft Vertical Mounted Trolley with Adjustable Mass (GAM) Impact Head Skull with Parietal Bone Exposed to Apparatus Positioning Fixture Figure 3.2. Drop test fixture set-up. Input energy was adjusted by changing the drop height of the 1.67 kg GAM. A second mass of 1.92 kg was necessary for specimens twenty-one days and older, as the drop height was limited in the laboratory. The mass of the impact interface for each impact was included in the total mass of the GAM, and the force data was checked for inertial effects. Preliminary data collected at several ages was used to estimate the input energy needed to initiate a fiacture using the rigid interface for each specimen age. In a few cases (n=7), the first impact with a rigid interface did not cause a fracture, so a second, higher energy impact was performed. Impacts were always conducted in pairs. All skulls impacted with the compliant interface were impacted at the same energy that caused fi’acture in the paired rigid impact on another skull of the same age. While the input energy was the same for rigid and compliant impacts at each age, the energy of impact required to cause fracture initiation had to be increased with specimen age. The forces and impact durations were recorded at 10,000 Hz. 36 Pressure sensitive film packets (Prescale, Fuji Film Ltd., Tokyo, Japan) were attached with tape to the skull at the impact site to measure contact area and pressures generated during impact. Low (0-10 MPa) and medium (10-50 MPa) range pressure films were stacked together and sealed between two sheets of polyethylene (0.04 mm thick) to prevent exposure to fluids during the impact (Atkinson et al, 1998). Using the previously documented procedure, the film was scanned and converted to pressure using a dynamically calibrated density-to-pressure scale. Only the low pressure film was analyzed in these experiments as the pressures were not high enough to record on the medium pressure film. Computed tomography (CT) scans were conducted for each skull after impact (625 rim-thick coronal slices). The width of the skull vault and average thickness of the parietal bone were obtained tom these scans using reconstuction software (Osirix version 3.0, Open Source General Public License). The width of the vault was measured at its widest point fi'om parietal bone to parietal bone. Average parietal bone thickness was determined from a slice near the center of the parietal bone. Visual inspection was conducted on each skull after the soft tissue and periosteum were removed from the impacted side. All suture damage and any visible bone fractures were recorded. Each specimen was then cleaned via standard anthropological procedures, removing all remaining soft tissue. Complete fracture diagrams and measurements were made for each specimen. Total length of the skull tactures was measured to the nearest millimeter with a flexible measuring tape, which conformed to the curvature of the skull. 37 The impact and pressure data were analyzed for age effects using linear regression analyses. Comparisons between the interfaces were performed using a two- factor (age, interface) AN OVA. Statistically significant effects were reported for p<0.05. RESULTS Linear regression analyses showed a significant increase (0.0516 mm/day) in average parietal bone thickness with age (p<0.001), but no change in skull vault width with age. The input energy required to cause fracture increased tom 1.5 J at two days of age to 11.5 J at twenty-eight days of age. A characteristic rapid drop in force appeared in the force-time plot for all rigid interface impacts that were associated with fiacture (Figure 3). In rigid interface impacts that did not cause fiacture, a more pronounced peak was observed. Rapid drops in force during compliant interface impacts could not always be associated with fiactures due to the deformation of the interface. 500 - N0 ../ 400 - mm" 2 300 ~ (D 8 LL 200 - 100 _ Fracture o I f I 0 0.01 0.02 0.03 Time (see) Figure 3.3. Overlay of force-time plots showing typical characteristics of an impact that caused fracture and one that did not. 38 For both impact interfaces, linear regression analyses showed a significant increase in impact force with age that was required to produce a fi'acture (p<0.001) (Figure 3.4). The slopes and intercepts of impact forces versus age were not different based on interface. 1600 - E 1200 - (D 2 1‘2 800 - *5 S g 400 - o I l I 0 10 20 30 Age (days) Figure 3.4. Peak force during impact for rigid and compliant interfaces. The area of contact for the compliant interface significantly increased with age (p<0.001), but no significant change was noted for the rigid interface (Figure 3.5). 1600 - 9 Rigid a Hexcell 0 NA E1200 - a 3 ° ° 5 800- , a g ,.‘D' a c 400 - a, ' n 8 e on Us 0 Age (daYS) Figure 3.5. Contact area recorded on pressure sensitive film during impacts for rigid and compliant interfaces. 39 Bone fractures for all ages, and for both interfaces, occurred at sites away from the point of impact (Figure 3.6). These bone fractures initiated at a suture-bone interface. Frequently, bone fractures were accompanied by diastatic fiactures that were generally located near the fi'acture tip at the bone-suture interface. Diastatic fiactures caused by a rigid interface impact were not present until specimens were ten days of age, while diastatic fractures caused by a compliant interface were present as early as four days of age (Figure 3.7). Diastatic fractures were common for both interfaces between ten and seventeen days. Neither interface showed suture damage after seventeen days. a b i i I I} - Impact Site Fracture Figure 3.6. Representative fracture sites for (a) rigid and (b) compliant interface impacts (five-day—old specimens shown). Fracture sites initiated at bone-suture boundaries. Compliant interface impact showed increased fracture lengths at similar initiation sites as the rigid interface impacts as well as diastatic fracture of coronal suture. 40 Ri' 100- _ _ I gld DHexceIl § 72’ 75— 8 .l-‘f o 50- a :5 S (I) .9 o 25- o IIIIII II I I I I I I I I I I I I 2 4 6 810121416182022242628 Age (days) Figure 3.7. Frequency of diastatic fractures caused by both rigid and compliant interfaces versus age. A two-factor (age, interface) AN OVA performed on the total fiacture length (bone and diastatic combined) revealed significantly more damage caused by the compliant interface than by the rigid interface on pig skulls less than seventeen days of age, similar amounts of damage fiom both interfaces between seventeen and twenty-two days of age, and significantly more damage caused by the rigid interface than by the compliant interface for specimens aged twenty-four to twenty-eight days (p<0.05) (Figure 3.8). 41 150 - _ I Rigid i , El Hexcell E g 100 :E m r: c -' i E 13 50 - E u. . -l l l 2 4 6 810121416182022242628 Age (days) Figure 3.8. (Average) Fracture length generated by rigid and compliant impacts at the same impact energy. DISCUSSION In an effort to better understand fiacture of the developing skull under known conditions, the current study examined characteristics of impact induced fiacture to the infant porcine skull with respect to age and impact interface. As would be expected, the peak impact force increased with age. While an increase in the bending modulus of the porcine skull with age would account for this increase in force with age, a previous study shows that the bending modulus of infant porcine bone and bone-suture-bone specimens under 4-point bending does not change significantly with specimen age (Baumer et al., in press). Measurements of the average parietal bone thickness, however, showed a significant increase with age, which seemed to correlate with changes in the impact force. It is expected that the increasing stength of the skull with age is due to changes in geomety (thickness), and not due to an increase in bone quality with age. 42 There has been debate in the forensic community as to the mechanism of fracture initiation on adult human skulls. Specifically, whether fiactures initiate at the point of impact (Kroman, 2004) or at remote sites of “outbending” (Gurdjian et al., 1950). In the current study, 70 of 76 infant porcine specimens displayed fiacture initiation at a suture- bone boundary, and not at the center of the parietal bone (the impact location). The other six specimens involved comminuted skull fiactures that obscured the site of initiation. The locations of fracture initiation on infant porcine specimens in the current study agreed with the findings of Gurdjian et al. (1950) for adult human skulls, however it is currently unclear if the fracture initiation sites observed in this study are due to the immature nature of the cranial bone, differences in porcine cranial geomety or a combination these factors. Despite potential differences limiting direct comparisons with adult human fracture information, the infant porcine model developed in the current study showed that fiacture initiation occurred in relatively predictable locations at bone- suture interfaces for impacts on the parietal bone against a flat surface. This information may be expanded in the future to other interface conditions and should be correlated with fiacture patterns in the human infant skull from case files. These findings suggest that skull fractures on the developing skull from impacts with rigid or compliant flat surfaces may be located remote tom the site of the impacting force and nearer to the cranial sutures. The current study produced a greater amount of fiacture on younger infant porcine skulls using a compliant than rigid interface. The concentrated force of an impact with a rigid interface was expected to produce a larger amount of bone fiacture than an impact with a compliant interface where the same force intensity would be 43 distributed over a larger contact area and contact stesses would be reduced. This study, however, showed that the relative amount of fracture between these interfaces was dependent on age. For infant pigs less than seventeen days of age, a compliant interface caused significantly more damage than a rigid interface for the same impact energy, while for specimens twenty-four days of age and older the rigid interface generated more damage than the compliant interface. Specimens younger than seventeen days of age also showed fiequent damage to the sutures, while specimens older than seventeen days of age had bone fractures alone. In a previous study on beam specimens cut fi'om infant porcine skulls and loaded in 4-point bending, it has been shown that there is a significant difference in the relationship between bone and bone-suture-bone material properties for specimens aged fourteen and eighteen days (Baumer et al., in