/1. ’ '7') 3' <3 / f [H533 This is to certify that the thesis entitled FRACTURE PATTERNS OF THE DEVELOPING SKULL A'ITRIBUTABLE TO DIFFERENT IMPACT SCENARIOS presented by BRIAN J. POWELL has been accepted towards fulfillment of the requirements for the MS. degree in EngineerinLMechanics @Aflajor Professor’s Signature W 0/ 0 Date MSU is an Affinnative Action/Equal Opportunity Employer LIBRARY Michigan State University 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 5/08 K:/Prclecc&Pres/ClRC/Dale0ue indd FRACTURE PATTERNS OF THE DEVELOPING SKULL ATTRIBUTABLE TO DIFFERENT IMPACT SCENARIOS By Brian J. Powell A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Engineering Mechanics 2010 ABSTRACT FRACTURE PATTERNS OF THE DEVELOPING SKULL ATTRIBUTABLE TO DIFFERENT IMPACT SCENARIOS By Brian J. Powell Forensic anthropologists and pathologists frequently rely on fracture pattern analysis to determine the causation of trauma in pediatric abuse cases. However, due to a lack of pediatric skull fracture tolerance data it is often difficult to diagnose whether the injury was inflicted or accidental. The research presented in this thesis utilizes an in situ porcine head model and finite element analysis to assess the degree and pattern of skull fracture in two different impact scenarios: entrapped and fiee fall. Chapter 2 documents the skull fracture patterns generated due to a high energy blunt impact using a rigid and compliant interface. The porcine specimens in Chapter 2 were entrapped in a bed of air- hardened epoxy in order to prevent translation of the head during impact. In Chapter 3, the porcine model was used to assess the degree and pattern of fracture due to a free falling head impact. These specimens were dropped with an equal level of impact energy as those in Chapter 2 to compare the patterns of fi'acture. In Chapter 4, the data from both impact scenarios was compared to assess the differences of fracture. A finite element model was constructed to provide theoretical insight of the principal tensile stress directions for better understanding of the fractures generated in each impact scenario. The information presented in this thesis may be helpful diagnosing whether the head was constrained in a forensic case. ACKNOWLEDGMENTS First and foremost, I would like to thank Dr. Roger Haut for his support and guidance in my research and education over the past two years. Also, Dr. Todd Fenton for helping me understand some basic concepts in forensic anthropology. I’d like to thank Cliff Beckett for all of his technical assistance and Ed Reed and Star Lewis for supplying the porcine specimens used for this research. I would like to thank Tim Baumer, M.S. and Nicholas Passalacqua, M.S. for their dedication to the project as a whole. I would like to thank my friends for all the laughs, support and good times through graduate school. Finally, I’d like to thank my family for their unwavering support through the good times and bad. I could not have done it without you! iii RESEARCH PUBLICATIONS BY THE AUTHOR PEER-REVIEWED MANUSCRIPTS Powell B, Passalacqua N, Baumer T, Fenton T, and Haut R. Fracture Patterns On the Infant Porcine Skull Following Severe Blunt Impact. J. Forensic Sciences. (in review) Baumer T, Passalacqua N, Powell B, Newberry 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 TG, Powell BJ, Fenton TW, Haut RC, 2009, “Age Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human,” Journal of Biomechanical Engineering, 131(11), pp. 111006-1-6. PEER-REVIEWED ABSTRACTS Powell BJ, Passalacqua NV, Baumer TG, Fenton TW, Haut RC. Fracture Patterns on the Infant Porcine Skull Following Severe Blunt Impact. American Society of Mechanical Engineers Summer Bioengineering Conference, Naples, Florida, 2010. F enton TW, Passalacqua NV, Baumer TG, Powell BJ, Baumer TG, Newberry WN, Haut RC. A Forensic Pathology Tool to Predict Pediatric Skull Fracture Patterns, Part 2: Fracture quantification and further investigations on infant cranial bone fracture properties. American Academy of Forensic Sciences (AAF S), Seattle, Washington, 2010. Van Wyhe RC, Powell BJ, Haut RC, Orth MW, Karcher DM. Reducing the growth rate in turkeys improves femoral bone quality. International Poultry Expo, Atlanta, GA, 2010. 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] iv TABLE OF CONTENTS LIST OF TABLES .......................................................................................................... vi LIST OF FIGURES ........................................................................................................ vii CHAPTER 1: Developing the Porcine Model — An Introduction .................................... 1 CHAPTER 2: Fracture Patterns on the Infant Porcine Skull Following Severe Blunt Impact ............................................................................................................. 10 Abstract .............................................................................................................. 10 Introduction ........................................................................................................ 12 Materials and Methods ....................................................................................... 13 Results ................................................................................................................ 17 Discussion .......................................................................................................... 23 References .......................................................................................................... 28 CHAPTER 3: Porcine Skull Fracture Length and Pattern Generated From a Free Fall Impact ................................................................................................................ 31 Abstract .............................................................................................................. 3 1 Introduction ........................................................................................................ 33 Materials and Methods ....................................................................................... 34 Results ................................................................................................................ 38 Discussion .......................................................................................................... 43 References .......................................................................................................... 47 CHAPTER 4: Validation of Whole Head Drops in Free Fall Using a Porcine Model ........................................................................................................................ 49 Abstract .............................................................................................................. 49 Introduction ........................................................................................................ 51 Materials and Methods ....................................................................................... 52 Results ................................................................................................................ 55 Discussion .......................................................................................................... 60 References .......................................................................................................... 64 CHAPTER 5: Conclusions and Recommendations for Future Work ............................ 66 References .......................................................................................................... 70 APPENDIX A: Raw Data fiom Chapter 2 ..................................................................... 71 APPENDD( B: Raw Data fi'om Chapter 3 ..................................................................... 77 LIST OF TABLES Table A. 1 . Raw data fi'om high energy rigid interface impacts ..................................... 72 Table A2. Thickness, fracture length, and contact area measurements for high energy rigid interface impacts ...................................................................... 74 Table A3. Raw data from high energy compliant interface impacts ............................. 75 Table A.4. Thickness, fracture length, and contact area measurements for high energy compliant interface impacts .............................................................. 76 Table 3.]. Raw data collected fi'om free fall rigid interface impacts ............................ 78 Table 3.2. Fracture length and contact area measurements for free fall high energy rigid interface impacts ...................................................................... 8O LIST OF FIGURES Figure 1.1. Anatomy of the infant (top) and adult (bottom) human skull ........................ 2 Figure 2.1 . Orientation of the right parietal bone with rigid impact interface ............... 14 Figure 2.2. The drop tower schematic with GAM shown .............................................. 15 Figure 2.3. Peak impact force versus age for both the rigid and compliant interfaces ........................................................... ‘ ........................................... 1 8 Figure 2.4. Contact area as a function of age for both rigid and compliant interfaces ...................................................................................................... 18 Figure 2.5. The average length of skull fractures as a function of age .......................... 19 Figure 2.6. GIS map of 2-9 day old rigid (a) and compliant (b) impacts at high energy ........................................................................................................... 20 Figure 2.7. GIS map of the 19-28 day old rigid (a) and compliant (b) impacts at high energy ................................................................................................... 21 Figure 2.8. GIS map of the 2-9 day old rigid (a) and compliant (b) impacts at low energy ........................................................................................................... 22 Figure 2.9. GIS map of the 19-28 day old rigid (a) and compliant (b) impacts at low energy .................................................................................................... 23 Figure 3.1. Schematic of the fi'ee fall drop tower ........................................................... 35 Figure 3.2. The drop trolley was raised to the necessary drop height and held with the electromagnetic solenoid clamp (a). The free fall impact was produced by disengaging the solenoid clamps (b). The mounting rod was free to move vertically until the impact force returned to zero. ............ 36 Figure 3.3. An overlay of force-time plots showing the characteristic rapid drop in force associated with skull fracture .............................................................. 39 Figure 3.4. Peak impact force with respect to age .......................................................... 39 Figure 3.5. Total fracture length with respect to age for free fall and entrapped impacts .......................................................................................................... 40 Figure 3.6. The average total fracture length versus age ................................................ 4O vii Figure 3.7. GIS map of the 2-9 day old age group for the flee fall (a) and entrapped (b) impacts ................................................................................... 41 Figure 3.8. GIS map of the 10-17 day old age group for the flee fall (a) and entrapped (b) impacts ................................................................................... 42 Figure 3.9. Fracture initiation sites located remote of the point of impact (represented by the bull’s—eye). Shaded areas represent flactures and each mark represents 5 mm of flacture length ............................................. 43 Figure 4.1. Simplified skull geometry used for finite element simulations. Rigid impactor positioned above the skull ............................................................. 53 Figure 4.2. Total flacture (diastatic and bone) length with respect to specimen age for both impact scenarios ............................................................................. 55 Figure 4.3. Peak impact force with respect to age for both flee fall and entrapped impacts .......................................................................................................... 56 Figure 4.4. Force versus time plots used to determine the duration of impact for both entrapped and flee fall impact scenarios .............................................. 56 Figure 4.5. Impact duration with respect to age for the flee fall and entrapped impacts .......................................................................................................... 57 Figure 4.6. Entrapped simulation showing impact site (bull’s-eye) and surrounding principal tensile stress directions. Four primary areas of maximum principal tensile stress were documented. Darker arrows represent higher magnitudes of tensile stress. .............................................. 58 Figure 4.7. Overlaid maximum tensile stress magnitudes and directions on a typical pattern of flactures flom an experimental entrapped porcine specimen. Shaded areas indicated documented flacture with each line representing 5 millimeters ............................................................................ 58 Figure 4.8. Free fall simulation showing impact site (bull’s-eye) and surrounding principal tensile stress directions. The largest magnitude of tensile stress was located near the point of impact .................................................. 59 Figure 4.9. Overlaid maximum principal tensile stresses on an experimental flee fall porcine skull ........................................................................................... 59 Figure 4.10. Gurdjian’s (1975) comparison of the degree of head trauma resulting flom a blunt impact to the head. The study suggest that longer duration impacts increase susceptibility of the skull to flacture ................................. 61 viii CHAPTER ONE DEVELOPING THE PORCINE MODEL — AN INTRODUCTION Head injury is the leading cause of death and permanent disability in children (Tabatabaei and Sedighi, 2008). It is difficult, however, to distinguish inflicted physical abuse flom accidental falls due to both scenarios producing similar types of head injuries (Billmire and Myers, 1985). Linear, complex, and depressed skull flactures have been documented in both cases (Reece and Sege, 2000; Wheeler and Shope, 1997). Indeed, skull flacture is diagnosed in 1 of 3 children who are investigated for physical abuse (Belfer et al., 2001). However, non-fatal injury resulting in skull flacture can occur in children less than 12 months of age flom a fall flom less than 3 feet (Gruskin and Schutzman, 1999). In a study by Belechri et al. (2002), 1,881 cases of falls flom bed height reported no fatalities. Attempts have been made in developing a model for pediatric trauma. One such model that has been met with some success scales adult head impact data to predict the impact response of the pediatric head (Prange et al., 2004). However, this model may be unreliable due to the head of a child being much smaller and geometrically different than an adult (Schneider et al., 1986). Also, little information is available on the similarity of skull fracture patterns between adults and children. Predicting trauma in infants using adult data is also problematic due to the significantly different mechanical properties and structural characteristics of the pediatric skull (Coats and Margulies, 2006). Fetal cranial bone is a very thin, non- homogeneous and highly curved material (McPherson and Kriewall, 1980) while adult bone is a three layered sandwich structure; two cortical bone plates surrounding a porous diploé center (Motherway et al., 2009). In addition, several soft tissue junctions (sutures) connect skull bones and allow growth of the brain during development. These sutures begin to ossify and interdigitate into adulthood and eventually fuse together, making the skull a single solid structure (Figure 1.1). The posterior and anterior fontanel are also present in the infant skull. These membranous tissues also allow expansion of the brain during development but eventually disappear after the skull bones fuse together. The posterior fontanel is typically closed between birth and eight weeks of age, while the anterior fontanel closes between nine and 26 weeks after birth (Knight, 1991). Anterolateral . fontanel Anterior fontanel Posterolateral ‘ fontanel Posterior Occipital bone Temporal fontanel Coronal SUIUIC Frontal . " Parietal Squamous suture ' ,3: {v . , . Coronal ‘_ Temporalh l .- a suture Lambdoid .. " ‘ ‘ suture ’ , , Sagittal Occrpital bone suture Figure 1.1. Anatomy of the infant (top) and adult (bottom) human skull. Because of ethical considerations, testing with infant humans is difficult. There is a current lack of studies that utilize infant cadaver specimens to produce data on the impact response of the pediatric head. Infant anthropomorphic surrogates have been used in a recent study (Prange et al., 2003) to document the rotational loading and deceleration conditions of pediatric head impacts. While these studies produce valuable quantitative data, there are no parts of these surrogates that fail at a given load to simulate bone flacture and the patterns of flacture that are generated during impact (Ewing et al., 1983). Animal models are commonly used for biomechanics studies to generate correlative data to humans. The porcine animal model, in particular, has been used flequently in recent studies. Margulies and Thibault (2000) used porcine skull samples to document age related changes in the mechanical properties of skull bone and suture in three-point bending. Baumer et al. (2009) performed similar studies using four- point bending of porcine skull beam specimens, documenting a skull development relationship of days in pigs to months in infant humans. In another study, Baumer et al. (In press) impacted the parietal bone on entrapped infant porcine heads using a gravity accelerated mass with rigid and compliant interfaces. The authors document flaeture initiation occurring at the bone-suture boundary, remote flom the location of impact. Another key finding was that a compliant interface generated more flacture than a rigid interface for a given impact energy. The study, however, was limited to only one, low level of impact energy. It has been thoroughly shown in the literature that the height of a fall (impact energy) has an effect on the degree of injury (Bertocci and Pierce, 2006; Barlow et al., 1983; Wilkins, 1997). A fall flom above 1.5 meters increases the severity of injury two fold, and a fall flom above 2.5 meters increases the severity of injury three fold (Macarthur et al., 2000; Chalmers et al., 1996). In the process of developing a more robust pediatric flacture model, the effect of impact energy on the length of skull flactures and the pattern of flacture must be examined. Additionally, the impact scenario must be investigated. Often children sustain head trauma flom accidental falls during the development of walking motor skills (Zimmerman and Bilaniuk, 1994). Death flom falls is the third leading cause of death in children aged 1-4 years (Hall et al., 1989). In the Baumer et al. (In press) study, the porcine head model was entrapped in a bed of air-hardened epoxy, preventing translation of the head after being impacted with a mass. In a study by Chason et al. (1966), a canine head model that was both fixed and flee to move during impact was struck with a rotary mass to assess the effect of head constraint on the degree of concussion. The authors document no skull flacture in the canine specimens, however, they did indicate that there was a higher susceptibility to concussive and neurological problems when the head was fixed during impact. It has been shown in the literature, however, that less force is typically required to produce a concussion than a skull flacture (Rosman, 2004), suggesting that the impact energy levels used by Chason et al. (1966) were not high enough to produce skull flacture. It is therefore unclear what effect constraint of the head has on the degree of skull flacture produced by a blunt impact. Investigation of this issue is also necessary to generate a more robust model of pediatric skull flacture. In addition to animal models, Finite Element Analysis (FEA) is a powerful and commonly used tool in biomechanics studies. While flacture simulations have extremely high computational costs, it has been noted in the literature that principal stress or strain directions found in finite element simulations correlate well with experimentally and clinically observed flactures (Bozic et al., 1994; Silva et al., 1998). Frank and Lawn (1967) propose that flacture propagates perpendicular to the direction of the greatest principal tensile stress as a method of dissipating the maximum amount of energy in a system. Thus, it is assumed that the maximum principal stress directions may be used as a predictor of flacture in a finite element simulation. This thesis presents research conducted on an in situ porcine head model, as well as several finite element simulations to assess the effect of impact energy and impact scenario on the degree and pattern of skull flacture. In Chapter 2, flacture length due to a high level of blunt impact energy is documented and compared to previous impacts by Baumer et al. (In press) at lower impact energy levels. Also, the patterns of flacture for the two studies are directly compared in the current study. This study hypothesized that flacture length would be a function of impact energy, specimen age, and impact interface, and that the pattern of flacture would change with a higher level of impact energy. In Chapter 3, a flee fall drop tower was built and used to assess the degree of skull flacture and the flacture pattern for a flee falling head onto a rigid interface. The flee fall flacture patterns were then compared to those in Chapter 2. It was hypothesized that the flacture patterns for these two impact scenarios would be similar. In Chapter 4, the total flacture length, peak impact force, and impact duration data flom Chapters 2 and 3 were compared and used in conjuncture with a finite element model to investigate the impact responses of a rigid mass striking a constrained skull and a flee falling head onto a rigid surface. The maximum principal stress magnitudes and directions were compared and used to discuss the experimental flacture patterns. The research conducted for this thesis provides insight into the flacture patterns of the developing porcine skull flom two different impact variables: impact energy and impact constraint. A finite element model was also used to validate the experimentally observed differences in the degree of flacturing. These data may ultimately aid in being able to determine whether the head in a forensic case was constrained or flee to move during impact. This information may ultimately prove to be extremely beneficial in helping forensic pathologists and medical examiners to determine whether a trauma was due to accident or abuse. REFERENCES Barlow B, Niemirska M, Gandhi RP, Leblanc W, 1983, “Ten years of experience with falls flom a height in children,” Journal of Pediatric Surgery, 18(4), pp. 509-511. Baumer TG, Passalacqua NV, Powell BJ, Newberry WN, Fenton TW, Haut RC, In Press, “Age-Dependent Fracture Characteristics of Rigid and Compliant Surface Impacts on the Infant Skull — A Porcine Model,” Journal of Forensic Science. Baumer TG, Powell BJ, Fenton TW, Haut RC, 2009, “Age Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human,” Journal of Biomechanical Engineering, 131(11), pp. 111006-1-6. Belechri M, Petridou E, Trichopoulos D, 2002, “Bunk versus conventional beds: a comparative assessment of fall injury risk,” Journal of Epidemiology and Community Health, 56(6), pp. 413-417. 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(2), pp. 122-4. Bertocci G and Pierce MC, 2006, “Applications of Biomechanics Aiding in the Diagnosis of Child Abuse,” Clinical Pediatric Emergency Medicine, 7(3), pp. 194- 199. Billmire ME and Myers PA, 1985, “Serious head injury in infants: accident or abuse?” Pediatrics, 75, pp. 340-342. Bozic KJ, Keyak JH, Skinner HB, Bueff UH, Bradford DS, 1994, “Three-Dirnensional Finite Element Modeling of a Cervical Vertebra: An Investigation of Burst Fracture Mechanism,” Journal of Spinal Disorders & Techniques, 7(2), pp. 102-110. Chalmers DJ, Marshall SW, Langley JD, Evans MJ, Brunton CR, Kelly A, Pickering AF, 1996, “Height and surfacing as risk factors for injury in falls flom playground equipment: a case-control study,” Injury Prevention, 2(2), pp. 98-104. Chason JL, Fernando 0U, Hodgson VR, Thomas LM, Gurdjian ES, 1966, “Experimental Brain Concussion: Morphologic Findings and a New Cytologic Hypothesis,” The Journal of Trauma, 6(6), pp. 767-779. Coats B and Margulies SS, 2006, “Material Properties of Human Infant Skull and Suture at High Rates,” Journal of Neurotrauma, 23(8), pp. 1222-1232. Ewing CL, Thomas DJ, Sances A, Larson SJ, eds., 1983, “Impact Injury of the Head and Spine,” lst Ed., Springfield, IL: Charles C Thomas. Frank F and Lawn B, 1967, “On the Theory of Hertzian Fracture,” Proceedings of the Royal Society of London, 299(1458), pp. 291-306. Gruskin KD and Schutzman SA, 1999, “Head Trauma in Children Younger Than 2 Years,” Archives of Pediatrics and Adolescent Medicine, 153, pp. 15-20. Hall JR, Reyes HM, Horvat M, Meller J L, Stein R, 1989, “The Mortality of Childhood Falls,” The Journal of Trauma, 29(9), pp. 1273-1275. Knight B, 1991, “Fatal child abuse,” In: Forensic Pathology. London: Edward Arnold, pp. 457-73. Macarthur C, Hu X, Wesson DE, Parkin PC, 2000, “Risk factors for severe injuries associated with falls flom playground equipment,” Accident Analysis and Prevention, 32(3), pp. 377-382. Margulies SS and Thibault KL, 2000, “Infant Skull and Suture Properties: Measurements and Implications for Mechanisms of Pediatric Brain Injury,” Journal of Biomechanical Engineering, 122(4), pp. 364-371. McPherson GK and Kriewall TJ, 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. Motherway JA, Verschueren P, Van der Perre G, Slotcn JV, Gilchrist MD, 2009, “The mechanical properties of cranial bone: The effect of loading rate and cranial sampling position,” Journal of Biomechanics, 42, pp. 2129-2135. Prange MT, Coats B, Duhaime A, Margulies SS, 2003, “Anthropomorphic simulations of falls, shakes, and inflicted impacts in infants,” Journal of Neurosurgery, 99(1), pp. 143-150. Prange MT, Luck JF, Dibb A, Van Ee CA, Nightingale RW, Myers BS, 2004, “Mechanical Properties and Anthropometry of the Human Infant Head,” Stapp Car Crash Journal, 48, pp. 279-99. Reece RM and Sege R, 2000, “Childhood Head Injuries: Accidental or Inflicted?” Archives of Pediatric and Adolescent Medicine, 154, pp. 11-15. Rosman NP, 2004, “Oski’s Essential Pediatrics,” In: M Crocetti and MA Barone, editors. Oski’s Essential Pediatrics. Philadelphia: Lippincott, Williams, and Wilkins. 