press). These changes resulted in a significant increase in the relative stength of the coronal suture at eighteen days to values more similar to that of bone. This effect correlates well with the disappearance of suture damage and a change in interface-dependent fiacture differences documented in the current study. The age-dependent response of the developing skull to impact with rigid and compliant surfaces is important in forensic cases for determining potential differences between testimony describing the events leading to injury and the observed fracture pattern. Specifically, the infant porcine model suggests that very young skulls may actually be more susceptible to extensive skull fiacturing tom an impact against a compliant surface than fiom an impact against a more rigid surface. Further investigations will be needed to create a more comprehensive set of patterns for various impact interfaces and energies. In an effort to explain the interface- and age-dependent results of the current study, it was noted that the contact area for the compliant interface impacts was generally greater than that for the rigid interface impacts. Yet, the generated impact force was not different between interfaces. It was hypothesized that the increased contact area for the compliant interface impacts resulted in a higher state of stess close to the boundaries of the parietal bone (sutures). It may then be suspected that the mechanical properties of the sutures were more critical during impacts with a compliant than with a rigid interface. For specimens under seventeen days of age, when the sutures are significantly more compliant than bone, large deformations at the sutures during impact for both interfaces may have generated high levels of stess near the boundaries of the parietal bone at the sutures. Therefore the risk of damage at the suture may have increased, resulting in the high frequency of diastatic fiactures. Furthermore, the higher state of peripheral stesses caused by the compliant interface may have promoted fracture propagation, resulting in the increased fracture length documented for the compliant interface versus the rigid interface. The increase in suture stength that occurs between fourteen and eighteen days of age may have allowed a better distibution of impact stesses across the boundaries of the parietal bone, resulting in lower overall levels of stess in the parietal bone and across suture lines. This effect may help explain why in the relatively older infants the compliant interface actually generated less fiacturing to the skull than for the rigid interface, since the rigid interface would be expected to generate more localized contact stesses nearer the site of impact. In summary, this study documented an increase in impact energy required to initiate skull fiactures as specimen age increased. Furthermore, there was relatively more 45 fracture damage generated for a compliant than a rigid interface with equal input energy at the younger infant ages. While a direct correlation with infant human fiacture patterns is not currently available fi'om case files, the experimental observations from the developing porcine model may aid in understanding some of the limited data that is available today. The experimental documentation of fiacture initiation energy and the extent of skull fracture with respect to interface as a fimction of specimen age may be of critical importance in fixture investigations of potential childhood abuse. 46 REFERENCES Baumer T, Passalacqua N, Powell B, Fenton T, Haut R., Accepted for Publication, “Age Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human,” Journal of Biomechanical Engineering, in press. Belechri M, Petridou E, Trichopoulos D., 2002, “Bunk versus conventional beds: a comparative assessment of fall injury risk,” J oumal of Epidemiology and Community Health, 56, pp. 413-7. Bertocci G, Pierce M, Deemer E, Aguel F, Janosky J, Vogeley E., 2003, “Using Test Dummy Experiments to Investigate Pediatic Injury Risk in Simulated Short-Distance Falls,” Archives of Pediatric Adolescent Medicine, 157(5), pp. 480-6. Billmire M. and Myers P., 1985, “Serious head injury in infants: accident or abuse?,” The American Academy of Pediatrics, 75, pp. 340-2. Case M, Graham M, Handy T, J entzen J, Monteleone J ., 2001, “Position Paper on Fatal Abusive Head Injuries in Infants and Young Children,” The American Journal of Forensic Medicine and Pathology, 22, pp. 112-22. Chalmers D, Marshall S, Langley J, Evans M, Brunton C, Kelly A, et al., 1996, “Height and surfacing as risk factors for injury in falls fi'om playground equipment: a case- contol study,” Injury Prevention, 2, pp. 98-104 Cooperman D. and Merten D., 2001, “Skeletal Manifestation of Child Abuse,” In: Child Abuse: Medical Diagnosis and Management. Philadelphia: Lippincott, Williams, and Wilkins, pp. 123-56. Dickerson J. and Dobbing J ., 1967, “Prenatal and postratal growth and development of the cental nervous system of the pig,” Proceedings of the Royal Society of London, 166(1005), pp. 384-95. Gruskin K and Schutzrnan S., 1999, “Head Trauma in Children Younger Than 2 Years,” Archives of Pediatric Adolescent Medicine, 153, pp. 1 5-21. Gurdjian E, Webster J, Lissner H., 1950, “The Mechanism of Skull Fracture,” Radiology, 54(3), pp. 313-39. Herring S. and Teng S., 2000, “Stain in the Braincase and Its Sutures During Function,” American Journal of Physical Anthropology, 112, pp. 575-93. Hobb C., 1984, “Skull Fracture and the Diagnosis of Abuse,” Archives of Disease in Childhood, 59, pp. 246-52. 47 Knight 3., 1991, “Fatal child abuse,” In: Forensic Pathology. London: Edward Arnold, pp. 457-73. Kroman A., 2004, “Experimental Study of Fracture Propagation in the Human Skull: A Re-Testing of Popular Theories,” Proceedings of the American Academy of Forensic Sciences, Dallas, TX, H79:314. Leventhal J, Thomas S, Rosenfield N, Markowitz R., 1993, “Fractures in young children: distinguishing child abuse from unintentional injuries,” American Journal of Diseases of Children, 147, pp. 87-92. Meservy C, Towbin R, McLaurin R, Myers P, Ball W., 1987, “Radiographic characteristics of skull fiactures resulting fi'om child abuse,” American Journal of Roentgenology, 149, pp. 173-5. Pierce M, Valdevit A, Anderson L, Inoue N, and Hauser D., 2000, “Biomechanical Evaluation of Dual-Energy X-Ray Absorptiomety for Predicting Fracture Loads of the Infant Femur for Injury Investigation: An In Vito Porcine Model,” Journal of Orthopaedic Trauma, 14(8), pp. 571-6. Prange M, Luck J, Dibb A, Van Ee C, Nightingale R, Myers 3., 2004, “Mechanical Properties and Anthropomety of the Human Infant Head,” Stapp Car Crash Journal, 48, pp. 279-99. Prange M. and Margulies S., 2002, “Regional, Directional, and Age-Dependant Properties of the Brain Undergoing Large Deformations,” Journal of Biomechanical Engineering, 124(2), pp. 244-52. Prange M, Meaney D, Margulies S., 2000, “Defining brain mechanical properties: Effects of region, direction, and species,” Stapp Car Crash Journal, 44, pp. 205-13. Reece R and Sege R., 2000, “Childhood Head Injuries: Accidental or Inflicted,” Archives of Pediatic Adolescent Medicine, 154, pp. 11-15. Reiber G., 1993, “Fatal falls in childhood: How Far Must Children Fall to Sustain Fatal Head Injury?,” American Journal of Forensic Medicine and Pathology, 14, p. 201-7. Schneider L, Lehman R, Pflug M, Owings C., 1986, “Size and Shape of the Head and Neck From Birth to Four Years,” Washington, DC. The Consumer Product Safety Commission, Report No.: UMTRI-86-2. Thibault K. and Margulies S., 1998, “Age-Dependant Material Properties of the Porcine Cerebrum: Effect on Pediatic Head Injury Criteria,” Journal of Biomechanics, 31(12), pp. 1 1 19-26. 48 Wheeler D. and Shope T., 1997, “Depressed skull fracture in a 7-month old who fell from bed,” Pediatrics, 100, pp. 1033-4. 49 CHAPTER FOUR CHARACTERISTICS AND PREDICTION OF CRANIAL CRUSH INJURIES IN CHILDREN ABSTRACT Experimentally applied bilateral crushing forces on adult cadaver skulls and even a few case reports on children have shown similarities in basilar fiacture locations. It has even been suggested that the overall path of tacture may be related to the location of applied force on the skull. This study documents 4 clinical cases of fatal crush injuries to young children between 1.5 and 6 years of age with correlations between modeled stess and clinically observed fracture patterns. The damage in the clinical cases was concentated in the basicranium. Each of the cases presented fracture patterns that bridged the impact sites and traversed the middle cranial fossa in the area of the spheno-occipital synchondrosis. The crushing forces observed in the clinical cases were recreated on a simplified finite element model of a cranium by applying bilateral pressures to regions correlating to the specific exterior injury locations of each case. Numerous trials were run within the elastic range of bone to develop a representative and clearly characterized pattern of principal stess directions. Areas of tensile stess were assumed to be the most likely location(s) of potential fracture. The regions of largest tensile stess developed on the skull were compared to the clinically observed fracture patterns in the 4 cases, revealing that locations of high tensile stess in all four cases were located on the basicranium with low levels of stess on the cranial vault. Additionally, regions of high tensile stess appeared to follow the fiacture path(s) extending tom the location of maximum stess. These results suggest that pre-failure stess field diagrams may give 50 some insight into the projected fracture propagation paths, although prediction will not be exact. Also, these analyses of the modeled stesses and the clinically observed fractures suggest that quasi-static bilateral loading of the cranium may lead to predictable fracture of the basicranium, possibly independent of age. 51 INTRODUCTION Fracture interpretation is a necessary component of forensic analysis and contributes to the determination of the cause and manner of death. Skeletonized remains, in particular, lack skin and scalp abnormalities that enhance the complete description of a head injury or injuries. Thus, a clear understanding of fracture propagation becomes vital to the forensic professional. Head injuries are often the result of dynamic forces in which a moving object strikes the cranium at a considerable speed or the cranium makes contact with a static object that causes it to rapidly decelerate (Duhaime et al., 1995). In cases of crushing injuries, however, a static load is applied relatively slowly to the cranium, initially causing deformation with eventual fiacturing if pressure continues or increases. In experimental tests on the effects of bilateral crushing forces on cadaver skulls, Russell and Schiller (1949) report that avulsion of the petous portion, fracture through the dorsum sellae, and principal fracture lines running in the direction of the compressive forces were common features observed. There are several case reports of basilar skull damage in the literature involving both crushing and impulse based loading scenarios mainly focusing on adult crania. A study of 15 cases of crushing injuries to adult skulls aged 14-63 years and experimentally induced crushing fractures on 11 cadaver crania revealed similarities in basilar fracture patterns (Russell and Schiller, 1949). However, a study by Harvey and Jones (1979) claimed that fractures through both petous bones was not an indicator of impact site. Another study by Yasue (1981) then suggested that the overall path of fiacture through the basicranium could, in fact, be related to the region of the skull impacted. More recently, a study by Kroman (2005) attempted to produce cranial base 52 (hinge) fractures in two unembalmed cadaver adult human crania specimens using a drop tower. Hi gh-speed video dernonstated that fi'actures initiated at the site of impact and propagated through the middle cranial fossa, with some additional minor fractures surrounding the forarnen magnum. In this study the vaults had been removed so it is unclear if the fiactures would have terminated at the points of applied forces (fixture and impactor). Kroman further notes a very wide separation of the failing bones during impact which no doubt would cause very severe soft-tissue damage in the cranial base; this correlates to the high amount of energy and lethality involved in these scenarios. It is assumed that fractures will generally follow the path of least resistance, but this path may be different between the child and adult cranium. While all of the aforementioned studies focused on adult crania, some recent case reports on quasi-static loading to the crania of children have also presented observations of basilar fiactures (T akeshi et al., 2005; Duhaime et al., 1995). The tend of quasi-static loading producing basilar tactures in both adults and children may suggest that prediction of these fractures is possible. The Finite Element (FE) method is a powerful analytical modeling tool that can be used to determine internal stesses and stains within a complex solid. Several studies have shown that maximum principal stess (or stain) directions from a modeled stucture correlate well with areas of experimentally or clinically observed bone fiactures (Graham et al., 2000; Silva et al., 1998; Doorly and Gilchrist, 2006). While modeling of fiacture propagation generally involves high computational costs, one theory suggests that there is a close, although not exact, relationship between pre-failure stess tajectories and the ultimate path of propagation. In brief, Frank and Lawn (1967) 53 proposed that crack growth occurs in such a way as to maximize the energy dissipation per unit length of growth, and that the greatest amount of dissipation occurs when the crack plane is perpendicular to the direction of greatest principal tensile stess. In other words, the directions of principal tensile stess developed on a solid may be a reasonable predictor of propagation direction. In this study, four cases of childhood fatalities from crushing head injuries inflicted by vehicle tires exhibited similar fracture patterns despite differences in age and extent of force. This raised the question as to whether certain fracture patterns are indicative of bilateral crushing injuries and if these could be predicted based on the type of loading and the biomechanical properties of the skull. MATERIALS AND METHODS Four clinical cases of fatal crush injuries to young children between 1.5 and 6 years of age under known conditions were used as a baseline for applied forces and resulting fiacture patterns. In all cases the skull was tapped between a vehicle tire and the ground. The contact sites and skull fractures were identified and diagramed during autopsy by the investigating forensic pathologist (M.N.). Biomechanical modeling was then conducted in an attempt to recreate and explain the similarities exhibited in the fracture patterns of these cases. A simplified model of a skull was created in a finite element package (Abaqus v.6.3, Hibbitt, Karlsson & Sorensen, Inc., Pawtucket, RI, USA) (Figure 4.1). This model consisted of synmrety about the sagittal plane, but not about the coronal or tansverse planes. Each major bone of the vault was modeled separately, assumed to have a uniform thickness, and constained at the suture joints to form the fill] cranial model. 54 The base of the cranium was modeled as a flat surface with the same thickness as the vault. The primary landmarks in the basal region were modeled: a hole was added to simulate the foramen magnum, dense oblong extusions lateral to the foramen magnum were generated to represent the petous portions, and anterior to the foramen magnum the thickness was reduced to represent the region of the spheno-occipital synchondrosis. Foramen Petous / C\ magnum portions *9 Posterior N m Spheno-occipital synchondrosis Figure 4.1. Simplified cranial model showing modeled landmarks of the basicranium. View of endocranial surface with vault removed. FE analysis was utilized in an effort to reproduce and examine possible relationships between the theoretical pre-failure stesses developed on the skull and the fracture patterns observed clinically. The magnitude of the force applied to the cranial model was the same in every case and was scaled down such that the modeled cranium would not actually fail. The crushing forces reported in the clinical cases were recreated by applying the force over a region of the cranium correlating to the specific exterior injury locations for each case. Numerous trials were run within the elastic range of bone to develop a representative and clearly characterized pattern of principal stess and stain 55 directions. Areas of tensile (principal) stess were assumed to be the most likely location(s) of potential tacture. RESULTS CASE REPORTS Case I A 5-year-old female child was stuck by her family’s sedan when she fell in an attempt to stop the vehicle after her brother accidentally engaged the tansrrrission. Her head was run over by the right front tire of the car. The autopsy revealed an intact cranial vault with a cranial base fracture that taversed the lateral right anterior cranial fossa, the midline of the middle cranial fossa, and the left lateral aspect of the posterior cranial fossa. This fracture bridged the blunt impact sites on the right anterior and left posterior aspects of the head. Case 2 An unhelmeted 6-year-old male was stuck by a car while riding his bike. His head was run over by the vehicle’s left rear tire. Autopsy revealed an intact cranial vault. The cranial base, however, was bisected by a gaping X-shaped tacture centered in the midline of the middle cranial fossa. There were two adjacent arms of the fiacture in the right middle cranial fossa. The other two adjacent arms were in the posterior aspect of the left middle cranial fossa and the anterior aspect of the left posterior cranial fossa. This fiacture bridged the blunt impact sites on the right anterior and lateral aspects of the head and the left lateral and posterolateral aspects of the head. Case 3 A 3-year-old male was run over attempting to run away from a farm grain wagon that had begun to roll. His head was run over by one of the wheels of the wagon. Autopsy revealed gaping linear fractures that taversed the parietal bones and the right 56 and left sides of the fiontal bone. One end of each fracture intersected the widely gaping right side of the coronal suture, other ends were continuous with fractures of the cranial base. A dominant cranial base fracture taversed the left petous bone, the midline of the middle cranial fossa, the anterior aspect of the right middle cranial fossa and the posterolateral aspect of the right anterior cranial fossa. Additional intersecting fractures were centered in the left middle and posterior cranial fossae. Case 4 A 1.5-year-old female was entapped under a small moving van and her head was run over by the right rear wheel. Autopsy revealed an intact cranial vault but extensive intersecting fractures of the cranial base that predominately taversed the middle and posterior cranial fossae fi'om side to side. A large fracture also bisected the anterior cranial fossae. Fracture Pattern Similarities Each of the cases presented tacture patterns that bridge the impact sites and taversed the middle cranial fossa in the area of the spheno-occipital synchondrosis. Additionally, the damage was concentated in the basicranium leaving the cranial vault intact in all cases except the 3-year-old boy impacted by the farm grain wagon. Furthermore, general amounts of fracture increased with inferred increase in magnitude of applied force (grain-wagon vs. car) or decrease in age. The four cases presented three contact scenarios with four different fracture patterns. Specifically, the contact forces in Case 1 were applied primarily to the frontal and occipi tal bones in a direction between anterior-posterior and lateral (Figure 4.2a), the contact forces in Cases 2 and 3 were also applied to the fiontal and occipital bones 57 but had an additional lateral component (Figure 4.2b, c), and the contact forces in Case 4 were applied to the lateral surfaces of the cranial model (Figure 4.2d). Figure 4.2. Diagrams of fractures due to quasi-static crushing forces to skulls aged (a) 5 years — Case 1, (b) 6 years — Case 2, (c) 3 years — Case 3, and (d) 1.5 years — Case 4. White cranial margins represent contact of applied forces (ground and tire). FE Analysis Modeling of all four cases revealed that locations of high tensile stess and high tensile stain were located on the basicranium. There were some areas of lower tensile stess at the sites of the applied load and in a few areas on the cranial vault, but due to the assumption that failure would occur at locations of high stess, these areas were ignored for this analysis. The maximum tensile stess for all models were similar (standard deviation of :I:5%). The tensile stesses developed on the skull were filtered to display only the highest 20% for correlation to fracture patterns. The leng th of the arrow at each location represented the relative magnitude of stess at that location. Case 1 presented the simplest fiacture pattern as a single linear fracture taversing the basicranium from the right greater wing of the sphenoid across the spheno-occipital synchondrosis to the left lateral portion of the occipital, connecting the loading sites (Figure 4.3a). An overlay of the directions of maximum principal tensile stesses revealed a linear concentation on the right side of sphenoid and in the area of 58 the spheno-occipital synchondrosis, which correlated well with sections of the observed fracture (Figure 4.3b). Figure 4.3. (Case 1) Crushing injury to the skull of a 5-year-old. Diagrams of (a) clinically observed fi’actures and (b) an overlay of the highest 20% of principal tensile stesses produced from a FE model. White cranial margins represent contact of applied forces (ground and tire). Cases 2 and 3 presented slightly more complex fracture patterns. In part, the complexity was increased due to the similarities in loading but dissimilarities in fracture (Figure 4.4a and Figure 4.5a, respectively). A similar fiacture in both cases stetched from the right aspect of the sphenoid through the sella turcica and along the superior aspect of the petous portion of the temporal. The overlay of the directions of maximum principal tensile stesses was identical in both cases due to the similarity of loading patterns. There was correlation between the stess and the fiacture for both cases, especially along the line of similar fiacture, although each had areas of high stess where fracture did not occur (Figure 4.4b and Figure 4.5b). 