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 MJ, Keaveny TM, Hayes WC, 1998, “Computed Tomography-Based Finite Element Analysis Predicts Failure Loads and Fracture Patterns for Vertebral Sections,” Journal of Orthopaedic Research, 16(3), pp. 300-8. Tabatabaei SM and Sedighi A, 2008, “Pediatric Head Injury,” Iranian Journal of Child Neurology, 2(2), pp. 7-13. Wheeler DS and Shope TR, 1997, “Depressed skull flacture in a 7-month old who fell flom bed,” Pediatrics, 100, pp. 1033-1034. Wilkins, B, 1997, “Head injury — abuse or accident?” Archives of Disease in Childhood, 76(5), pp. 393-397. Zimmerman RA and Bilaniuk LT, 1994, “Pediatric head trauma,” Neuroirnaging Clinics of North America, 4(2), pp. 349-366. CHAPTER TWO FRACTURE PATTERNS ON THE INFANT PORCINE SKULL FOLLOWING SEVERE BLUNT IMPACT ABSTRACT Traumatic injury to the head accounts for 80% of fatal child abuse cases. However, it is difficult to distinguish abuse flom accidental injury is difficult due to similarities in the type of injury. In this chapter, porcine specimens were impacted with a rigid mass at a high level of energy to assess characteristic flacture patterns related to impact interface, specimen age, and level of impact energy. A single impact was delivered to the right parietal bone of 57 specimens aged 2 to 28 days. The impact was generated by releasing a gravity accelerated mass flom a controlled height. Paired rigid and compliant impacts of equal energy were conducted for each specimen age. Impact force and contact area were recorded for each impact. Also, skull flacture length was measured to the nearest millimeter. Geographic Information Systems software was used to monitor the flequency of flacture initiation and propagation. Individual maps were generated based on impact interface, specimen age, and impact interface. It was found that impact force increased with age, regardless of interface. Contact area increased with age but was significantly higher in the compliant interface impacts than the rigid. Skull flacture length was greater for the rigid interface than the compliant at all ages except for 2 days. The degree of skull flacture also increased with an increase in impact energy. GIS maps documented additional sites of flacture initiation at a high level of impact energy dependant upon impact interface and level of impact energy. Several unique characteristic flacture patterns were noted as a function of specimen age, impact 10 interface, and impact energy. These characteristics may prove to be extremely beneficial in establishing the causation of trauma in child abuse cases where the validity of the testimony is questionable or unclear. ll INTRODUCTION Head injuries account for 80% of fatal child abuse in young children (Case et al., 2001). In a study of 89 children under the age of 2 years, 19 of the 20 fatalities were due to abuse (Hobbs, 1984). In contrast, short falls rarely cause serious injury or death in young children (Reiber, 1993). A study of 1,881 falls flom bed height reported no deaths (Belechri et al., 2002). Short falls (less than 3 feet), however, can still produce skull flacture in infants less than 12 months of age (Gruskin and Schutzrnan, 1999). There is currently a lack of data regarding pediatric skull flacture tolerance data and, thus, pediatric trauma with related cranial flacture due to a single-event represents one of the greatest challenges to forensic pathologists and anthropologists. Distinguishing between accidental and abusive trauma can be difficult, as both may produce similar types of injuries (Billmire and Myers, 1985). Specifically, linear, complex, and depressed skull flactures have been seen in both cases (Reece and Sege, 2000; Wheeler and Shope, 1997). The most commonly flactured cranial bone in accidental and abuse cases is the parietal (Hobbs, 1984; Meservy et al., 1987; Leventhal et al., 1993). The risk of injury is also dependent on the contacting surface (Bertocci et al., 2003). Other variables, such as the area struck, thickness of the skull, thickness of the scalp and hair, and impact direction can also affect the pattern of skull flacture (Knight, 1991; Cooperrnan and Merten, 2001). To better distinguish pediatric abuse flom accident, the aforementioned variables and the magnitude in which they affect skull flacture must be better understood. A recent study by Baumer et al. (In press) assessed the effects of interface and age using an infant porcine skull impact model, looking specifically at the location of 12 flacture initiation on the parietal bone. The study shows that impacts flom Iow heights (low energy) typically initiate flactures at a bone-suture boundary. However, in many pediatric death cases there are multiple skull flactures which sometimes extend across suture boundaries (Meservy et a1, 1987; Stewart et al., 1993). Multiple, wide or cross- suture flactures are indicative of high energy trauma (Hobbs, 1984). It is therefore necessary to study the effect of high energy impacts as a function of age and interface on the flacture patterns for the pediatric skull. There were two hypotheses in the current study. First, that the locations of flacture initiation would not depend on the level of impact energy. Rather, it would only depend on age and interface. Secondly, that an increase in impact energy would increase the amount of flacture via propagation for all ages and for both rigid and compliant interfaces. These data will ultimately provide insight into the effect of impact energy on skull flacture patterns due to a single, blunt impact and may provide valuable information that may help distinguish pediatric abuse flom accident. MATERIALS AND METHODS Porcine specimens were received flom a local supplier and stored at -20°C. A total of 57 specimens (aged 2 to 28 days) were used for this study. The animals died of natural causes and were flozen within 12 hours of death. All specimens were flee of head injury, which was confirmed during preparation. The test procedure was described in a previous study (Baumer et al., in press). Briefly, the head was allowed to thaw at room temperature for 24 hours before the scalp and facial tissues on the left side were removed. The specimens were transversely and rotationally constrained in a bed of air-hardened epoxy (Fibre Strand, Martin Senour 13 Corp., Cleveland, OH). Phosphate-buffered saline (PBS) solution was applied regularly during the preparation. The specimen was placed in a four degree of fleedom fixture that allowed adjustments of the impact site (Figure 2.1). The skull was oriented such that the center of the right parietal bone was normal to the impact interface. Impact Pressure Head Film Epoxy Bed Positioning Fixture Figure 2.1. Orientation of the right parietal bone with rigid impact interface. The specimens were impacted using a gravity accelerated mass (GAM) (Figure 2.2). A single impact was delivered using an operational amplifier comparator circuit to monitor the impact force and energize an electromagnetic solenoid to catch the GAM immediately after the impact force returned to zero. The forces were recorded using a load transducer (4.45 kN capacity, model AL311CV, Sensotec, Columbus, OH) mounted immediately behind the impact interface. Two interfaces were used in this study: rigid and compliant. The rigid interface was a solid aluminum cylinder with approximately 16 cm2 of surface area (Figure 2.1). The compliant interface was a deformable aluminum block (1.10 MPa crush strength Hexcel, Hexcel Corp, Stamford, CT), approximately 3 cm thick with a 16 cm2 surface area, attached to the rigid interface. 14 Stainless Steel Shaft Vertically mounted trolley with adjustable mass (GAM) Impact Head Skull with parietal bone exposed to apparatus Positioning fixture Figure 2.2. The drop tower schematic with GAM shown. Impact energy was controlled by varying the drop height of a 1.67 kg GAM. A slightly larger, 1.92 kg mass, was used to generate flacture in specimens aged twenty- one days and older. The mass of the impact interface was included in the GAM. Energy levels for each age were double those of a previous study (Baumer et al., in press). The impact energy was doubled by raising the height of . the GAM to twice the height of the previous study. The input energy for the compliant and rigid interfaces was equal at each age, however, the impact energy was increased with specimen age. For cases in which the first impact did not cause flacture of the skull (n=5 in the current study), the skull was impacted a second time at a slightly higher energy. The force data was sampled at 10,000 Hz. Pressure sensitive film packets (Prescale, Fuji Film Ltd., Tokyo, Japan) were attached to the impact site of each specimen to capture contact area. Two sheets of polyethylene were used to protect low (0-10 MPa) and medium (10-50 MPa) range pressure films stacked on top of one other flom fluids (Atkinson et al., 1998). The 15 medium pressure film was not used in the current study as the impact pressures were too low to record on the film. Afler impact, the remaining periosteum and sofl tissues were removed flom the skull and it was visually inspected for bone flacture and suture damage. The remaining soft tissue on the skull was then removed using standard anthropological procedures. The length of skull flactures was measured to the nearest millimeter using a soft, flexible measuring tape, which contoured to the curvature of the skull. Complete flacture diagrams were constructed manually for each specimen. In order to compare the patterns of flacture between specimens and interfaces, a Geographic Information System (GIS) method was utilized in the study. The pattern of flacture flom each skull was constructed using a projected view of the porcine cranium which best highlighted the right side of the skull with flacture configurations superimposed on it for each specimen. A second view of the posterior aspect of the cranium was also included as many high energy flactures involved the occipital bone. Fracture data flom Baumer et. al (In press) was also revisited to compare low energy rigid and compliant flacture configurations to the current study. Porcine specimens were separated into two different age groups (2-9 and 19-28 days) for both the rigid and compliant impact interfaces and at low and high energy levels to better demonstrate flacture pattern changes in relation to porcine growth and development, impact interface, and input energy. These age groups were chosen based on general observations of gross flacture and material property changes for the skull and suture tissues documented in the literature (Baumer et al., in press; Baumer et al., 2009). The flacture pattern for each porcine cranium was traced into individual shape files (Marean et al., 2001). The GIS 16 model then counted overlaid flacture patterns on each cranium, generating a map of where flactures appeared most flequently. After each map was constructed, the GIS model was used to discuss the differences in flacture patterns between specimens of different age, impact energy, and interface. The impact data were analyzed for age effects using linear regression analyses. Comparisons between the interfaces were performed using a two-factor (age, interface) ANOVA. Statistically significant effects were reported for p<0.05. RESULTS Impact energies were doubled flom those used by Baumer et al. (In press) by doubling the drop height at each age. The drop heights ranged flom 0.2 m for a 2 day old specimen to 1.2 m for a 28 day old specimen. The values of impact energy ranged flom 3.1 J to 22.6 J, respectively. The impact force on the skulls increased with age for both interfaces, and there was little difference in the peak impact force (within 100 N at a given age) between the two interfaces (Figure 2.3). Linear regression analysis indicated a significant increase in impact force with age for the compliant (p<0.001) and the rigid interfaces (p=0.006). l7 H Rigid E] -- £1 Compliant O 2500 - 2000 - 1500 - 1000 ~ Impact Force (N) 500 ~ o T I I I I I 0 5 10 15 20 25 30 A99 (Days) Figure 2.3. Peak impact force versus age for both the rigid and compliant interfaces. The contact areas generated during impact were found to significantly increase (p=0.003) with age at a similar rate for both interfaces (Figure 2.4). A two-factor AN OVA (age, interface) for the contact area showed a significantly larger area of contact generated with the compliant than rigid interface. 1600 ~ HRigid D—UCompliant A1200 ~ m2 Contact Area (m Age (Days) Figure 2.4. Contact area as a function of age for both rigid and compliant interfaces. 18 The length of flacturing (in bone and along sutures) versus age plot showed, on average, a significantly larger (p=0.034) amount of flacturing for the rigid than the compliant interfaces (Figure 2.5). 200 — IRigid DCompliant A160 - E E 1 £120 - a: c: o .1 w .1 g 80 o E '1 404 O 7 I r r r I I 2 3 4 5 6 7 8 910111213141516171819202428 Age (Days) Figure 2.5. The average length of skull flactures as a function of age. The GIS flacture maps confirmed that the length of flactures was greater for rigid than compliant interface impacts for the younger age group in the current study (Figure 2.6a and 2.6b). For the compliant interface experiments the pattern maps showed flactures primarily appearing to initiate at 4 sites adjacent to sutures along the perimeter of the parietal bone. However, for the rigid, there appeared to be numerous initiation sites. There was also significantly more diastatic flacturing in the rigid than compliant interface experiments, specifically along the coronal suture. 19 Frequency Figure 2.6. GIS map of 2-9 day old rigid (a) and compliant (b) impacts at high energy. In the older goup of specimens (19-28 days) more flacturing was again confirmed for the rigid than compliant interface experiments (Figure 2.7a and 2.7b). Yet, in these experiments, no diastatic flactures were noted. Sites of flacture initiation were evident in the parietal bone along the coronal and lambdoid sutures for the compliant interface experiments. Two of these sites were similar to those documented in the younger age goup. These sites were also noted in the rigid interface impacts, however there were more propagated flactures with the rigid interface. Interestingly, in the current study using high energy impacts to the parietal bone, significant flacturing was also documented in the occipital bone for botln age goups and interfaces. 20 Frequency Figure 2.7. GIS map of the 19-28 day old rigid (a) and compliant (b) impacts at high energy. The GIS maps of the revisited Baumer et al. (In press) data showed three primary areas of flacture irnitiation, regardless of interface. For the younger age goup (2-9 days old), the compliant interface produced more flacture of the skull than the rigid at the same impact energy level (Figure 2.8a and 2.8b). 21 Frequency Figure 2.8. GIS map of the 2-9 day old rigid (a) and compliant (b) impacts at low energy. For a low energy of impact there was little to no skull flacture with the compliant interface for the older age goup (Figure 2%). Two flacture initiation sites were noted along the coronal and lambdoid sutures. The rigid impacts produced more propagated flactures initiating at approximately the same locations as in the compliant interface experiments (Figure 2.9a). 