59 Figure 4.4. (Case 2) Crushing injury to the skull of a 6-year-old. Diagrams of (a) clinically observed fractures and (b) an overlay of the highest 20% of principal tensile stesses produced from a FE model. White cranial margins represent contact of applied forces (ground and tire). Figure 4.5. (Case 3) Crushing injury to the skull of a 3-year-old. Diagrams of (a) clinically observed fractures and (b) an overlay of the highest 20% of principal tensile stesses produced from a FE model. White cranial margins represent contact of applied forces (ground and tire). Case 4 presented the most complex fracture pattern extending both longitudinally and tansversely through the basicranium (Figure 4.6a). Fractures ran longitudinally across the midline of the anterior cranial fossa, and horizontally through the middle and posterior cranial fossa with various areas of comminution. The overlay of the directions of maximum principal tensile stesses revealed an area of complex stess in the same area as the highly comminuted fracture of the right side of the cranial base. Additionally, 60 a more linear pattern of stesses occurred near the fiactures on the left side of the skull (Figure 4.6b). Figure 4.6. (Case 4) Crushing injury to the skull of a 1.5 year old. Diagrams of (a) clinically observed fiactures and (b) an overlay of the highest 20% of principal tensile stesses produced from a FE model. White cranial margins represent contact of applied forces (ground and tire). DISCUSSION The analyses performed in this study represent a quantitative, theoretical attempt to correlate stess and stain intensities and their distributions to fracture patterns on the skull caused by quasi-static bilateral pressures. A linear, small stain, isotopic, uniform thickness model of a human calvarium was constructed to maintain some manageability in the preliminary modeling. Analyses of the modeled stesses and the clinically observed fractures suggested that quasi-static bilateral loading of the cranium may lead to predictable fracture of the basicranium, although the predictions were not exact. The promising nature of these results suggest that it may be useful to develop a more complex model for a more accurate prediction of fractures on a case-by-case basis. The location of maximum tensile stess would be expected to be the site of fracture initiation, and the location of maximum stess in each of the four models correlated well with the area(s) of fiacture in each of the clinical cases. However, Frank 61 and Lawn (1967) also suggest that paths of high tensile stess may approximate the path of fi‘acture propagation. In the four cases presented in this study, regions of high tensile stess appeared to follow the path of the fracture extending from the location of maximum stess. Fracture propagation, however, requires increasing levels of stess to overcome the dissipation of energy that occurs with the previous fiacturing. These increasing levels of stess may ultimately result in additional fiacture sites on the material surface (Paterson and Wong, 2005). Two or more isolated regions of the skull developed high tensile stesses in Cases 2-4, which may have resulted in multiple sites of fracture initiation with propagation extending from each site. Multiple initiation sites with subsequent propagation joining the sites could explain the complexity of the fiactures seen in Cases 2-4, as opposed to the singular region of fiacture and stess intensity developed in Case 1. The four cases presented in this study suggest that pre- failure stess field diagrams may, in fact, give some insight into the projected fiacture propagation paths and, ultimately, fracture patterns developed during quasi-static bilateral loading of the cranium. The cases and analyses presented in this study supported some general tends for quasi-static bilateral loading scenarios. Autopsy revealed tactures through the middle cranial fossa in the region of the spheno-occipital synchondrosis in all cases. Modeling displayed high tensile stesses through this region in Cases 1-3; with Cases 2 and 3 developing maximum levels of stess at this landmark. However, modeling of Case 4 showed no region of high tensile stess passing through the spheno-occipital synchondrosis. Interestingly, Case 4 involved the youngest child and one of the most massive vehicles (moving van). This result may suggest that the material model should 62 be altered as a fimction of age, especially for the very young pediatic patient. Autopsies also revealed fiactures connecting the loading sites through the basicranium in all four cases. The models showed regions of high tensile stess in this region for Cases 1-3, although the tensile stesses in Case 4 did not seem to follow the same pattern. Intact cranial vaults were also present in three of four cases. The models displayed no significant levels of stess on the vault that would be expected to cause fiacture. It is possible that the mass of the vehicle (grain wagon) caused excessive fracture propagation that was not able to be modeled in Case 3. All of the fiacture related observations seen in the current study showed similarities to experimentally developed fiactures on adult cadaver skulls (Russell and Schiller, 1949) suggesting that these patterns of quasi-static loading of the cranium may be somewhat age independent, except for the very young patient; likely as a function of cranial geomety which would remain essentially the same throughout life. In summary, a simple model of a cranium subjected to clinically observed bilateral crushing forces was largely able to predict the characteristics of the developed fi'acture patterns. The results of this study were not meant to propose that the theoretical stesses developed on this simplified cranial model actually predict an arbitary fracture pattern developed clinically. These results do show that a simplified analysis of tensile stesses developed tom some basic information on boundary loads and locations on the cranium can generate patterns of fracture that are consistent with clinical fractures. The four cases in conjunction with the FE model suggest that craniums subjected to bilateral crushing forces tend to fracture in predictable ways. Further, the regions of high tensile 63 stess that develop between the locations of loading on the pediatric cranium coincide with the locations of tensile fracture. The predictability of this pediatric cranial fi'acture pattern has practical application. Medical examiner and coroner death investigations attempt to establish circumstances of injury. In cases of unknown or questionable circumstances, this fiacture pattern would suggest a particular mechanism of injury (relatively slow application of a static load) consistent with a heavy wheeled vehicle. The described fiacture pattern would be less consistent with other, more common injuries such as those sustained in motor vehicle collisions (ejection or internal tumbling) and in the setting of inflicted blunt tauma. This information may prove relevant in clarifying unexplained perimortem events in cases of crushing injuries to the pediatric cranium, particularly if information from examination of skin and scalp is not possible due to skeletonization. 64 REFERENCES Doorly M, Gilchrist M., 2006, “The use of accident reconstuction for the analysis of taurnatic brain injury due to head impacts arising from falls,” Computer Methods in Biomechanics and Biomedical Engineering, 9(6), pp. 371-7. Duhaime A-C, Eppley M, Margulies S, Heher K, Bartlett S., 1995, “Crush Injuries to the Head in Children,” Neurosurgery, 37(3), pp. 401 -7. Frank F, Lawn B., 1967, “On the Theory of Hertzian Fracture,” Proceedings of the Royal Society of London, 299(1458), pp. 291-306. Graham R, Oberlander E, Stewart J, Griffiths D., 2000, “Validation and use of a finite element model of C-2 for deterrrrination of stess and fiacture patterns of anterior odontoid loads,” Journal of Neurosurgery (Spine 1), 93, pp. 117-25. Harvey F, Jones A., 1980, “‘Typical’ Basal Skull Fracture of Both Petous Bones: An Unreliable Indicator of Head Impact Site,” Journal of Forensic Sciences, 25(2), pp. 280- 6. Paterson M, Wong T., 2005, “Griffith Theory of Brittle Failure,” In: Experimental Rock Deformation — The Brittle Field. Berlin: Springer, pp. 45-57. Russell W, Schiller F ., 1949, “Crushing Injuries to the Skull: Clinical and Experimental Observations,” Journal of Neurology, Neurosurgery, and Psychiaty, 12, pp. 52-60. Silva M, Keaveny T, Hayes W., 1998, “Computed Tomography-Based Finite Element Analysis Predicts Failure Loads and Fracture Patterns for Vertebral Sections,” Journal of Orthopaedic Research, 16, pp. 300-8. Takeshi M, Okuchi K, Nishiguchi T, Seki T, Watanahe T, Ito S, et al., 2006, “Clinical Analysis of Seven Patients of Crushing Head Injury,” The Journal of Trauma-Injury, Infection, and Critical Care, 60(6), pp. 1245-9. Tortosa J, Martinez-Lage J, Poza M., 2004, “Bitemporal Head Crush Injuries: Clinical and Radiological Features of a Distinctive Type of Head Injury,” Journal of Neurosurgery, 100, pp. 645-51. Yasue M., 1981, “Clinical and experimental investigation of mediobasal skull fiacture,” Neurol Med Chir (Tokyo), 21, pp. 1041-9. 