22 Frequency Figure 2.9. GIS map of the 19-28 day old rigid (a) and compliant (b) impacts at low energy. There were many specimens in botln high energy age goups where flactures appeared in the occipital region. These flactures were not present in the revisited Baumer et al. (In press) data. DISCUSSION The current study focused on skull flacture patterns under a high impact energy as a function of both age and interface using a porcine model. It was hypothesized that a high level of impact energy would not change the locations of flacture irnitiation flom those documented previously in low energy experiments (Baumer et al., in press). However, the locations of flacture initiation were found to be a function of impact energy. Baumer et al. (In press) document three primary sites of flacture initiation for botln interfaces at a low impact energy. While these sites were present in the current study, new initiation sites emerged at the increased level of impact energy. It was also hypothesized in the current study that there would be a geater amount of flactuning via 23 propagation for these higher energy impacts. The amount of flacture produced in the current study was significantly geater than that produced in the low energy experiments conducted by Baumer et al. (In press). An interesting finding in the Baumer et al. (In press) study was the equal amount of flacture produced by the rigid and compliant interfaces at approximately 18 days of age for a given impact energy. Prior to 18 days, the compliant interface produced more flacture than the rigid interface, but thereafter there was less for the same impact energy. Baumer et al. (In press) suggest that the compliant interface generated higher states of stress near sutures. These stresses were high enough to produce diastatic flactures and therefore a larger amount of total flacture with the compliant interface. In the current study, however, the rigid interface produced more flacture than the compliant interface at all ages, except for the 2 day old specimens. This change may be attributable to alterations in bone and suture sensitivities to rate of loading. It has been shown in the literature that human bone and suture exhibit mechanical property sensitivities to loading rate (Wood, 1971; Yoganandan et al., 1995; Motherway et al., 2009). Young soft tissues, in particular, are more sensitive to changes in loading rate than older tissues (Haut, 1983). Margulies and Thibault (2000) also determined the material properties of young porcine crarnial suture at two different rates of loading and document significant increases in rupture modulus, elastic modulus and rupture energy at the higher rate. Therefore, at higher rates of loading, sutures may become more “brittle-like”, especially in the younger aged specimens. The rigid interface may have produced more flacture due its smaller contact area, which produced higher impact stresses than the compliant interface. Furthermore, with an increase in suture stiffness for these high rates of loading 24 in the current study higher stresses were likely transmitted across sutures to produce flacture in the occipital bone with botln interfaces. These results contrast with the lack of occipital flacture documented in the Baumer et al. (In press) study at low levels of impact energy. These findings suggest that high levels of impact energy generate more areas of flacture remote to the impact site. The geater degee of total skull flacture in the current study was also due to new sites of flacture initiation with increased levels of impact energy. Baumer et al. (In press) document three primary sites of flacture initiation, regardless of interface (Figure 2.8a and 2.8b) at low energy. These three sites were also documented in the current study, however, there were one or more additional sites of flacture irnitiation depending on interface. One possible explanation for these new sites is the need to dissipate a larger amount of energy through flacture of the bone. In the current study, the initiation sites documented by Baumer et al. (In press) were fully propagated through the parietal bone for the rigid interface. Fracture propagation appeared to be one method to dissipate impact energy, however, additional flactures sites were likely needed to allow the bone to dissipate all of the impact energy generated with the higher drop heights in the current study. These new sites were more flequent with the rigid interface due to the impact force being distributed over a smaller area, generating larger impact stresses in the skull. The larger contact area produced with the compliant interface likely attenuated impact stresses resulting in fewer flacture initiation sites. This difference in flacture initiation sites suggests that the level of impact energy affected a characteristic feature of the flacture pattern for a given interface. 25 The commonality in flacture patterns for the current and Baumer et al. (In press) studies between specimens of the same age and impacted with the same interface was documented using GIS software. This GIS image-analysis approach has been previously used for both archaeological cut-mark distribution on fauna (Abe et al., 2002) and carnivore modification to fauna] remains (Hodgson et al., 2009), however, while Darnann et al. (2009) have examined flacture patterns flom human aircraft crashes, this is the first forensic application of the Marean et al. (2001) GIS image-analysis technique for bone flacture pattern analysis. Furtlner, this project is unique as it is one of the first attempts to compile flacture pattern data flom a relatively large sample of documented experimental impacts. Lee (1992) examined flacture patterns of human laryngeal structures after impacting tlnem with a drop-tower, but this analysis did not employ GIS, instead using simple tracings of flacture lines. The analyses in the current study provided some insight into the discrimination of flacture characteristics as a fimction of specimen age, impact interface and energy. These characteristics were described by assessing the flequency of flacture on each GIS map for a given set of impact conditions. For example, high energy impacts in the younger age goup (2-9 days) tended to produce occipital flacture for both interfaces. Again, occipital flacture was not seen in the low energy impacts. Additionally, each interface produced characteristic flacture attributes. The rigid interface generated much diastatic flacturing at the higher impact energy, whereas the compliant interface did not. These findings contrast with those noted by Baumer et al. (In press) where the compliant interface produced more diastatic flactures than the rigid interface for the younger aged specimens. One could then say, if a given flacture pattern for a younger aged victim involves occipital and diastatic flacture, the 26 causation of injury could have been due to a high energy, rigid impact. Future work should focus on the study of uniqueness of the characteristic features of flacture patterns on the infant porcine skull as a function of impact interface, energy, and specimen age. While there are certain age limitations of the porcine model (after approximately 24 days the skull geometry begins to differ significantly flom humans), it could be used as an experimental model to help develop a better understanding of these characteristics. In cases of suspected child abuse, the medical examiner faces a difficult task in determining the causation of trauma. Age, interface and impact energy each appear to affect the pattern and degee of skull flacture. In the current study, rigid and compliant interfaces were compared at a given specimen age under a high impact energy. The rigid interface produced as much or more flacture as the compliant at each age. This was in contrast to previously reported data for low energy level impacts where a compliant interface produced more flacture of the infant porcine skull for specimens less than 18 days of age (Baumer et al., in press), and less for more aged specimens. The current study also showed that a higher versus lower levels of impact energy in previous studies altered the pattern of skull flacture by generating new sites of flacture irnitiation and causing flacture in adjacent bones of the skull. Some unique characteristics of these flacture patterns as a function of energy and interface were also assessed in the study using GIS software. Development of these characteristics may prove to be extremely beneficial in the task of investigating cases of potential abuse to infants. 27 REFERENCES Abe Y, Marean CW, Nilssen PJ, Assefa Z, Stone EC, 2002, “The analysis of cutrnarks on archaeofauna: A review and critique of quantification procedures, and a new image-analysis GIS approach,” American Antiquity, 67(4), pp. 643-663. Atkinson PJ, Newberry WN, Atkinson TS, Haut RC, 1998, “A method to increase the sensitive range of pressure sensitive film,” Journal of Biomechanics, 31(9), pp. 855- 859. Baumer TG, Passalacqua NV, Powell BJ, Newberry WN, Fenton TW, Haut RC, in press, “Age—Dependent Fracture Characteristics of Rigid and Compliant Surface Impacts on the Infant Skull — A Porcine Model,” Journal of Forensic Science. Baumer TG, Powell BJ, Fenton TW, Haut RC, 2009, “Age Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human,” Journal of Biomechanical Engineering, 131(11), pp. 1-6. Belechri M, Petridou E, Trichopoulos D, 2002, “Bunk versus conventional beds: a comparative assessment of fall injury risk,” Journal of Epidemiology and Community Health, 56, pp. 413-417. Bertocci GE, Pierce MC, Deemer E, Aguel F, J anosky J E, Vogeley E, 2003, “Using Test Dummy Experiments to Investigate Pediatric Injury Risk in Simulated Short- Distance Falls,” Archives of Pediatrics and Adolescent Medicine, 157(5), pp. 480- 486. Billrnire ME and Myers PA, 1985, “Serious head injury in infants: accident or abuse?” Pediatrics, 75, pp. 340-342. Case ME, Graham MA, Handy TC, Jentzen JM, Monteleone JA, 2001, “Position Paper on Fatal Abusive Head Injuries in Infants and Young Children,” The American Journal of Forensic Medicine and Pathology, 22, pp. 112-122. Cooperman DR and Merten DF, 2001, “Skeletal Manifestation of Child Abuse,” In: RM Reece & S Ludwig, editors. Child Abuse: Medical Diagnosis and Management. Philadelphia: Lippincott, Williams, and Wilkins. Darnann FE, Adler R, Benedix DC, Kontanis EJ, 2009, “Patterns of perimortern flacture flom military aircraft crashes,” Proceedings of the American Academy of Forensic Sciences; Washington DC. Gruskin KD and Schutzrnan SA, 1999, “Head Trauma in Children Younger Than 2 Years,” Archives of Pediatrics and Adolescent Medicine, 153, pp. 15-20. 28 Haut RC, 1983, “Age-Dependent Influence of Strain Rate on the Tensile Failure of Rat- Tail Tendon,” Journal of Biomecharnical Engineering, 105(3), pp. 296-299. Hobbs CJ, 1984, “Skull Fracture and the Diagnosis of Abuse”, Archives of Disease in Childhood, 59, pp. 246-252. Hodgson JA, Plummer TW, Forrest F, Bose R, Oliver J S, 2009, “A GIS-based approach to documenting large canid damage to bones,” Proceedings of the 78th annual meeting of the American Association of Physical Antlnropologists; Chicago, IL. Knight B, 1991 , “Fatal child abuse,” In: Forensic Pathology. London: Edward Arnold, pp. 457-73. Lee SY, 1992, “Experimental blunt injury to the larynx,” The Annals of Otology, Rhinology, and Laryngology, 101(3), pp. 270-274. Leventhal JM, Thomas SA, Rosenfield NS, Markowitz RI, 1993, “Fractures in young children: distinguishing child abuse flom unintentional injuries,” American Journal of Diseases of Children, 147, pp. 87-92. Marean CW, Abe Y, Nilssen PJ, Stone EC, 2001, “Estimating the Minimum Number of Skeletal Elements (MNE) in Zooarchaeology: A Review and a New Image-Analysis GIS Approach,” American Antiquity, 66(2), pp. 333-348. Margulies SS and Thibault KL, 2000, “Infant Skull and Suture Properties: Measurements and Implications for Mechanisms of Pediatric Brain Injury,” Journal of Biomecharnical Engineering, 122(4), pp. 364-371. Meservy CJ, Towbin R, McLaurin RL, Myers PA, Ball W, 1987, “Radiogaphic characteristics of skull flactures resulting flom child abuse,” American Journal of Roentgenology, 149, pp. 173-175. Motherway JA, Verschueren P, Van der Perre G, Sloten JV, Gilchrist MD, 2009, “The mechanical properties of crarnial bone: The effect of loading rate and crarnial sampling position,” Journal of Biomechanics, 42(13), pp. 2129-2135. Reece RM and Sege R, 2000, “Childhood Head Injuries: Accidental or Inflicted?” Archives of Pediatric and Adolescent Medicine, 154, pp. 11-15. ' Reiber GD, 1993, “Fatal falls in childhood: How Far Much Children Fall to Sustain Fatal Head Injury.” American Journal of Forensic Medicine and Pathology, 14, pp. 201-207. Stewart G, Meert K, Rosenberg N, 1993, “Trauma in infants less than three months of age,” Pediatric Emergency Care, 9(4), pp. 199-201. 29 Wheeler DS and Shope TR, 1997, “Depressed skull flacture in a 7-month old who fell flom bed,” Pediatrics, 100, pp. 1033-1034. Wood J, 1971, “Dynamic Response of Human Cranial Bone,” Journal of Biomechanics, 4(1), pp. 1-12. Yoganandan N, Pintar FA, Sances A, Walsh PR, Ewing CL, Thomas DJ, et al., 1995, “Biomecharnics of Skull Fracture,” Journal of Neurotrauma, 12(4), pp. 659-669. 