65 CHAPTER FIVE FINITE ELEMENT MODELING OF RIGID AND COMPLIANT SURFACE IMPACTS TO THE PEDIATRIC PORCINE PARIETAL BONE ABSTRACT Head injuries account for 80% of fatal child abuse in younger children. Yet, in many cases the ability to determine child abuse from accidental injury may be difficult. Head injuries may occur when a moving object stikes the cranium at a considerable speed or when the cranium strikes a static object that causes the skull to rapidly decelerate. This study documents theoretical principal stess and stain directions developed on a cranial model with respect to age and interface in an effort to identify fiacture characteristics for the developing skull under known conditions. The impact conditions generated experimentally for rigid and compliant interface impacts to porcine skulls of various ages were recreated on a simplified finite element model. The model accounted for changes in bone thickness as well as material properties of bone and suture with age. Numerous trials were run within the elastic range of the materials to develop a representative and clearly characterized pattern of pre-failure principal stess and stain directions. Maximum principal stess and stain directions revealed distinct locations of expected fracture regardless of age or interface. However, the extent and magnitude of stains at any given location were dependent on both age and interface. Experimental results showed that blunt force tauma to the parietal bone consistently resulted in fiacture at the bone-suture boundaries of the parietal bone, which correlated well with areas of high stain. Young specimens showed high levels of stain in the sutures while old specimens showed limited frequency and magnitude of stain within the sutures. 66 These findings correlated well with previous findings of significantly higher rates of suture failure in young specimens than in old specimens. However, suture damage only occurred experimentally in 29% of younger specimens, while bone tacture occurred in 97%. Due to the relative differences in the elastic moduli of bone and suture it was expected that similar levels of stain could cause failure of the bone prior to failure of the sutures. Interestingly, the larger contact area generated during compliant interface impacts also seemed to produce stains within the bone distributed nearer the parietal bone boundaries. This distribution of high stains may have promoted propagation of fractures initiating at the boundaries, resulting in the greater amount of observed damage on the cranium produced from compliant interface impacts. 67 INTRODUCTION Skull fracture interpretation is a necessary component of forensic analysis and contributes to the determination of the manner and cause of death. Head injuries are often the result of dynamic forces in which a moving object strikes the cranium at a considerable speed or the cranium makes contact with a static object that causes the skull to rapidly decelerate (Duhaime et al., 1995). Head injuries account for 80% of fatal child abuse in younger children (Case et al., 2001). Yet similar types of injuries can be generated by abusive and accidental tauma (Billmire and Myers, 1985) and, due to the paucity of cranial fracture tolerance data for infants and young children, pediatric tauma involving single-event head injuries with related cranial fi'actures represent one of the greatest challenges to forensic pathologists and anthropologists. Thus, a clear understanding of fracture patterns becomes vital to the forensic professional. While several studies have documented important information regarding the response of the infant cranium and brain to blunt force impact, these data are limited to one or two age groups. And because of ethical considerations, experimentation on human pediatric specimens is very limited. Scaling of the adult cranium has met with some success in predicting impact response of the pediatic cranium (Prange et al., 2004) but, the head of an infant is smaller and geometrically unlike that of an adult (Schneider et al., 1986) and the validity of predicting cranial fiacture patterns in infants from adult data has not been investigated. Using adult data to predict fracture patterns in the pediatric cranium may also be problematic due to the different stuctural (ease of deformation and decreased threshold to fiacture) and mechanical properties of the infant cranium (Thibault and Margulies, 1998; Prange et al., 2000; Prange and Margulies, 68 2002). However, one animal model of the human infant has begun to emerge regularly in the current literature. A study by Dickerson and Dobbing (1967) shows a similarity in development of the cental nervous system when correlating months of the human to weeks of the pig. Other studies have used porcine models to predict fracture loads for the infant human femur (Pierce et al., 2000) and strain in the braincase and sutures (Herring and Teng, 2000). More recently, age-related rates of change in the bending rigidity of the infant human and infant porcine craniums has been found to be similar for a correlation based on months in humans to days in pigs (Baumer et al., in press). The Finite Element (FE) method is a powerful analytical modeling tool that can be used to determine internal stesses and stains within a complex solid. Several studies have shown that maximum principal stess (or stain) directions fi'orn a modeled stucture correlate well with areas of experimentally or clinically observed bone fiactures (Graham et al., 2000; Silva et al., 1998; Doorly and Gilchrist, 2006). While modeling of fracture propagation generally involves high computational costs, one theory suggests that there is a close, although not exact, relationship between pre—failure stess/stain tajectories and the ultimate path of propagation. In brief, Frank and Lawn (1967) propose that crack growth occurs in such a way as to maximize the energy dissipation per unit length of growth, and that the greatest amount of dissipation occurs when the crack plane is perpendicular to the direction of greatest principal tensile stess. In other words, the directions of principal tensile stess developed on a solid may be a reasonable predictor of the propagation path. In this study, blunt impact tauma to the center of the parietal bone was modeled for comparison to fiacture patterns experimentally generated on the developing porcine 69 cranium. The hypothesis was that age and interface characteristics of a given impact would define the stess and stain fields developed on the cranium, leading to predictable fracture patterns. A secondary goal was to examine possible stess or stain related causes for the experimentally generated fractures that consistently occurred away from the impact site at the bone-suture boundaries (Chapter 3). MATERIALS AND METHODS A simplified model of a porcine cranium was created in a finite element package (Abaqus v.6.3, Hibbitt, Karlsson & Sorensen, Inc., Pawtucket, RI, USA) using dimensions from CT scans of impacted crania Most elements of the cranial geomety were accounted for, although some regions were simplified due to the distance tom the area of interest and complexity of meshing (Figure 5.1). The curvature of the parietal bone was determined from CT scans and a 5-point spline was used to simulate the measured curvature. The average bone thickness of the parietal bone was taken from specimens tested in 4-point bending and was altered with age, but the modeled bone was assumed to have a uniform thickness. The sutures were given a more pronounced width than observed on the actual cranium since the material properties used in the model were not directly obtained from sutures, but involved the bone on both sides of the sutures as well. The primary landmarks in the basal region were included, but the specific undulations were removed and replaced by a flat plate with the same thickness as the rest of the cranium. A region of the frontal bones was modeled, but the snout and orbits were removed. The model consisted of symmety about the sagittal plane, but not about the coronal or tansverse planes. Each bone and suture of the cranium was modeled separately and constained at the bone-suture interfaces to form the full cranial model. 70 Contact area (varied) Parietal Bone # FE Parietal Sutures Frontal «7",— B (smooth line) ~ one Occipital Bone Foramen Magnum Base Orbit* CT Parietal Figure 5.1. Simplified cranial model with overlays of modeled stuctures on true stuctures. Coronal cut of CT used for thickness comparison. *Feature not modeled. TFeature simplified. #Suture width altered for consistency with material property data obtained previously. Dotted line represents location of frontal bone removal. Finite Element (FE) analysis was utilized in an effort to produce and examine the expected stesses and stains developed on the cranium and to compare these stesses and stains to experimentally observed fracture patterns with respect to both age and interface. The thickness of the cranium and the material properties of the bone and suture were altered in the model to represent changes in age. Age related changes in the elastic moduli of bone and suture were taken from experimental results of a previous study (Baumer et al., in press), and average impact force and contact area data from another study were used to simulate the difference between a rigid and a compliant interface 71 (Chapter 3) (see Table 1). Briefly, the rigid interface was constained to produce smaller contact areas than the compliant interface with both interfaces producing the same magnitude of impact force. Numerous trials were run within the elastic range of bone to develop a representative and clearly characterized pattern of principal stess‘and stain directions. Table 5.1. (Average) properties used in development of the FE model and impact conditions. Elastic Modulus Impact Force Contact Area Specimen Age Esut Ebene Rigid Compliant Rigid Compliant Young (3-7 days) 3.38 7.21 561 508 94 119 Old (18-21 days) 3.37 5.29 930 860 176 322 The magnitude of principal stress and stain directions were filtered to remove the lowest 20%. The remaining stesses and stains were compared to the experimentally produced fracture patterns. A representative cranium from young and old specimens impacted by each interface was selected. Each chosen cranium displayed all the common fracture sites from the age/interface and no instances of uncommon sites. A photo editing program (Photoshop ..) was used to overlay the modeled stesses and stains onto a photograph of the chosen, experimentally fiactured cranium. RESULTS An overlay of maximum principal tensile stess directions revealed a concentation at the impact site for all age-interface combinations (Figure 5.2). The length of the arrow at each location represents the relative magnitude of stess at that location. Comparison of these overlays revealed distinct lines of expected fracture between suture boundaries and the impact point, regardless of age or interface. However, the proximity of the stess field to the boundary varied by interface. Specifically, larger 72 regions of high tensile stesses were located nearer the suture-bone boundary for compliant interface impacts than for rigid interface impacts. Furthermore, tensile stesses developed in the sutures of the young specimens were significantly greater for compliant interface impacts than for rigid interface impacts. The old specimens had similar levels of stess in the sutures for both interfaces. Figure 5.2. Overlay of maximum principal tensile stesses from FE models on representative experimental cranial fractures for (a) rigid interface impact to a young specimen, (b) rigid interface impact to an old specimen, (c) compliant interface impact to a young specimen, and (d) compliant interface impact to an old specimen. Longer lines represent higher stesses. An overlay of the directions of maximum principal tensile stain revealed high stains at or near bone-suture boundaries, with lower levels of stain at the impact site for all age-interface combinations (Figure 5.3). The rigid interface impacts produced higher levels of stain closer to the impact site than the compliant interface impacts, however, the highest strains were still located away from the impact site. Furthermore, the 73 magnitude of the tensile strains developed in the sutures of the young specimens from compliant interface impacts was considerably greater than those from rigid interface impacts. The old specimens had similar levels of stain in the sutures for both interfaces. Comparison of these overlays revealed some possible paths of expected fracture between bone-suture boundaries and the impact point. Figure 5.3. Overlay of maximum principal tensile stains from FE models on representative experimental cranial fractures for (a) rigid interface impact to a young specimen, (b) rigid interface impact to an old specimen, (c) compliant interface impact to a young specimen, and (d) compliant interface impact to an old specimen. Longer lines represent higher stains. DISCUSSION Modeling of a cranium with sutures resulted in a heterogeneous structure and, therefore, regions of high stess did not necessarily correspond to regions of high stain. Specifically, high stesses were generally located near the point of impact in the center of the parietal bone, while high stains were located at or near bone-suture boundaries. 74 Experimental results showed that blunt force tauma to the parietal bone consistently resulted in fracture at the bone-suture boundaries of the parietal bone (Chapter 3), which correlated well with areas of high stain. The correlation of fracture to stain has been previously documented in a study on failure of vertebral sections by Silva et al., (1998). The study suggests that regions of high stain may be a better predictor of failure than regions of high stess for heterogeneous stuctures. However, the current study utilizes two distinct materials resulting in different failure criteria for each. It has been shown that stains generated in the suture of the developing porcine skull during normal frmction can be 6 times greater than the stain developed in the bone (Herring and Teng, 2007) and in 4-point bending failure tests the stain in the suture can be 2 to 3 times greater than the stain in bone (Baumer et al., in press). Thus, it would be expected that initial material failure would occur in bone prior to suture for conditions of similar stain. Therefore, the stains in the bone near the boundaries of the parietal bone, which were slightly lower than the stains developed in the sutures, likely resulted in failure prior to failure of the suture stucture in most cases. Age effects played an important role in the distibution of stains on the cranium. The young specimens showed more frequent and greater magnitudes of stain in the sutures than the old specimens, which correlated well with the experimentally observed patterns of suture failure (Chapter 3). Specifically, the developing skull fi'om birth to sixteen days of age showed significantly higher rates of suture failure than specimens over sixteen days of age. The old specimens showed much higher fi'equency and magnitude of stain within the bone than the young specimens with some lower magnitude stains in the sutures. This shifting of stain from the sutures to the bone with 75 increasing age was likely related to the changes in suture properties observed between fourteen and eighteen days of age (Baumer et al., in press). Overall, the modeled patterns of stain seemed to support the general characteristics of experimentally generated fiacture with regard to age. The compliant interface produced high stains within and near the sutures while the rigid interface produced more localized stains around the impact site with lower magnitudes of stain in the sutures. The larger contact area during compliant interface impacts seemed to result in strains distributed nearer the parietal bone boundaries. This difference in the proximity of stain in the bone to the bone-suture boundaries may help explain the greater extent of fracture in compliant interface impacts at young ages. Specifically, higher levels of stain near the boundaries may have promoted propagation of fractures initiating at the boundaries, resulting in greater damage to the skull from compliant interface impacts. Despite the difference in proximity to the parietal bone boundaries, both interfaces produced higher levels of stain away from the point of impact than at the point of impact. At young ages, both interfaces produced high stains that were largely constained to the coronal and temporal sutures. However, suture damage was only seen in 29% of specimens aged under 16 days, while bone fracture occurred in 97% of these specimens (Chapter 3). This suggests that the sutures were better able to absorb the high stains while the slightly lower levels of stain observed in the parietal bone resulted in material failure. Of course, some suture damage was generated experimentally, suggesting that the sutures were not always able to fully absorb the high stains. It is currently unclear if the sutures in these cases failed before of after the bone, but it may 76 be reasonable to assume, based on the modeled stains, that the suture damage occurred after bone failure. The large stains developed in the suture model with relatively low fiequency of experimental damage may suggest that prediction of suture failure on a case-by-case basis may be difficult. However, the correlation between a reduction in magnitude of stain and the disappearance of damage in the sutures for older skulls suggests that some element of predictability may exist. In summary, a simple model of a porcine cranium subjected to the effects of experimentally induced blunt force tauma to the parietal bone was largely able to predict characteristics of observed fracture patterns. The maximum principal stain directions generated fnom the model seemed to correlate well with locations of fiactures seen in a previous study (Chapter 3). The results of this study were not meant to propose that the theoretical stesses or stains developed on this simplified cranial model actually predict an arbitary fracture pattern developed experimentally. The results do show that a simplified analysis of tensile stains and, to some extent, stesses developed fi'om some basic information on material properties of the cranium and interface specific loading patterns can generate patterns that are consistent with some aspects of experimental fractures. The experimentally obtained fiactures in conjunction with the FE model suggest that blunt impact tauma to the parietal bone of the developing cranium leads to predictable fiacture locations. This information may prove relevant in determining perimortem events in forensic cases of head injuries to the pediatric patient with unknown or questionable circumstances. 77 REFERENCES Baumer T, Passalacqua N, Powell B, Fenton T, Haut R., Accepted for Publication, “Age Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human,” Journal of Biomechanical Engineering, in press. Billmire M. and Myers P., 1985, “Serious head injury in infants: accident or abuse?,” The American Academy of Pediatrics, 75, pp. 340-2. Case M, Graham M, Handy T, Jentzen J, Monteleone J ., 2001, “Position Paper on Fatal Abusive Head Injuries in Infants and Young Children,” The American Journal of Forensic Medicine and Pathology, 22, pp. 112-22. Dickerson J. and Dobbing J ., 1967, “Prenatal and postnatal growth and development of the cental nervous system of the pig,” Proceedings of the Royal Society of London, 166(1005), pp. 384-95. Doorly M, Gilchrist M., 2006, “The use of accident reconstuction for the analysis of taumatic brain injury due to head impacts arising from falls,” Computer Methods in Biomechanics and Biomedical Engineering, 9(6), pp. 371-7. Duhaime A-C, Eppley M, Margulies S, Heher K, Bartlett S., 1995, “Crush Injuries to the Head in Children,” Neurosurgery, 37(3), pp. 401-7. Frank F, Lawn B., 1967, “On the Theory of Hertzian Fracture,” Proceedings of the Royal Society of London, 299(1458), pp. 291-306. Graham R, Oberlander E, Stewart J, Griffiths D., 2000, “Validation and use of a finite element model of C-2 for determination of stess and fiacture patterns of anterior odontoid loads,” Journal of Neurosurgery (Spine l), 93, pp. 117-25. Herring S. and Teng S., 2000, “Stain in the Braincase and Its Sutures During Function,” American Journal of Physical Anthropology, 112, pp. $75-93. Pierce M, Valdevit A, Anderson L, Inoue N, and Hauser D., 2000, “Biomechanical Evaluation of Dual-Energy X-Ray Absorptiomety for Predicting Fracture Loads of the Infant Femur for Injury Investigation: An In Vito Porcine Model,” Journal of Orthopaedic Trauma, 14(8), pp. 571-6. Prange M, Luck J, Dibb A, Van Ee C, Nightingale R, Myers 8., 2004, “Mechanical Properties and Anthropomety of the Human Infant Head,” Stapp Car Crash Journal, 48, pp. 279-99. Prange M, Meaney D, Margulies S., 2000, “Defining brain mechanical properties: Effects of region, direction, and species,” Stapp Car Crash Journal, 44, pp. 205-13. 78 Prange M. and Margulies S., 2002, “Regional, Directional, and Age-Dependant Properties of the Brain Undergoing Large Deformations,” Journal of Biomechanical Engineering, 124(2), pp. 244-52. Schneider L, Lehman R, Pflug M, Owings C., 1986, “Size and Shape of the Head and Neck From Birth to Four Years,” Washington, DC. The Consumer Product Safety Commission, Report No.: UMTRI-86-2. Silva M, Keaveny T, Hayes W., 1998, “Computed T omography-Based Finite Element Analysis Predicts Failure Loads and Fracture Patterns for Vertebral Sections,” Journal of Orthopaedic Research, 16, pp. 300-8. Thibault K. and Margulies S., 1998, “Age-Dependant Material Properties of the Porcine Cerebrum: Effect on Pediatic Head Injury Criteria,” Journal of Biomechanics, 31(12), pp. 11 19-26. 