30 CHAPTER THREE FRACTURE LENGTH AND PATTERN OF THE PORCINE HEAD MODEL IN FREE FALL ABSTRACT Head trauma is the leading cause of death in infants. Often, children sustain head trauma during normal activities and play which are deemed accidental and innocent. In the previous chapter, the degee and pattern of skull flacture was investigated after a porcine head model, entrapped in a bed of air-hardening epoxy, was struck with a blunt mass. This scenario, however, does not reflect the common injury occurrence flom accidental falls. In the current study the porcine head model was dr0pped in flee fall to assess the pattern of flacture generated flom an unconstrained fall flom a controlled height. A mounting plate and clamp was used to orient the right parietal bone normal to the rigid impact interface. A drop trolley translating along a vertical stairnless steel shaft attached to a drop tower provided motion in only the vertical direction. Drop heights were calculated flom previously used impact energy values in Chapter 2 in order to produce an equal amount of flee fall impact energy to the skull. The Geogaphic Information Systems method was again used to map the flequency of fractures in the parietal, flontal and occipital bones. Total flacture length steadily decreased with age in the flee fall head impacts. The results also showed that the pattern of flacture differed significantly between the flee fall impacts and entrapped impacts regardless of age. The extensive levels of occipital bone flacture documented in Chapter 2 was not seen in the flee fall impacts, even though the same impact energy was used. Additionally, there were significantly fewer sites of flacture initiation in the flee fall impacts than in the 31 entrapped impacts. These sites were age dependent with younger specimens having flacture initiation along the coronal suture and older specimens having flacture initiation sites on botln the posterior and anterior edges of the parietal bone along the coronal and larndoid sutures. Diastatic flacture was documented in both age goups but not in specimens older than 14 days of age. The patterns of flacture presented here are significantly different and are representative of two cases of equal high level of energy head impacts. One case is a fall with a given energy onto a rigid surface and the second case is a head that is on a rigid surface being struck with a blunt object. The research presented in this chapter may be extremely beneficial to medical examiners and forensic pathologists in determining the cause of skull flacture in cases of potential pediatric abuse. 32 INTRODUCTION Traumatic injury to the head is the leading cause of death in infant humans (Tabatabaei and Sedighi, 2008). Often children sustain head trauma due to accidental falls during the development of walking motor skills or flom child support devices such as highchairs (Zimmerman and Bilaniuk, 1994; Mayr et al., 1999). Pediatric falls typically result in impacts to the head due to the increased weight ratio of the head to the body as compared to adults (Y oganandan and Pintar, 2004; Snyder, 1977; Smith et al., 1975; Cory et al., 2001). Death resulting flom falls is the third leading cause of death in infants aged 1-4 years of age (Hall et al., 1989). However, in a study of 89 children under 2 years of age, 19 of the 20 fatalities were attributed to physical abuse (Hobbs, 1984). Linear, complex, and depressed flactures have been documented in both abuse and accidental cases (Reece and Sege, 2000; Wheeler and Shope, 1997). Thus, distinguishing an accident flom inflicted abuse is difficult due to both scenarios producing similar types of injury (Billmire and Myers, 1985). Due to a lack of pediatric cranial trauma data, correctly diagnosing skull flacture due to abuse or flom an accidental fall poses a significant challenge to medical examiners and forensic pathologists and anthropologists. Injury biomechanics are used in case-based investigations of suspected child abuse (Bertocci and Pierce, 2006). Animal models are often used in these cases to correlate data to humans. A porcine head model has been recently used in this laboratory to assess the effect of a blunt impact to the parietal bone of the skull. Baumer et al. (In press) document flacture initiation sites occurring at the bone-suture boundary, remote of the point of impact. Similarly, in Chapter 2, skull flacture patterns were assessed to 33 distinguish unique characteristics due to impact interface, age of specimen, and impact energy. In both studies, the porCine head was entrapped in a bed of epoxy and impacted with a rigid mass. Chason et al. (1966) impacted a carnine head model that was both fixed and flee using a rotary striker. Interestingly, the authors documented that there was a higher susceptibility to concussion when the head was fixed rather than flee to move. The authors also documented a lack of skull flacture in all of the specimens. It has been shown, however, that there is typically less force required to produce a concussion than a skull flacture (Rosman, 2004) and thus the energy levels used in the Chason et al. (1966) study may not have been high enough to produce skull flacture. Thus, it is unclear what effect constraint of the head during impact has on the degee and pattern of skull flacture. Therefore, in the current study, the porcine head model was dropped in flee fall to assess the degee of skull flacture length as well as the flacture pattern. The hypothesis of this study was that the pattern of flacture in the flee fall impact would be similar to those found in the entrapped impacts flom Chapter 2 for an equal impact energy. These data may ultimately provide utility for forensic pathologists and anthropologists in comparing flacture data flom potential pediatric abuse cases to known impact conditions using the porcine head model. If there are differences in fracture pattern between a constrained head impact versus a flee fall head drop, it may be possible to determine if the victim’s head was constrained or flee to move during the impact. MATERIALS AND METHODS A total of 31 porcine specimens were received flom a local supplier and stored at a temperature of -20°C. All animals had died of natural causes and were flozen within 34 12 hours of death. The specimens ranged flom 2-17 days of age. Each animal was inspected for initial head injury by palpating the parietal bone prior to experimentation. Each specimen was fastened to a mounting plate using Velcro straps. A four degee of flwdom clamp directly attached to the mounting plate was used to orient the parietal bone normal to the impact interface. The mounting plate was fastened to a hollow aluminum rod which was supported by a gravity accelerated drop trolley. The rod was clamped with an electromagnetic solenoid, acting as a catch and release mechanism, attached to the drop trolley. The drop trolley was constricted to linear translation in the vertical direction by a stairnless steel shaft attached to a drop tower chassis (Figure 3.1). To produce the necessary impact energy, the drop trolley was raised to the necessary drop height and held in place by a second electromagnetic solenoid. The solenoid was fastened to a crossbar to adjust the drop height to a maximum of 9 feet. During the experiment, both solenoids disengaged, allowing the drop trolley and mounting rod to fall fleely (Figure 3.2). Stainless Steel Shaft Vertically mounted drop trolley Mounting plate with porcine head Trolley Impact Padding Impact interface and load cell - ... _ ... . , ,.... _ V r I‘ ‘1 7.3.” v-7": r, ., .,, ,_ t if . . J‘“n ~.— +1 «ate: ra'ricrjmx-i? .. ‘z I .~;-r-_ _ I r‘ Figure 3.1. Schematic of the flee fall drop tower. 35 Upon impact, the trolley base struck a soft padded surface to dissipate the drop energy in the trolley. The skull impacted a rigid, aluminum interface with an approximate surface area of 324 cm2. A load cell (2.22 kN capacity, model IOIOAF-SOO, Interface, Scottsdale, AZ) mounted immediately behind the impact interface recorded impact force, duration, and energy. The skull was allowed to impact only once by using an operational amplifier comparator circuit to monitor the impact and reenergize the electromagretic solenoid to catch the rod immediately after the impact force returned to zero. The force data were sampled at 10,000 Hz. Solenoid clamps Solenoid clamps engaged disengaged a b Figure 3.2. The drop trolley was raised to the necessary drop height and held with the electromagnetic solenoid clamp (a). The flee fall impact was produced by disengaging the solenoid clamps (b). The mounting rod was flee to move vertically until the impact force returned to zero. In Chapter 2, entrapped porcine specimens were impacted using a high level of impact energy. Energy values for each age of specimen were matched in this chapter to those used in Chapter 2 by varying the drop height of each specimen. Due to the 36 differences in mass of the head with age, each specimen’s head mass was measured and given a respective drop height to match the Chapter 2 impact energies at each age using U = m * g * h (4.1) where U was the potential impact energy, m was the mass of the head, g was the gavitational acceleration, and h was the respective drop height. Each skull was dropped only once regardless if skull flacture was present. After impact, the scalp and soft tissues of the skull were removed using standard anthropological procedures. Each skull was visually inspected for flactures and photo documentation was taken. The periosteum and remaining soft tissues were then removed. The length of skull flacture was measured to the nearest millimeter using a soft, flexible measuring tape, which contoured to the curvature of the skull. Complete flacture diagarns were manually constructed for each specimen. The Geo gaphic Information System (GIS) method was again used in this chapter to map the flequency of flactures. The pattern of flacture flom each skull was constructed using a projected view of the porcine cranium which best highlighted the right side of the skull with superimposed flacture configurations for each specimen. A second view of the posterior aspect of the cranium was also included to display flactures of the occipital bone. The porcine specimens were again separated into two different age groups (2-9 and 10-17 days) to demonstrate flacture pattern changes in relation to porcine gowth and development, as well as the impact scenario. Fracture pattern data flom Chapter 2 were revisited to compare the flacture patterns of the 2-9 day old and 10- 17 day old age goups from high energy, rigid interface, entrapped impacts to the flee fall experiments. The age goups were chosen based on the available specimens, as well 37 as general observations of goss flacture and material property changes for the skull and suture tissues documented in the literature (Baumer et al., in press; Baumer et al., 2009). The pattern of flacture for each porcine cranium was traced into individual shape files (Marean et al., 2001). The GIS software then counted overlaid flacture patterns on each crarnium, generating a map of where flactures appeared most flequently. After each map was constructed, the GIS maps were used to compare the patterns of flacture produced flom flee falls and entrapped impacts using the same energy. Impact force and total flacture length were analyzed for age effects using linear regession analyses in a statistics software (SignaStat 2.03, Aspire Software International, Ashbum, VA). Statistically significant effects were reported for a p-value of less than 0.05. RESULTS The aim of the current study was to assess the impact characteristics of a porcine head dropped in a flee fall scenario. There was a characteristic rapid drop in force documented in the force-tirne plots, which was associated with skull flacture (Figure 3.3). A more pronounced peak was noted if no flacture was present. 38 1400 - 1200 - 1000 i Fracture 800 - Force (N) 600 - 400 - 200 - 0 r . I 0 0.001 0.002 0.003 Impact Duration (seconds) Figure 3.3. An overlay of force-time plots showing the characteristic rapid drop in force associated with skull flacture. Peak impact force significantly increased with age at a rate of 40.9 N/day of age (p<0.001) (Figure 3.4). Also, impact energy significantly increased with age at a rate of 0.19 J/day (p=0.009). 0 Impact Force - Impact Energy 1600 — —— 9 @1200 — 3 3 " 65 '5 ~ 2 I: 800 m 0 H 3 . . -- 3% _§ 400 - . a I E 0 I . I I 0 o 5 1o 15 20 Age (Days) Figure 3.4. Peak impact force with respect to age. 39 Total flacture length (bone and diastatic) was found to significantly decrease with age at a rate of -l.83 mm/day (p<0.001) (Figure 3.5). In 5 of the 31 specimens, no skull flacture (bone or diastatic) was documented. 80~ O) O r I Total Fracture Length (mm) B 8 O 10 15 20 Age (Days) 0 (1'1 Figure 3.5. Total flacture length with respect to age for flee fall and entrapped impacts. Average diastatic and bone flacture were compared at each age (Figure 3.6). Diastatic flacture was documented as early as 3 days of age. However, no diastatic flacture was noted in specimens older than 14 days of age. 0) O J I Bone El Diastatic .h o r N O r Average Fracture Length (mm) 2 3 4 5 6 7 8 91011121314151617 Age(Days) Figure 3.6. The average total flacture length versus age. 0 r 40 In the younger age goup (2-9 days old), extensive diastatic fracture was documented along the coronal suture in the flee fall impacts (Figure 3.73). Additionally, several bilateral (occurring on both sides of the suture) bone flactures extended across the coronal suture into the flontal and parietal bones. Bone flacture tended to initiate at the coronal suture. There were no documented initiation sites along the posterior or superior edges of the parietal bone. Also, there was no documented occipital flacture in the flee fall head impacts contrasting the extensive occipital flacture noted in Chapter 2 (Figure 3.7b). Frequency CEO-1 -2-2 -3-4 -5-6 Figure 3.7. GIS map of the 2-9 day old age goup for the flee fall (a) and entrapped (b) impacts. In the older age goup (10-17 days old), several flacture initiations were documented at the lamdoidal and squamosal suture intersection (Figure 3.83). However, flacture initiation was documented more flequently along the anterior parietal bone as in the younger age goup. In Chapter 2, the sites of flacture initiation were flequent and numerous along the perimeter of the parietal bones (Figure 3.8b). 41 Frequency Figure 3.8. GIS map of the 10-17 day old age goup for the flee fall (a) and entrapped (b) impacts. As with the young age goup, no occipital flacture was documented in the 10-17 day old age goup when dropped in flee fall. This again contrasted with the extensive occipital flacturing documented for the entrapped head impacts of Chapter 2. Diastatic flacture in the coronal suture was again recorded in the older age goup. However, the degee of diastatic flacture was significantly less in the older age goup than in the younger. Several of the older specimens had flacture irnitiation on either side of the parietal bone at locations remote flom the point of impact (Figure 3.9). This was also documented in Chapter 2 and in a previous study using entrapped porcine head rigid impacts (Baumer et al., in press). 