79 CHAPTER SIX CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE WORK The previous chapters describe the material and fracture characteristics of the developing porcine skull. The bone and suture properties were obtained through 4-point bending experiments on beam specimens removed from the cranium. The parietal bone was impacted under contolled laboratory settings to produce fi'acture patterns with regard to age and surface interface. A simplified model of the infant cranium was developed in a finite element program for comparison of theoretical stesses and stains with observed fracture patterns. The model was used to simulate clinical cases of bilateral crushing injuries as well as blunt impact tauma to the parietal bone. In Chapter 2, beam specimens cut tom porcine crania were tested in 4-point bending at 25mm/sec until failure to obtain material properties of the bone and suture. While the mechanical properties of parietal bone and coronal suture did not change significantly with age, bone specimens showed an increase in bending stiffness with age. Bone-suture-bone specimens showed an increase in bending stiffness only between 14 and 18 days of age. Correlation analyses using existing and new data to compute the bending rigidity of infant parietal bone specimens suggested that days of pig age may correlate with months of human age during the most common time fiarne of childhood abuse cases. Future studies should investigate the possible rate sensitivity of the developing bone and suture, specifically focusing on high rate loading to simulate impact rates. In Chapter 3, skull fracture on infant porcine specimens with respect to age and interface were documented in an effort to identify fracture characteristics for a 80 developing skull under known conditions. Fractures on the infant porcine skull due to impact with a flat surface initiated at points outside the zone of contact for both rigid and compliant impact interfaces. The skulls of young infants suffered relatively more extensive tacture damage stiking a compliant versus a rigid interface for the same impact energy. As the animal aged, however, the damage caused by the compliant interface decreased until no damage was present at energies causing fracture with a rigid impact interface. Future studies should examine additional interfaces of impact, especially involving common surfaces of potential accidental injury such as carpeting or playground rubber. Additionally, the extent of propagation and fracture patterns developed under higher energy impacts should be documented. In Chapter 4, four clinical cases of fatal crush injuries to young children between 1.5 and 6 years of age were documented with correlations between modeled stess and clinically observed fiacture patterns. Regions of high tensile stess in all four cases were located on the basicranium and appeared to follow the fiacture path(s) extending tom the location of maximum stess. These results suggest that pre—failure stess field diagrams may give some insight into the projected fracture propagation paths, although prediction will not be exact. Also, these analyses of the modeled stesses and the clinically observed fiactures suggest that quasi-static bilateral loading of the cranium may lead to predictable fracture of the basicranium, possibly independent of age. Additional case studies would be helpful in validating the model further. Future studies may also be able to obtain more accurate results using CT scans of infant crania. With CT scans, the actual density distribution within the skull could be accounted for rather than the approximation used in this study. Also, the skull dimensions and basal landmark 81 geomety were approximated. Approximate thickness changes with age were the only geometric property altered. In Chapter 5, the impact conditions generated experimentally for rigid and compliant interface impacts to porcine skulls of various ages were recreated on a simplified finite element model of a cranium. Experimental results showed that blunt force trauma to the parietal bone consistently resulted in fracture at the bone-suture boundaries of the parietal bone, which correlated well with areas of high stain. The theoretical stains also showed similarities with the significantly higher rates of suture failure in young specimens than in old specimens. Interestingly, the larger contact area generated during compliant interface impacts also seemed to produce stains within the bone distributed nearer the parietal bone boundaries, which may have promoted propagation of fractures accounting for the greater amount of observed damage on the cranium produced from compliant interface impacts. Future studies should develop a model of the skull from exact dimensions of a CT scan. The information presented in this thesis documents baseline information for fi'actures generated on the developing porcine skull. An age-based correlation with porcine specimens may provide an important biomechanical tool for the study of pediatric skull fracture in efforts to investigate cases of childhood abuse. This information may prove relevant in clarifying unexplained perimortem events in cases of injuries to the pediatric cranium. Modeling cranial fiacture is particularly important in unwitnessed injury cases or in cases where information tom examination of skin and scalp is not possible due to skeletonization. These data present several factors that must be examined on a case-by-case basis and may ultimately lead to the deciding factor in a courtoom verdict. 82 APPENDIX A RAW DATA FROM CHAPTER 2 83 00... 0 ...0 ..000.0 00.00.. .000 0 ...000 .00 P0. ..0 3. 0000 00. .. 00.0 0000.0 00.0? 000.0 v0.09. 00.0 00.0 3. ..000 00. F 0.0 00.0.0 00.00 .. 00 .. .0 00. ..00 0 0.0.. 00.00 3. ....00 00. .. 00. _. 3.00.0 00. .0? 00. ...0 .000 0 .. ....0 0.. 0000 0 ...0 0v. .. 030.0 00.00 F 000 .0 #0000 00.0.. 00.00 0.. 0000 00.0 00. 0 0000.0 00.3. 000.0 .0003. 00 ... 0.. .0 .1. .0000 #0.. .....F 030.0 00.00 0.120 00.....0 00.0.. 00.... 0.. 0.00 ..0... 00.0 300.0 8.3.. 00 F .0 00.000 00.0.. 00... .. 0.. 0 ..00 0... .. 006 .0000 00.00 .. 80.0 00.000 00. .. .. 00 .0 0 0000 00... v0... 00.0.0 0 ....V0.. .00 .0 00.00. 00.0.. 00 ..V r 0 0.00 00... 00... 0000.0 00.00v 00.? 00.000 00.0.. .40.; 0 0.00 .V ..... 006 0000.0 0 ...00.. 00 .. .0 0v.000 0 w.0.. 00 .3. 0 0000 0... .. 00.0 0¢00.0 V0000 V0000 V0. .. _. 00.00 0 0000 00. .. 00.0 0.00.0 00.8 0 ..0.0 00.0.... 0 ...0.. 0 .. .0 .. 0 0000 00.0 3.0 00 ..0.0 00.00.. 8 .. .0 0.0000 00... v 00.0.. 0 0000 00. .. ..0.0 0000.0 0.0.000 00.. .0 0.0000 0 ...0.. .00.. 0 0000 0.0.? 0 ...0 .030 0 ...000 000.0 00.0 ..0 .00.. 0.000 0 0000 00... 00.0 0000 .0 0 ...000 8V0 00.00.. ..0..... 00.0.. V 300 V ...P 0.00 0.00.0 00.00.. 00 .. .0 0 0.000 00... .. 00.0.. v 0000 00.0 00... 0.00 .0 00.00.. 00 .. .0 00.000 00... .. .. ...0.. v 0000 00.0 0.0.00 0000.0 3.00... 000. .. 00.000 00 .00 00 .v0 0 @000 ..0.0 0000.0 00. v0.0 00 .. .0 00.000 00.0.. 0.0.3. v 0000 .88. $00. €8.20 30.2. .3 .82. .EE. $9.on 8332 868m 59m as 0.85 ...: Foam 58m 38505 328985 .2. .08.. a? 85.8% .3088on 3000508050 me new»: 050:2. genre :80 028:8 8% 33. ...< 03:. 84 85 «3 SN Named 3.8 82. ~38 3m «5. a mean. 3..- 88 88a 8R mood ~38 8.4 :3 N Eda 9N Ea 888 8.0a Sod were 8.... E. a mean. 2a 3.. 288 8.2 mood 88a ta LB a mean. 8a and 823 8.8 Sod 8.8a 8s «3 a. amen. 84. 8a $55 8.8 Sod 82$ 5... 5m 2 Ram New as 885 8.8 :3 9.52 :6 8s t 5%. mod an. 888 «tea as! 3.8.. 5.9 ES 8 8a.... 01 «3 83¢ 3.8..” $3 meta and. 8.9 a. 2am Es. .aaov Eases. 802. 365$ .52. .55. 85.95 8.38.2 Pecan Ease ...: macaw ...: 52w «.8an 558935 2:83 a? 858% cos—5:00 .~.< «Bah 8.: 3.0 8I~00 :88: 8% 00.08: 8.0 8.0: 8: 8%. 0:: 5.0 8:00 :50 80.0 8:8: 80... 0:8 8: :0%. 80.: 8.0 80.0 888 8.8: $8 8.0: : :%. 00.: 8.8 8:0.0 :8: 0.60 0%: 8.0 00.0 8 0:%. 8.: 8.0 080.0 8.8: 80.0 80.80 8.0 8.: 8 8%. 8.8 :8 080.0 8.8 80.0 8.08: :00 8.0 0: 00a... 00.: :0 080.0 8.8 :30 ::.:::: 8.0: 8.8 0: 8%. 0:.: 8.0: @800 8.88 08.0 8.80 8.: 8.0: 0: 8%. 8.: 8.: :80.0 8.8: 80.0 8:0 00.: N08 0: 8%. :N: 8.0 :::0.0 80.88 80.0 8.88 00.8 8.0 0: 3%. 8.: 88 080.0 8:8 80.0 8.80 8.: 80.8: 0: 8%. 8.: 8.0 0800 8:0: 08.0 8.08 8.0: 0:8 0: 8n... 8.: «0.0 00:00 8.8:. 83 8E 8.0: :08 0 8%. 8.: 8.8 8:00 8.80 :00 8.088 8.8 8.8 0 8n... 8.: «0.0 8:00 8.88 80.0 8.80 8.: 8.0 0 8%. 8.: :08 88.0 8.08 8:0 8.08 8.8 00:0 8 00%. 8.: 8.0 80.0 8.80 80.0 8.88 00.: 8.8 8 Sn... 8.: 8.8 080.0 8.80 80.0 8.08 88 00.: 0 00%. 8.: M08 0800 8.5 :00 8.88 8.0 8.8 0 00%. 8.0 8.: 880.0 8.800 80.0 :38 8.0: 8.8: 0 8%. 8.0 8.0: 880.0 8.88 08.0 00.08 8.8 8.8 0 8%. 00.: 8.0 080.0 8.8 08.0 8:8 8.0 00.8 8 8%. 8.0 :8: 88.0 8.88 80.0 8.800 8.0 8.0 0 88 8.0 :E 88.0 8.80 80.0 8.08 00.0 8.: 0 8%. 80.: 8.0 800.0 8.08 ::0.0 8.08 8.0 8.8 :. 8%. 00.: 8.0 00:00 00.0I0w. 80.0 00.80 8.8 8.0 v 8%. AEEV Ann—0v AEE\EEV 3&2 g. 36cm AEZV AEEV m8§2£ 8.38.2 8.0000 0.90m .05 880.5 0.80 w8080 000580.85 2.08.. 03 005.080 858608 ocoo. 8339833 MO $08 9:083 «509+ :88 380:8 33 30m .N.< 239 86 87 8d mg 955 2.3 «8.0 8.29 NS. 3d a man. one «:3 8.8 Bod 52$ 8m «mm a 3%. Ba 8+ 885 3% mad 2.39 96 8d a 9%. and mom 823 8.8 $8.0 «3 8.2 E «Eu 3.. an 835 2.5 «36 ~33 3.» m3 2 8%. m? Re 25o 8.2. Bod «tamfi m3 3m 2 8%. one 385 2.0: mood $.88 8.» 5o 2 NE SN 8s 885 8.5? Bod 3.” mod S 3%. 84 So 285 8.8 Sod $.on an EN 2 2% a: 3m E86 :afi «Mug 38$ 8.» 3.2 3 Ram AEEfi ANQOV $55595 awn-2v A3 5cm AEEV AEEV ”85.2.: 3:32 96:8 526 .5 33w .5 595 amoéam “528335 2 83 o? cogoogm 62:5ch .~.< 033,—. ..0.0 00.0 0000.0 8.00 000.0 00.0 0v.0.. 3. 3mm 00... 3.0 00 50 00.000 00.0.. 00.0.. 3. 001m 00. .. ..0.0 800.0 006 ..0 00.005 00.0.. 00.00 S. ..0n.n_ ..h. .. h ...v 00 ..00 00.00 500.0 00 .000 00.0 00..q 3 ....Qn. 0V. .. 00K #0 ..00 00.000 000.0 00.000 00.0 ..0.0 0.. 0..n.n. 00. .. 00.0 ..n ..0.0 00.00.. ..000 $.03. ..v.0 00K 0 .. 00n.n. 0n. .. 00K 00.00 000.0 00.00 .... _. ...0.. 5.00 0.. 00am 00.0 000 00 ..00 0 ...00.. 000.0 00,000.. 0.}‘ 00.0 0.. 00n.n_ 00. .. 00... 00 ..00 00. v0? 0000 .000.» 00.0 010 0.. 0_.n_a 0 ... .. RN ~“.0000 00.00? 00.0.. 0.000 or 0..n.n. 00. .. 00.0 00 ..00 3.000 000.0 00 ..K0 .0... .. 00.0 0.. Sin. 00... 00.0 0.. ..0.0 00.00 000.0 0v.000 0h .0 510 0.. 0am 00. .. 00.0 0000.0 000 ..0 by .000 0 .. .0 .. 00.0.. 0.. 0......n. 00. .. 00..‘ 0h ..00 00.03 5 ...0 00.0 ..0 00.0.. 00.0 0 0_.n_n. 0..... 0.00.. 00 5.0 3.. ..0.. 05.0 50.30 50.0 50.0 0 tan. #0. .. 00.0 ..0 ..00 00.00 V ..00 00.000 «0.0 50.0 h 00n_n. ..0.0 0000.0 010.0 0 ..00 3000 00.5 00.0 n noun. 00. .. 0.00 #0000 00.00.. 000.0 0.000.. ..00 V0.2. 0 00mm 00.0 00... ..0000 000 ..v 0 ..00 00.0 0 0.20.. 00.0? 0 00n.n. 00.0 00.0 #0 ..00 00 ...0 00.90 0.0.. @0.3 0 vomn. N0. .. 00.0 0000.0 0000.. ...V00 0.. .000 00.0.. 00.0.. 0 81¢ 3V. .. 00.0 00 ..00 00.00.. 0 5.0 2.000 3 N 00.0 v 00n.n. 00.0 00.0 0000.0 00.0 ..0 000.0 00. .....0 00.0 {.0 0 00mm 8.0 00.0 00 ..00 00.000 0.000 0 .. .000 010 .. 00.0 0 01¢ 00. _. 00.0.. 0000.0 00.000 v 5.0 5.000 00.0 N00 v 00.: .0; 0 ..K 00 ..00 v ...000 v ..00 0.000% 00.0 0 V0 v 00%. .EE. ammo. €5.88. Ana—E A... .82W .EE. 822%: 3.822 86:8 595 .5 83m ...: 3.8m 59% 88:5 23883 .20 83 a? 553m £88.30». 280 .2330 ..o 38. 050:2. 3.00-.“ SE... 098:8 8% 33. .03.. 0.05.. 88 o..~ 8.” 8.3. 8.... .3... 3.08. .3 3.. .N 3%. ...... on... $5... 8.8. as... $33. mom was .N 3a.... a... 8.. .32. «I: Sod .35 8... o... .N my"... mum 2a 2.2. 2.8 .5... 3.8.. 93 8.. .N NE: m... an... .~.o.o 8.3 Bod 8.8.. 8... ......” m. 8a.... 6.. 8.. ...od N33 .8... «33. men an m. an: 3.. R... 38... 8.8 m8... 8.3.. and B... t .Nna 8.. ...m 0.8... .....m ...od 8.8.. om.» :3 m. 8% 8.. 8d 88... 8% mm... 3.8.. 8... SN w. 2%. AEE. 3&0. AEEEE... 3&2. 0... 5cm $52. ASE. «.852... 3332 9.68m swam ...: 32% ...: 52w 8&5 2958.35 .2. 83 on... 8588 02:55.0 .0.< 0..—ah. 89 3‘. F 00.0 08.0 00.05 00.0 «N000 v06 3.0 N 335 05—. _._..N #80 00.000 00.0 00.8w 00.9 0.. RN 0.. noun—n. 00... 00.N 000.0 v0.Nm_. 0.. .0 00.8» 00.2. 00.0w 0_. gown—n. 05.? 50... 08.0 0 _. .000 00.9 8.50m SN? 3200 0.. 0 En. 00... 00.N m 5.0 «5505 0.. .0 9.50m 50.? _. 00.3. 0.. v En. 0v; NNN 50.0 05.0NN 0N.0 «0.8N N _. .0_. 0.. .0? 0.. N En. 00d 8N0.0 50.000 «5.0 NEEN 00.N_. 9.0—. 0.. Sin. 00. _. 00... 3.00.0 mm. .0? 30.0 00. SN KN? m —. .vw 0? 00am 5N. _. mod h 5.0 50.0NOF v0.0 03.0mm 00... _. ..0.v~ 0 00am 00.0 ..NN 0 v0.0 3.0 ..0 5.0 0N.00 MK .0 _. No.0? 0 Von—n. v0.0 0 ...0 v v0.0 50.000? 00.0 «0.2.0 NONF 5.0m V sown—n. 00.0 00.0 _. V0.0 N0 .000 :20 00.0VN F0... _. 5.0—. 0 mun—n. A880 089 AEEEEV fins: 3 0520 AEEV 3.8on 3382 965m 226 .5 32% ...: 3.8m gnaw chEm 52.8335 20 28.. o? 559% .825 833. E mat—smog 0.02 2.258% oconéhfiambcoa .00 3m»: 0505: “500% 890 038:8 83 Bad 6.4 03:. 90 91 8a mg 83 8.8 5.0 8.08 80 N3 3 38%. m3 F3. 23 9.8 55 8.8: 23. 8.5 8 39.; 3.: 3.0 :3 8.3 So $.32 23 5m 3 £3; SN 8.” 22. 8.8 36 8.2a .3. m3 2 £8“. «2 8.0 83 ”EN 86 8.89 «B 8.: 3 88¢ o? 53 8.: 5o {.33 5e 20 3 32“. m: E 93.0 8th 5.? $.58 E: 3.8 3 8.8... NE SN Rod 83 Q3 9.3” F; 8a 3 «NE; 8.? Ea BS :3” ”3 EN: :2 8% 3 £8; m: Bu ~86 SN... 85 3.8,. 8.» 8.0 3 «5%. :4. ta 82. 8.8 So 3.8m 8.5 8.0 2 88.”. a: 955 No.8? 3.0 33% 8.2 3.8. 2 «3%. m: «.3 $3. 3.8? 2.0 2.9% 3.: 8? 0. 88a «2 8.” 83 $.52 ”We 5.8.... 3.9 8.0 .V «8%. 2:5 fine 3.52.5 was 3 8m E20 :55 $05.05 3.332 320m 595 .5 $26 .5 :05 $9an $58520 A20 28.. m9. cmgoam .8355 25$ 20 E00 :33 25863.9. econ—653-080 20038 .80 $30 050:3 0500-0 80.0 028:8 300 Bad .m.< 0:35 Table A.6. Porosity data from stained cross-sections of perpendicular bone specimens. Bone Total Porosity (Pixels) (Pixels) (%) PP02 33547 41699 19.5% PPO3 25797 31059 16.9% PP04 61503 68346 10.0% PP05 69804 81977 14.8% PP06 76844 79827 3.7% PP07 52030 56810 8.4% PP08 45056 50665 11.1% PP09 43168 56322 23.4% PP10 43379 62302 30.4% PP12 37707 43774 13.9% PP13 67578 74024 8.7% PP14 66050 76634 13.8% PP15 29341 34137 14.0% PP16 66006 85252 22.6% PP17 29099 33402 12.9% PP18 87797 100901 13.0% PP19 75149 99736 24.7% PP20 67324 86982 22.6% PP21 84893 105391 19.4% PP22 76138 84575 10.0% PP23 52362 57918 9.6% PP24 10420 1 1466 9.1% PP25 41367 45319 8.7% PP26 36817 40503 9.1% PP27 91413 111205 17.8% PP28 55797 71386 21.8% PP29 62829 72158 12.9% PP31 109448 1 18586 7.7% PP32 123319 160033 22.9% PP33 55829 61288 8.9% PP35 111136 139891 20.6% PP39 236070 305571 22.7% 92 Table A.7. Porosity data from stained cross-sections of parallel bone specimens. Bone Total Porosity (Pixels) (Pixels) (%) PP02 59955 74687 19.7% PP04 23619 25450 7.2% PP05 32821 38448 14.6% PP06 121544 136713 11.1% PP07 47076 51903 9.3% PP08 72309 85020 15.0% PP09 52524 58510 10.2% PP10 69059 86167 19.9% PP11 135612 168833 19.7% PP12 68397 76470 10.6% PP13 67312 76059 11.5% PP14 62774 72731 13.7% PP15 67010 77393 13.4% PP16 50708 66943 24.3% PP17 65382 71443 8.5% PP18 89707 97636 8.1% PP19 51343 65791 22.0% PP20 1 12397 137864 18.5% PP21 39143 47067 16.8% PP22 51157 57322 10.8% PP23 27210 29493 7.7% PP24 12117 13427 9.8% PP25 21 753 24577 1 1.5% PP27 67091 105630 36.5% PP28 37794 42043 10.1 % PP29 72273 85481 15.5% PP30 51137 56511 9.5% PP32 100422 125993 20.3% PP34 122032 141013 13.5% PP35 146688 168171 12.8% PP39 5921 19 889588 33.4% 93 APPENDIX B RAW DATA FROM CHAPTER 3 94 mm 8.8 8.. 8.8~ .8.” 8888 8.8 8.88 88 V. BE 8.8. K. ~88 .88 880.0 85. 8.88 88 8 ... 08E 8- 8.. 8.88 o8.» 88... 8.8 2.88 88 8 ~. 88: m... 8.. 8.8. ~8.~ 88... 8.8 8.88 88. 8 ~. 82.. 88 8.. ~..8. ~8.~ 88.0 ~..o. ...8~ m8. 8 .. 82¢ -- 8.. .38 88.. ~83. 8.8 8.8.. 88 8 8 82m 8.8 8.. 8.8 8~.~ 88.0 8.» ~38 88 8 8 ~85 8.8 8.. 8.5.. §.~ 2.8.0 8.8 8.8.“. 88 8 m .8: 8.8 ~... 8.8. ~8.~ 88... 8.8 8.88 88 8 h 88: 8.8 .... 8.8 ~8.~ 888 8.8 8.8 88 8 8 ~85 ...8 8.. 8.8 ~8.~ 88... 88 8.88 88 8 8 82¢ .8. 8.. 8.8. .8.~ 88... 8.8 8.88 88 8 o 82.. o.-. .... 2.8. 8¢~ 888 o... 8.88 88 8 m 82.. 8.8. 8.. m~.o.~ ~88 .83. 8.8 8.8 88 8 8 ~88. 8.8. 8.. 8.8. ~8.~ 888.0 8.8 8.5, 88 8 m 82... 8.28 8.. 8.8. 8.8 880.0 8.8 8.88 88. 8 .. 82a 8.8 8.. 8.8. 8- ~80... 8.8 888 88. 8 m 82“. Q8 3.. 8.8 ~8.~ E88 8... £88 88 8 m .8... 8.8. 8.. 8.8 8.. 8808 8.0 8.8.. .8. 8 ~ 08: 8.8 8.. 8.8. 8.. 88.0 E... 8.8.. v8. 8 . ~ 88: -- -.. 8.8.8 88... .83. 8..., 883 .8. 8 ~ .8: £55 8.5 8525 3.82m E 88.8 8.5 83 ea: .2. 83”. a. 8a: 85 8.89 85.883 82 $8on 8858 2.8“. .085 8 52.8885 .08.... 8.8“. .88: 85 8< 3862 .88.... 8.83:. a»... 828 8.02.8 38 as. ...m 2.3. 95 odov N. ..m modmm .V. ..N _ m.oo.o _ an...» 8.33. mo .N mm mm m. .5 BEN end .0. .8 wood owood 8.0 Eden. oo.w ow mm x. . ._n_ QR mm. not . w mEmu =38 - 8882 .0: Emu .cOzocémE .oSano moo. mm m. omoE 8.8 8.. 8:89 085.8 =33 - 8282 .9. 98 .5888... .28 .80 88 8 ~ 0 SE «.mm 86 notomm. wmemu =83 - 3882 no: 8% 60:05.8... .55an0 mom. on H oon «.5. RN 3.8. m3... 886 .mN. Node... m5. om mm on ._n_ 3.3 n ..m 8.0.... mmmd ovood .wd. wndmm. m5. om mm cm :1 mdmm vmd NmNo. $5.0. «mood 3N. wvdmw. m .m. on om mv ._n_ m. ...N 3N 3.3. Nam...“ 9.8.0 o... .. 5.3.0. m5. om mm 8 ._n_ mdo. mm... 5.09. m .m.m Nmood ts 868 m5. on em ov ._n_ nee 3.. 8.3% mom. m .86 pm... 8.08 m .m. 0.. wm an ._n_ Qmmm Q... m ..EN 8.6 38.0 .vd m5. on «N 3:; 5.9. 3... Q... .m. EON 58.0 and 8.3K m5. om mm 9. .E Qo§ $8. $58 ~86 good mmd. ondww m .m. om om 95?. 9%. 3.. 8.8. 30... omood and 0.95 80. mv m. N. ._n_ 3%. 3... 868 mom. .Nood mo.m ...08 wow. mm 5. 59¢ 98. 8.. mm. .m. End Rood and 3&5 80. mm o. w .oE @va 8.. NNNE 08.. owood 09m 8. . .m 80. mm m. 391 me. 2.. modmm $68 886 m ..m modmw moo. mm m. amen. A~EEV .EEV €28 3 8.8m 3 cozlmso Ass. 83 xm.‘ A2. 85“. 8. 84.5. Es ENE :2 _ Spam 8?. 30.2th 30:85 228“. .08.; um EwEoomfiwE .08.; 9:8”. 290: no.0 09.. mmmhm>< 62.5250 ...m Saab 96 Table B.2. Skull damage measurements for rigid interface impacts. Bone Suture Total Fracture Damage Damage Specimen (mm) (mm) (mm) P001 8 0 8 P1003 11 0 11 P1050 31 0 31 P1004 47 0 47 P1079 47 0 47 P1016 4 0 4 P1023 38 0 38 P1062 7 0 7 P1076 0 0 0 P1013 28 0 28 P1049 26 22 48 P1007 25 0 25 P1008 25 0 25 P1071 80 0 80 P1032 17 0 17 P1066 8 12 20 P1057 66 10 76 P1030 25 35 60 P1045 15 21 36 P1046 21 30 51 P1077 88 0 88 P1028 20 0 20 P1064 79 10 89 P1018 22 0 22 P1067 83 10 93 P1112 5 0 5 P1053 29 0 29 P1142 10 0 10 P1144 10 0 10 P1138 60 0 60 P1140 85 0 85 P1146 105 0 105 P1148 85 0 85 P1134 0 0 0 P1136 85 0 85 P1080 30 0 30 P1010 14 0 14 P1020 23 0 23 97 0.00: 2.0 8.80 :80 3.8.0 8.0 00.08: 82 00 2 80.0 0.08 8.: 8.000 000.0 008.0 8.: :0.:0: 82 8 0: 20.0 00:0 8.: ::.0:.: 80.0 88.0 8.:: 8.80 82 8 0: 20.0 0:8 :0: 000:: 8:0 880 ::.:: 00.000 80: 00 0: :8: 0.0 :0 8.: 000:: 20.0 88.0 00.:: 80:0 82 8 N: :00: :8: 8.: 8.0:: 0::.0 88.0 2.:: 8.08 82 8 0: :00: :0: :0: 8.:0: 08.0 88.0 8.: 8.80 80: 00 :: 80E «:0 8.: 8.:0 0:00 880 :02 8.08 82 8 2 000.0 0.80 2.: 00:0: 08.0 88.0 8.0 00.8: 82 0: 0 08.0 0.000 8.: 00:2 80.: 008.0 8.0 000:0 82 00 0 «:0: 0.8: :00 8.00 80.0 :080 8.: 8.20 82 0: : 08.0 0.00 :0: 0:0: 0::.: 008.0 00.0: :88 82 0: : :2“. 0.8 8.: 8.8 80.: :080 8.0 2.80 82 00 0 08.0 0.0: 8.: 8.0: 0:00 0:80 8.0: 00.80 82 00 0 0:0: 0:0: 000 :08 02.0 88.0 8.: 8:8 82 00 0 08.0 :8: 8.: 8.:: 80.0 :080 8.0 808 82 00 0 80: 0.00: 8.0 :08 80.0 0:8.0 :0 8.8:. 82 00 0 08.0 0.00 8.: :08 0:00 88.0 8.0 8.80 82 00 0 0:2... 0.8 00.0 8.0: 80: 008.0 8.0 00.08 82 0: 0 08.0 0.08 00.: 8.0: :00: :80 8.: 8.08 82 00 0 :00: 0.00 00.: 2:0 0::.0 88.0 :00 8.00:. :8: 0m 0 08.0 0.0: 2.: :08: 0::.: :080 ::.0 00.20 80: 0: 0 80.0 :80 00.: 88: 0:0: 88.0 0:.0 0:.:0:. 82 0: N 08.0 0.8: :0: :30 08.: 88.0 8.0 2.08 80: 2 0 80.0 meg 8.5 828 3 3080 888050 8.5003 x02 A2820“. av 0002 0:9 8:03 85800 8a 085.2: 08080 05:0". 8005. :0 8580.85 .805. 05:0“. 203: 020 00< omEo>< .3805: 8835 0800: 82 a: 080088 :50 02808 300 300 .0:— 20.; 98 00: ::.0 v0.00: 000.0 300.0 00. :: :0.:00: 00:0 00 00 00 :_n_ 0. :0:. ::.0 :0.:.0 :000 0:00.0 0:. :0 00.000: 0 :0: 00 mm :0 :E 0.0:: : 00.0 v :.0: 0 ::.0: 0000.0 ::.00 :0. :0:: 0 :0: 00 00 00 :0. 00:0 0:.0 00.3: 00:.0: 0000.0 00.:: 00.00:: 0 :0: 00 00 03.1 0.00: 0:. : :0. ::: :000 030.0 :0. :: 0:000 0 :0: 00 00 :v :n. 0.:0v: :0.0 00.00: 00:.0 0000.0 :00: 00.:00 0:0: 00 v0 :35 0.000: 00.: 000.00 000.0 0000.0 ::.0: 00.00: 0 :0 : 0:. v0 00 :E 0.::: 0:.0 v0.00: 00:.0 300.0 00.0 0:00: 0 :0: 00 00 0:. :5 0:0:. 00. : 00. : :: :00.: 030.0 00.: 0:0: 00 00 00:5 0.0 :0 :0. : 00.00: 00 :.v 3000 00.0: 00.000 0 :0: 00 00 «00?. 0.00: :0. : :0.00: v ::.0 0000.0 00.0 00.0 :0 000: 00 :: 000E 0.0:0 00.: 00.000 :0:.0 0:000 00.0 00.00 000: 00 0: 000E 0.: :0 0:. : 00.00: 0.00.: 0000.0 0 :0 0:.:: 000: 00 0: 000E :. :00 00.: 00.00: 000.: 0000.0 :00 00.:00 000: 00 0: 000E :000 0:.0 00.00: 000.0 3.000 00.0 03000: 000: 00 0: :00_n_ €55 055 0520 0.10.000 E 000050 025 08.. x05. :20 020“. :00 80.2 0:3 3:03 002800 00.:: 002222.: 03:50 83:0”. 8008. :0 Em E80305 800 E. 05:00”. £081 no.5 00< 008020. 60:50:00 .05 030,—. Table B.4. Skull damage measurements for compliant interface impacts. Bone Soure Total Fracture Damage Damage Specimen (mu!) (mm) (min) P1034 53 0 53 P1035 117 0 117 P1056 52 0 52 P1005 112 0 112 P1081 30 0 30 P1060 9 11 20 P1078 30 0 30 P1022 57 30 87 P1024 17 20 37 P1063 36 16 52 P1012 7 14 21 P1025 34 40 74 P1011 190 0 190 P1040 78 0 78 P1072 16 0 16 P1069 11 0 11 P1065 132 0 132 P1059 129 15 144 P1031 141 0 141 P1037 0 41 41 P1021 75 0 75 P1014 103 0 103 P1015 65 22 87 P1041 12 53 65 P1027 105 0 105 P1029 95 0 95 P1052 41 0 41 P1082 50 0 50 P1068 27 10 37 P1054 23 0 23 P1143 0 0 0 P1145 0 0 0 P1139 60 0 60 P1141 0 0 0 P1147 0 0 0 P1149 40 0 40 P1135 0 0 0 P1137 0 0 0 P1122 0 0 0 100 T 1| “11111111111111Wiles 0 3 0 6 2 L 8 3 2 3 ..."..l..|.w.".. 3 12 9