42 Figure 3.9. Fracture initiation sites located remote of the point of impact (represented by the bull’s-eye). Shaded areas represent flactures and each mark represents 5 mm of flacture length. DISCUSSION The GIS maps generated in the current chapter did not validate the hypothesis that the pattern of skull flacture was similar between free fall and entrapped head impacts. In Chapter 2, significant occipital flacture was documented in both age groups (2-9 days old and 10-17 days old) for entrapped head impacts at the high level of impact energy. However, irn the current chapter the flacturing generated flom a flee fall head drop, with an equivalent level of impact energy, was confined to the parietal and frontal bones. Thus, it might be hypothesized that the presence of occipital flacture is attributable to an increased state of overall skull stress when the skull is rigidly constrained in the entrapped impacts. Further investigation into this claim will be conducted in Chapter 4. Additionally, diastatic flacture was localized to only the coronal suture for all ages in the flee fall impacts. In the entrapped impacts, diastatic was documented extensively in all sutures surrounding the parietal bone for young ages. For older aged specimens, diastatic flacture was documented only in the coronal and sagittal 43 sutures. These differences further validate that the pattern of flacture changed with different impact scenarios. In the 2-9 day old age goup, the flacture patterns generated flom a blunt impact to the entrapped head produced flequent and diffuse flacture initiation sites along the parietal bone boundaries. In the flee falls, however, the flacture initiation sites were localized along the coronal suture. However, the two sets of flacture patterns may be used in tandem to diagnose if the head was constrained during the impact by comparing the amount of fracture initiation sites of the victim to the data. The entrapped head fracture patterns provide a better indicator of skull flacture pattern generated flom an impact where the head may be resting against a rigid surface. An example of this may be when a victim’s head is pinned against a concrete wall and is struck with a blunt weapon such as a club. The flacture patterns generated flom the flee fall head impacts would be more representative of cases where the victim fell flom a highchair or countertop or flom a trip and fall incident during the development of walking motor skills. If in a case of potential child abuse, the victim has few initiation sites located along the coronal suture, flom the current study data, it can be reasoned that the trauma was due to some sort of flee falling head impact with a rigid surface. While abuse cannot be ruled out (the child may have been pushed over), it does rule out a blow to the head while the child’s head was constrained with an equal impact energy. In the current study, the pig model was dropped onto the right parietal bone for better comparison to the data in Chapter 2, which also involved impact trauma to the right parietal bone. The two studies were conducted to record differences in flacture patterns between the two impact scenarios. The literature has shown that the most 44 commonly flactured cranial bone in both accidental and abuse cases is the parietal (Hobbs, 1984; Meservy et al., 1987; Leventlnal et al., 1993) and thus it was reasonable that the targeted impact area of the skull in Chapter 2 and 3 should be the parietal. However, it has been documented in the literature that accidental falls in children typically result in vertex impacts to the head (Yoganandan and Pintar, 2004). This is attributed to the larger head to body mass ratio of children than adults (Snyder, 1977). It must be noted, however, that not all accidental head injuries occur in this fashion. In a study of 18 children with cranial trauma (Plunkett, 2001), several children fell onto the occipital region and also onto the parietal region in innocent playgound accidents. In order for the medical examiner to validate the testimony of the caregiver, the pattern of skull flacture must be known for all impact conditions, including impact locations. In this regard, the current study was limited to parietal bone impacts. In the flee fall impacts, flacturing was typically linear and remote flom the point of impact. The flactures in the young specimens were generally bilateral, spanning into both the flontal and parietal bones. However, in the older age goup, the flactures were unilateral and present in only the parietal bone. Hobbs (1984) documents that accidental injury typically generated narrow, linear flactures in the parietal bone. The data in this chapter supports this claim but it is unclear why there are bilateral flactures and unilateral flacture differences between the young aged and the older aged specimens. It was suggested in Chapter 2 that higher rates of loading cause young suture tissue to behave more “brittle-like”, allowing higher stress distributions into surrounding bones of the skull. This same suggestion can be made for the young aged fleely dropped specimens in the current chapter. The impact stresses were distributed into the flontal 45 bone causing flacture because of the suture’s ability to transmit stress at higher rates of loading. In the older aged specimens, the sutures have become more stiff and the skull behaves as a single, complete structure rather than separate entities of bone. This allows the impact stresses to be transmitted over the entire area of the skull reducing the potential of flacture. The results of this chapter document that a porcine model dropped in flee fall produced a different pattern of flacture, for a given energy, than an entrapped head impacted with a blunt object. Interestingly, there was a lack of occipital bone flacture in the flee fall cases, which was attributed in Chapter 2 to high levels of impact energy. It was also found that the site of fracture initiation was age-dependent with older specimens having flacture initiation on opposite sides of the parietal bone. Accidental falls are a common occurrence in child development and it is necessary to know the pattern of flacture generated flom a flee fall head impact. The information produced in this research could prove extremely valuable to forensic pathologists and medical examiners in distinguishing the cause of skull flacture in a potential child abuse case when the circumstances of injury are questionable or unknown. 46 REFERENCES Baumer TG, Passalacqua NV, Powell BJ, Newberry WN, Fenton TW, Haut RC, in Press, “Age-Dependent Fracture Characteristics of Rigid and Compliant Surface Impacts on the Infant Skull — A Porcine Model,” Journal of Forensic Science. Baumer TG, Powell BJ, F enton TW, Haut RC, 2009, “Age Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human,” Journal of Biomechanical Engineering, 131(11), pp. 1-6. Bertocci G and Pierce MC, 2006, “Applications of Biomechanics Aiding in the Diagnosis of Child Abuse,” Clinical Pediatric Emergency Medicine, 7(3), pp. 194- 199. Billmire ME and Myers PA, 1985, “Serious head injury in infants: accident or abuse?” Pediatrics, 75, pp. 340-342. Chason JL, Fernando OU, Hodgson VR, Thomas LM, Gurdjian ES, 1966, “Experimental Brain Concussion: Morphologic Findings and a New Cytologic Hypothesis,” The Journal of Trauma, 6(6), pp. 767-779. Cory CZ, Jones MD, James DS, Leadbeatter S, Nokes LDM, 2001, “The potential and limitations of utilizing head impact injury models to assess the likelihood of significant head injury in infants after a fall,” Forensic Science International, 123, pp. 89-106. Hall JR, Reyes HM, Horvat M, Meller JL, Stein R, 1989, “The Mortality of Childhood Falls,” The Journal of Trauma, 29(9), pp. 1273-1275. Hobbs CJ, 1984, “Skull Fracture and the Diagnosis of Abuse”, Archives of Disease in Childhood, 59, pp. 246-252. Leventhal JM, Thomas SA, Rosenfield NS, Markowitz RI, 1993, “Fractures in young children: distinguishing child abuse flom unintentional injuries,” American Journal of Diseases of Children, 147, pp. 87-92. Marean CW, Abe Y, Nilssen PJ, Stone EC, 2001, “Estimating the Minimum Number of Skeletal Elements (MNE) in Zooarchaeology: A Review and a New Image-Analysis GIS Approach,” American Antiquity, 66(2), pp. 333-348. Mayr JM, Seebacher U, Schimpl G, Fiala F, 1999, “Highchair accidents,” Acta Paediatrica, 88(3), pp. 319-322. Meservy CJ, Towbin R, McLaurin RL, Myers PA, Ball W, 1987, “Radiogaphic characteristics of skull flactures resulting flom child abuse,” American J oumal of Roentgenology, 149, pp. 173-175. 47 Plunkett J, 2001, “Fatal Pediatric Head Injuries Caused by Short-Distance Falls,” The American Journal of Forensic Medicine and Pathology, 22(1), pp. 1-12. Reece RM and Sege R, 2000, “Childhood Head Injuries: Accidental or Inflicted?” Archives of Pediatric and Adolescent Medicine, 154, pp. 11-15. Rosman NP, 2004, “Oski’s Essential Pediatrics,” In: M Crocetti and MA Barone, editors. Oski’s Essential Pediatrics. Philadelphia: Lippincott, Williams, and Wilkins. Smith MD, Burrington JD, Woolf AD, 1975, “Injuries in children in flee falls: an analysis of 66 cases,” Journal of Trauma, 15, pp. 987-991. Snyder RG, F oust DR, Bowman BM, 1977, “Study of impact tolerance through flee fall investigations,” III-IS, Washington DC. Tabatabaei SM and Sedighi A, 2008, “Pediatric Head Injury,” Iranian Journal of Child Neurology, 2(2), pp. 7-13. Wheeler DS and Shope TR, 1997, “Depressed skull flacture in a 7-month old who fell flom bed,” Pediatrics, 100, pp. 1033-1034. Yoganandan N and Pintar FA, 2004, “Biomechanics of temporo-parietal skull flacture,” Clinical Biomechanics, 19, pp. 225-239. Zimmerman RA and Bilaniuk LT, 1994, “Pediatric head trauma,” Neuroirnaging Clinics of North America, 4(2), pp. 349-366. 48 CHAPTER FOUR COMPARISON OF FREE FALL AND ENTRAPPED IMPACTS WITH FINITE ELEMENT ANALYSIS ABSTRACT Significant differences exist in the pattern and degee of flacture in head impacts onto a rigid surface and impacts to a constrained head with a rigid interface. In this chapter, flacture length, peak impact force, and impact duration data flom the previous two chapters were compared to assess the differences between the two scenarios. Also, finite element simulations were performed using a simplified skull geometry in order to further investigate the differences flom a theoretical viewpoint. Peak impact forces were found to be similar regardless of the impact scenario. However, significantly more flacture length was documented in the entrapped porcine head impacts than in the fleely dropped heads. Impact durations were significantly longer in the entrapped impacts leading to the assumption that a longer impact duration was the likely explanation for more propagated flactures. Several sources in the literature confirm that a longer duration impact increases the susceptibility of material flacture. The maximum principal tensile stresses were compared between the two impact scenarios using the finite element model. The results showed that the stresses in the entrapped scenario were distributed over a larger area of the skull and were geater in magnitude at locations remote flom the point of impact. The flee fall simulation produced maximum principal tensile stresses at or nearer to the point of impact. Overlays of the principal stresses onto an experimentally impacted entrapped porcine head showed a strong correlation of flacture pattern between the theoretical finite element model and the experimental data. 49 However, the flee fall principal stresses contrasted the experimental flacture pattern in which flacture initiated at the bone-suture boundary and not at the point of impact which the finite element results suggested. The model was limited, however, to a homogeneous material and simple geometry. A more complex geometry with bone-suture interfaces may provide a better insight into the flacture patterns in the fleely dropped porcine heads. 50 INTRODUCTION In the previous chapters, two different impact scenarios of the infant porcine skull were tested. In Chapter 2, an entrapped porcine head was impacted with a blunt, rigid mass. These impacts were conducted with a high level of impact energy and generated extensive propagated flactures in the parietal and adjacent bones of the skull. The impact scenario was representative of a constrained head that was impacted with a blunt object, most likely occurring in an inflicted situation. However, a majority of pediatric head trauma is due to accidental falls (Kim et al., 2000; Hall et al., 1989), and thus the porcine head model was tested in flee fall in Chapter 3. Significant differences in the flacture patterns were documented in the two studies. Since the level of impact energy in the two studies was the same, the variation in flacture patterns must have been attributed to head constraint. Often in biomechanics studies, Finite Element Analysis (F EA) is used to determine material stresses and strains in complex geometries or dynarrnic situations. Several studies have shown that finite element models can be used to correlate clinically or experimentally observed flacture patterns with the stress or strain distributions found flom finite element simulations (Silva et al., 1998; Doorly and Gilchrist, 2006; Baumer, 2009; Baumer et al., in press). Modeling flacture of a material typically requires a large computing power, but Frank and Lawn (1967) proposed a theory that principal stress directions could be used to predict the path of flacture propagation. They propose that flacture, as a means of dissipating the maximum amount of energy in a system, occurs perpendicular to the maximum principal stress direction. Therefore, it may be possible to 51 predict areas of a model that are more susceptible to flacture by investigating the principal stress directions. In this chapter, a comparison was made between the entrapped head and flee fall impact data flom the previous two chapters. Fracture length, impact duration, and peak impact force data were revisited flom Chapters 2 and 3. Also, finite element analysis was used to simulate the two impact scenarios using a simplified skull geometry. It is thought that by assessing the principal stress magnitude and directions found in the finite element analysis, the change in flacture pattern between entrapped and fleely dropped heads could be explained. MATERIALS AND METHODS In partnership with Wayne State University, a simplified skull geometry was meshed using the Hexmesher function in DEP Hexmorpher v4.0 (Detroit Engineered Products, Detroit, Michigan) (Figure 1). The diameter of the simplified skull geometry (80 millimeters) was based on the skull size of a 7 day old porcine specimen using CT scans flom specimens in Chapter 2. An element size of 0.5 mm was selected, yielding 36,525 solid elements. The outer faces of the hexahedral solids were duplicated to create an overlying skull shell layer with nodal connectivity. 52 [IIIIIIIIILILILTLLLIJ Figure 4.1. Simplified skull geometry used for finite element simulations. Rigid impactor positioned above the skull. Pre-processing was performed in Altair Hypermesh v9.0 (Altair Engineering, Inc., Troy, Michigan). The solid elements were assigned viscoelastic properties representative of brain tissue (p=l,040 kg/m3, K=211 MPa, (30:1.72 kPa, G1=0.51 kPa, decay constant = 20 ms) (Mao et al., 2006). The bulk modulus has limited effect on model predictions, but reducing it by a factor of 10 (flom 2.1 GPa) saves computational time (Kleiven, 2002). The brain was added in the model to include intracranial pressure and displacement effects developed by the brain during a blunt impact. This method has been used in previous studies to generate a better predictive model for head impacts (Horgan and Gilchrist, 2004). The skull shell was assigned elastic properties of E=5 GPa (Baumer et al., 2009), p=2,150 kg/m3, and u=0.28 (Margulies and Thibault, 2000) flom the literature. The total head mass was 359 g, which was the average mass of a 2-9 day old pig head (see Appendix B). A skull thickness of 2 mm obtained flom the CT scan data was assigned to the shell layer. An 80x80 millimeter rigid impactor with aluminum material properties (p=2700 kg/m3, E=70 GPa, u=0.35) was positioned 1.4 millimeters 53 above the sphere to allow initial translation before contact. The rigid impactor had a mass of 1.67 kg, the same mass as the impactor in Chapter 2. Automatic surface-to- surface contact in LS-DYNA (Livemore Software Technology Corporation, Livermore, California) with no penalty stiffness was used to define interaction between the impact interface and sphere. A simulation length of 4 milliseconds was set to show the entire impact duration for both scenarios. The simulations were solved on an Opteron cluster using 1 node (8 CPUs) in LS-DYNA 971r4. Post-processing software (LS PrePost 3.0, Livermore Software Technology Corporation, Livermore, California) was used to assess impact duration, maximum tensile and compressive stress magnitudes, and principal stress directions for each simulation. Two scenarios were simulated to compare the impact conditions in Chapter 2 and 3. In the first simulation, half of the shell nodes were constrained in the simplified skull to simulate the entrapped head. The constraint condition prevented the simplified skull flom translating during the impact. The rigid impactor was assigned an initial velocity of 2.8 m/s to represent a drop height of 40 centimeters. This drop height was used for 3-9 day old entrapped specimens in Chapter 2. In the second simulation, the rigid impactor was constrained to represent an infinite mass. The simplified skull displacement constraints were removed and were given an initial velocity condition of 6.1 m/s to represent a drop height of 1.87 meters. The drop heights were representative of equal impact energies with respect to the mass of the impactor and the head and a static gavitational acceleration (9.81 m/sz). Fracture length, impact force, and impact duration data were revisited flom Chapters 2 and 3 for comparison. Effects of the impact scenario were assessed for 54 statistical significance using statistics software (SignaStat 2.03, Aspire Software International, Ashbum, VA). Two-way ANOVA tests were used to compare the statistical data between the two scenarios. Statistically significant effects were reported for a p-value of less than 0.05. RESULTS Total flacture length, impact duration, and peak impact force data flom Chapters 2 and 3 were compared. There was significantly more total skull flacture at each age in the entrapped head impacts than in the flee fall experiments (p<0.001) (Figure 4.2). Comparing the linear regession slopes and intercepts, the slopes were not found to be significantly different (p=0.2106), however, the intercepts were statistically different (p<0.05). _ H Entrapped D - -EI Free Fall 200 a ' ' é1eo- - I 5 I U) I E l 5120- - ' ' . - J “...-”M e _ . :r '8 804 n 5 ' E n . '3 ' ll. 0 . . ' .‘3 401 "Ira-0 ----- a ........ 0 c1 0 3 O 0 O ....... D I- n . I 8"'°Q'1E--a.., El 0 a—C a H . 0 5 10 15 20 Age(Days) Figure 4.2. Total flacture (diastatic and bone) length with respect to specimen age for both impact scenarios. A two-way ANOVA analysis showed that the revisited peak impact force data flom Chapters 2 and 3 was not statistically different between the flee fall and entrapped impacts (Figure 4.3). 55 2500 _ H Entrapped E} - -Cl Free Fall g 2000 J 0 I I 2 ,2 1500 - ‘6 «I @1000 - 3 to g 500 - D O I I l I 0 5 10 1 5 20 Age (Days) Figure 4.3. Peak impact force with respect to age for both flee fall and entrapped impacts. Impact duration was defined as the total time taken until the peak impact forces began to decrease (Figure 4.4). 1000 '1 —Entrapped - Free Fall Impact Durations g 750 - N 0 g u. 500 - ‘6 (U Q g 250 l , O I I I f I I I “ITAI'AMTTT'TT 1.... I 0 1 2 3 4 5 6 7 8 9 10 11 Time (milliseconds) Figure 4.4. Force versus time plots used to determine the duration of impact for both entrapped and flee fall impact scenarios. 56 The impact duration was significantly shorter in the experimental flee fall impacts for a given age than the experimental entrapped impacts (p<0.001) (Figure 4.5). Linear regession analysis showed shorter impact durations with an increase in age for botln scenarios. The theoretical impact durations taken flom the finite element simulations were 1.1 ms and 4 ms for the flee fall and entrapped impacts, respectively. I—I Entrapped D - 4:! Free Fall 8- I 10‘ I 5.6- ' . ' . I g I '54- 5 n :- D e D I 0 ' 8 ........... r3 u I ' Q2- 0"D-n ..... E] """" Ifl""‘P--- B B D E 0 I: D U ....... — U 0 :1 :1 B D 0 D I I I I O 5 10 15 20 Age(Days) Figure 4.5. Impact duration with respect to age for the flee fall and entrapped impacts. The finite element simulations for the entrapped simplified geometry showed maximum tensile stresses in areas remote of the impact site (Figure 4.6). 57 Figure 4.6. Entrapped simulation showing impact site (bull’s-eye) and surrounding principal tensile stress directions. Four primary areas of maximum principal tensile stress were documented. Darker arrows represent higher magnitudes of tensile stress. These principal tensile directions and magnitudes were overlaid on a photo of an entrapped porcine skull to compare the distribution of stresses to the experimental flactures (Figure 4.7). Figure 4.7. Overlaid maximum tensile stress magnitudes and directions on a typical pattern of flactures flom an experimental entrapped porcine specimen. Shaded areas indicate documented flacture with each line representing 5 millimeters. 58 For the flee fall simulations, maximum principal stress magnitudes were located nearer the point of impact (Figure 4.8). At points further flom the point of impact, the tensile stresses dimirnished rapidly. I I Figure 4.8. Free fall simulation showing impact site (bull’s-eye) and surrounding principal tensile stress directions. The largest magnitude of tensile stress was located near the point of impact. The flee fall maximum principal stresses were also overlaid onto an experimentally impacted flee fall specimen to compare the observed patterns of flacture with the principal stress magnitudes and directions (Figure 4.9). Figure 4.9. Overlaid maximum principal tensile stresses on an experimental flee fall porcine skull. 59 Maximum pressures of the brain were an order of magnitude higher in the entrapped simulation than the flee fall pressures (5.41 MPa in the entrapped versus 500 kPa in the flee fall). DISCUSSION In this chapter, data flom the previous two chapters was compared to assess the differences between an entrapped head impacted with a mass and a fleely falling head impacting a rigid surface. The total flacture length for the flee fall impacts was significantly less than that for the entrapped impacts for an equal high level of impact energy. The larger flacture length in the entrapped scenario was attributed to flacture in the surrounding bones of the skull. In the flee fall scenario, the flactures were primarily located in the parietal, although some flacture occurred bilaterally in the flontal bone (Figure 3.7a). Peak impact forces were typically similar for a given age and impact energy, and thus the increased amount of flacture in the entrapped scenarios was not attributed to a higher force of impact. Impact duration, however, was significantly longer in the entrapped specimens than in the flee fall. Additionally, the maximum principal stress distributions in the entrapped impacts encompassed a much larger area whereas the flee fall stresses were localized around the point of impact. In the entrapped impacts, a mass was dropped onto a porcine head embedded in epoxy to prevent translation of the head. These impacts were significantly longer in duration than the flee fall impacts. In the literature, it has been shown that the density and size of cracks in a brittle solid is affected by the magnitude and duration of an applied dynamic elastic stress field (Evans et al., 1978). Seaman et al. (1976) suggest that propagation of flacture is related to the stress and strain duration. Davis (2000) 60 documents that the duration and magnitude of the impact forces will determine the extent of crarnial injury. Further, Gurdjian (1975) documents that higher and longer intracrarnial pressure changes can increase the susceptibility of the skull to flacture (Figure 4.10). The intracranial pressures found in the entrapped finite element model support this claim. The entrapped impact scenario produced ten times as much intracranial pressure as the flee fall. COMPARISON OF FIXED AND FREE HEAD DEGREE OF CONCUSSION MIN MOD SEVERE ks: REMARKS Higher and longer duration intracranial pressure changes more susceptible to fracture. j.‘ j /7 Figure 4.10. Gurdjian’s (1975) comparison of the degee of head trauma resulting flom a blunt impact to the head. The study suggests that longer duration impacts increase susceptibility of the skull to flacture. These claims seem reasonable to apply to the current study where there are significant differences in experimental impact duration between two different impact scenarios which generate dramatically different amounts of skull flacture length. The time durations recorded in the theoretical finite element simulations firrther imply that the process of skull flacture is a time-dependent phenomenon. Mechanically, a longer impact duration places the skull under an elevated stress field for an extended lengtln of time. Assuming the bone flactured at a flaction of the applied stress field, the initiation flacture would continue to propagate under the applied stress field irn order to maximize 61 the energy dissipation. This may explain why there was more propagated flacturing in the entrapped high energy impacts than the flee fall impacts where only flacture initiation seemed to have been documented. Further study of flacture models is necessary to verify these claims. The maximum principal stress magnitudes and directions correlated with the experimentally observed flacture patterns, especially for the entrapped scenario. F rank and Lawn (1967) claim that the geatest amount of energy dissipation occurs when the crack plane is perpendicular to the maximum principal stress direction. They also propose that flacture is the key mechanic in maximizing energy dissipation. In the entrapped scenario, the maximum tensile stresses were documented in areas surrounding the impact site. When overlaid onto a porcine skull, the maximum stresses were not only in the parietal bone but in the occipital bone as well. The principal stress directions were also perpendicular to the key flacture sites. In the flee fall simulations, the maximum tensile stresses were located primarily near the point of impact. Using the Frank and Lawn approach, this would suggest that the skull flactures should have been located near the point of impact or in the immediate surrounding area. However, the flee fall impacts typically generated small flacture initiation sites away flom the point of impact near the sutures surrounding the parietal bone. The mechanism of flacture is therefore somewhat uncertain. The simulation geometry is restricted in that it is a homogeneous bone layer with a brain solid interior. There were no sutures with appropriate material properties to affect the principal stress & strain distributions across the skull (Herring and Teng, 2000). It may be possible that the flacture initiation occurs at the bone-suture boundaries 62 due to stress and strain concentrations in the bone-suture interface (Baumer et al., in press; Yu et al., 2004; Alexandridis et al., 1985). In summary, the flacture length, impact duration and peak impact force data flom the previous chapters was compared. Fracture length was much geater in the entrapped impacts than the flee falls. This was attributed to the entrapped impacts having longer impact durations. Longer impact durations promote flacture propagation in an attempt to maximize the energy dissipation in the system. Also, a finite element software package was used to simulate the two impact scenarios using a simplified skull geometry to validate the increased amount of flacture length documented in the experimental entrapped skull impacts. It was found that an entrapped impact produced a larger distribution of maximum principal tensile stresses across the entire skull, whereas the flee fall impacts generated a more localized maximum principal stress pattern near the point of impact. The model, however, was limited in geometry and lacked interfacial changes between the bone and sutures. In order to better understand the mechanisms of remote flactures near the sutures in the experimentally observed flee fall scenarios, a more complex model of the immature skull is likely required. Furthermore, the current study suggests that mechanisms of bone and suture damage in the pediatric skull may be very dependent on temporal stress-strain effects. A purely elastic analysis may therefore not be appropriate. 63 REFERENCES Alexandridis C, Caputo AA, Thanos CE, 1985, “Distribution of stresses in the human skull,” Journal of Oral Rehabilitation, 12(6), pp. 499-507. Baumer TG, 2009, “Material Property Documentation and Fracture Analyses of the Developing Skull,” Masters Thesis Dissertation. East Lansing (MI): Michigan State University. Baumer TG, Nashelsky M, Hurst CV, Passalacqua NV, Fenton TW, Haut RC, in press, “Characteristics and Prediction of Cranial Crush Injuries in Children,” Journal of Forensic Science. Baumer TG, Powell BJ, Fenton TW, Haut RC, 2009, “Age Dependent Mechanical Properties of the Infant Porcine Parietal Bone and a Correlation to the Human,” Journal of Biomechanical Engineering, 131(11), pp. 1-6. Davis AE, 2000, “Mechanisms of Traumatic Brain Injury: Biomecharnical, Structural and Cellular Considerations,” Critical Care Nursing Quarterly, 23(3), pp. 1-13. Doorly MC and Gilchrist MD, 2006, “The use of accident reconstruction for the analysis of traumatic brain injury due to head impacts arising flom falls,” Computer Metlnods in Biomechanics and Biomedical Engineering, 9(6), pp. 371-377. Evans AG, Gulden ME, Rosenblatt M, 1978, “Impact damage in brittle materials in the elastic-plastic response regime,” Proceedings of the Royal Society of London, A361, pp. 343-365. Frank F and Lawn B, 1967, “On the Theory of Hertzian Fracture,” Proceedings of the Royal Society of London, 299(1458), pp. 291-306. Gurdjian ES, 1975, “Impact Head Injury,” Charles C. Thomas, Springfield, IL. Hall JR, Reyes HM, Horvat M, Meller JL, Stein R, 1989, “The Mortality of Childhood Falls,” The Journal of Trauma, 29(9), pp. 1273-1275. Herring SW and Teng S, 2000, “Strain in the braincase and its sutures during function,” American Journal of Physical Anthropology, 112(4), pp. 575-593. Horgan TJ and Gilchrist MD, 2004, “Influence of FE model variability in predicting brain motion and intracrarnial pressure changes in head impact simulations,” International Journal of Crashworthiness, 9(4), pp. 401-418. Kim KA, Wang MY, Griffith PM, Summers S, Levy ML, 2000, “Analysis of pediatric head injury flom falls,” Neurosurgical Focus, 8, Article 3. 64 Kleiven S, 2002, “Finite Element Modeling of the Human Head,” Doctoral Thesis Dissertation. Stockholm, Sweden: Royal Institute of Technology. Mao H, Zhang L, Yang K, King A, 2006, “Application of a finite element model of the brain to study traumatic brain injury mecharnisms in the rat,” Stapp Car Crash Journal, 50, pp.583-600. Margulies S and Thibault K, 2000, “Infant Skull and Suture Properties: Measurements and Implications for Mechanisms of Pediatric Brain Injury,” Journal of Biomechanical Engineering, 122(4), pp. 364-371. Seaman L, Curran DR, Shockey DA, 1976, “Computational models for ductile and brittle flacture,” Journal of Applied Physics, 47(11), pp. 4814-4826. Silva MJ, Keaveny TM, Hayes WC, 1998, “Computed Tomogaphy-Based Firnite Element Analysis Predicts Failure Loads and Fracture Patterns for Vertebral Sections,” Journal of Orthopaedic Research, 16(3), pp. 300-8. Yu JC, Borke JL, Zhang G, 2004, “Brief synopsis of cranial sutures: Optimization by adaptation,” Seminars in Pediatric Neurology, 11(4), pp. 249-255. 65 CHAPTER FIVE CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE WORK This thesis documented the flacture characteristics of the infant porcine skull as a function of specimen age, impact interface, level of impact energy, and impact scenario. The flacture patterns resulting flom a high energy blunt impact to the parietal bone were documented using GIS software. Porcine heads were dropped flom a controlled height onto a rigid interface to assess the differences in the pattern of skull flacture between two impact scenarios: flee fall and entrapped head impacts. Finally, a simplified skull geometry was modeled in a finite element software package to investigate potential explanations for differences in the flacture lengths and distributions of flacture documented in the entrapped and flee fall head experiments. In Chapter 2, high energy blunt impacts of infant porcine skulls were conducted in order to study characteristic patterns of propagated skull flacture. It was found that new flacture initiation sites were generated in these higher energy impacts that were dependent on impact interface. A rigid interface produced numerous new flacture irnitiation sites, while the compliant interface produced only one. Additionally, the rigid interface produced more skull flacture than the compliant at all ages except for 2 days for an equal amount of impact energy. GIS was used to document the flequency of characteristic flactures sites generated for both interfaces and at high and low (revisited data flom Baumer et al, in press) impact energy levels. Several urnique characteristic flactures were documented for each interface, age and energy level impact. One example of these characteristic flactures was a rigid mass striking a constrained head with a high level of impact energy produced a pattern of skull flacture including parietal, occipital 66 and diastatic flactures. This combination and location of flacturing was unique to only the high energy, rigid impacted specimens and was tlnus a characteristic flacture pattern for those impact conditions. If given a similar pattern in a case of potential child abuse, it may help in diagnosing whether the trauma was accidental or inflicted. Further studies should focus on continued development of these characteristic flactures using different ages, interfaces and impact energy levels. A comparative database describing flacture characteristics based on the input conditions may be used to eliminate or validate causes of trauma in cases of potential abuse. In Chapter 3, porcine heads were dropped onto a rigid mass using a gavity accelerated drop trolley. It was found that the pattern of flacture was different than that of the entrapped head impacts for an equal level of impact energy. Fractures in the flee fall head drops were typically located in the parietal bone, however in younger aged specimens the flactures extended bilaterally into the flontal bone, producing flacture on either side of the coronal suture (Figure 3.7a). The suture is thought to transmit the impact stresses to surrounding bones by its inherent stiffening during high rates of loading which may help explain the bilateral flactures shown in the younger specimens. In the older age goup, the sutures are more developed and the impact stresses are likely distributed more uniformly across the entire skull, reducing the potential for skull flacture. Extensive damage at the coronal suture was documented regardless of age so it is reasonable that a large portion of the impact stresses were located near or were transmitted through the coronal suture. Occipital flacture was not documented in the flee fall impacts which significantly contrasted the high energy entrapped impact flacture patterns. These results suggest that the anterior parietal bone and coronal suture 67 boundary is the weakest section of the skull in tlnese impacts (Macdonald et al., 1988; Burstein and Frankel, 1971). Ultimately, the differences in flacture patterns documented in this chapter could help in future diagnoses of whether a victim’s head was constrained during impact. Future work should investigate the differences in skull flacture patterns due to different impact interfaces, impact energies, and impact locations using the flee fall porcine head model. A different impact energy or impact interface may produce a different pattern of skull flacture than documented in this study for a high energy, rigid, flee fall impact. A larger range of flacture patterrns may produce characteristic features of skull flacture depending on each set of input conditions. These characteristic features may prove to be extremely valuable for forensic pathologists and medical exanniners to compare to potential child abuse cases to better diagnose the causation of trauma. In Chapter 4, a simplified skull geometry was modeled using a finite software package. Two simulations were conducted to assess the differences in maximum principal tensile magnitude and directions between the impact scenarios presented in Chapters 2 and 3. Total flacture length, peak impact force, and impact duration data flom Chapters 2 and 3 were also compared. Total flacture length was found to be significantly geater in the entrapped impacts than in the flee fall. However, there was little difference in peak impact force for a given age, and thus the increased flacturing was not attributed to impact force. The impact duration was significantly geater in the entrapped scenarios tlnen the flee falls for a given age. The literature suggests that a longer impact duration under a dynamic stress field makes the skull more susceptible to flacture. Also, the maximum principal stresses were distributed over a larger area of the entrapped simplified skull in the finite element simulations than the fleely falling skull. 68 Using an overlay of these principal stress magnitudes and directions, it was noted that the experimental skull flactures compared well with the theoretical stress directions for the entrapped scenario, however the flee fall experimental flactures did not resemble the theoretical predictions. Further studies should use a more complex and realistic skull geometry taken flom a computed-tomogaphy (CT) scan to investigate principal stress differences due to the more complex geometry. The model should also include sutures to assess how the impact stresses are affected by a non-homogeneous material and interface conditions between the bone and suture. It has been shown that increased levels of strain exist at the sutures during impact which may be a key indicator of why flacture initiation occurs at the bone-suture boundary in the flee fall impacts (Baumer, 2009). This research was based on assessing the degee and pattern of flacture of an infant porcine head model impacted in two injury scenarios. It was found that specimen age as well as impact interface, energy, and scenario all contribute to the pattern and degee of skull flacture in a porcine head model. This work suggests that investigating characteristic patterns of flacturc may aid in producing a more robust predictive flacture model to help diagnose accidental trauma flom inflicted. Pattern recognition techniques currently used for fingerprints could potentially help quantify these characteristic flacture patterns for each set of impact conditions. With a database of characteristic flacture patterns for a wide range of impact conditions, a forensic pathologist or medical examiner may be able to find a causation of trauma based on a victim’s skull flacture pattern. 69 REFERENCES Baumer TG, 2009, “Material Property Documentation and Fracture Analyses of the Developing Skull,” Masters Thesis Dissertation. East Lansing (MI): Michigan State University. Baumer TG, Passalacqua NV, Powell BJ, Newberry WN, Fenton TW, Haut RC, in press, “Age-Dependent Fracture Characteristics of Rigid and Compliant Surface Impacts on the Infant Skull — A Porcine Model,” Journal of Forensic Science. Burstein AH and Frankel VH, 1971, “A standard test for laboratory animal bone,” Journal of Biomechanics, 4(2), pp. 155-158. Macdonald W, Skirving AP, Scull ER, 1988, “A device for producing experimental flactures,” Acta ortlnopaedica Scandinavica, 59(5), pp. 542-544. 70 APPENDIX A RAW DATA FROM CHAPTER 2 71 72 o F .8N ova _ Nmood ON. 3 _ no.2: 83 ON 3 N rowIE Snow 23 _ 28.0 o; _ 38$ 8.2 E 3 38:3 863 2.0m _ mvood om: _ 8.3? map on 3 voomIE chmEmn :36 9 can corona. EB 68:: - 02:89 6: See .cOnocEfiE BSanO mom? on 3 881.1 deN vmvd :86 E: New: 0 5? co me 805 2.8? mmmé 886 SN? vaNo 33 oo 2 82¢ omdNN mmmd Nmood one”: mode moor 0N NF 9 Ed woNNF 086 9.8.0 2.3 omdmo 89 on Z on w E EN: 3.2m mmood mva ovNE moor ov m 9.815 9.va NmN.m Rood 0N.m oodom mom: ov m 381?. 8.09 Nmné Nmood mvNF NN. F 5 moor ov m N F _n_ 5.2 ONm.m $86 NmNF :68 wow? ov w or :a $69 mwod $86 3.2. 3.39. we? O... m Somzi wad? NONB mvood mm. 3 omhnw moor ov w meow—... a 5.8? meme 9.8.0 no.2. wo.mmm moor ov w Noomri IcmmEme :36 2 can corona. :5 SEE - noncooo. Co: San 60:95sz .93 E00 meme ov w 581?. mNm. 5v momN 586 or .m N F .moo mom? 9... N 881?. mu. 5N Now.”c 9.8.0 vaF 56o? moor ow N 8315 ENNN 25m 88.0 8.3 moeNNF moor ov N moomzi vo. 5F vONd Nvood NN.o_. mN.mmo moor ov N 2. :a tee mmmb mvood me. 5 2.33 mom: ov m N Four-n. $.va Secs ovood Ede mmNNm moor ov m mN:n_ 8.3; mmud Nvood 8d mNdmN moor ov m mNza NN.m_.N mine mmood mNdF 5. SK moor ov v m—omIE mon {mm 3.8.0 8.3 FNdNN moor ov m 82d No.0: NmNN mmood Rd NNdmv vow? 0N N 89d .53 :5sz 3 3 Ass. 80.. .3: A2. neon a. news. 292.. was om< 558nm mmoctzw >925 95.3.... c2359 BEE. “a EoEoermE Seas: mcEen. no.0 .3895 BEBE: Emu Q85 :mE 88m See 3mm .~.< «Ssh. 73 3.3m moms 386 mm. 3 mo. r EN m5? 8.. N 531:. NméFN 5% m Sod Noe? 5.82. 32. ON? mN omomIE mm. 5m Se. 3 $86 8.3. $.89 33 on N NNomIE 3.6mm. can.» 38.0 mo. 3 3.3.: a SF on N oNomIE mNémF N _.m.m 2.86 8.2. Sam FNF m ..mp co ON mNomIE Iodemc =3...“ 0. can Stone. :5 “08:: - 3280.. Doc Emu 60:05sz .65an0 9.3 cm oN NNomI_n_ mN. FNN vmod 58.0 «6.9 N tmmNF 2.3 oo 3 38:5 VNNt. Vde oeood em. 2. madam a z: co 3 32a endoN ONNé Need 96 mmNNm meme om 5 BBQ 3N2. Ems 2.8.0 NN.3 v0.89 mom? on or 98.13 :63 ommd $.86 mm.N_. 8.8m mom? on or E :a m... :.m N 56 $86 3.5 mN.vmmN mom? on or m _.om_I_n_ A53 ...:sz S 3 Es. .23.. ca: .2. 83a. E 8e: E 22.. in... eu< 55.8% 32.55 >925 95.3“. coach—5 «can—E. «a EoEoanwS «owns: 9.5a". no.5 was—5.80 .m.< 035—. Table A.2. Thickness, flacture length, and contact area measurements for high energy rigid interface impacts. Age Average Bone Suture Total Contact Specimen (da ) Thickness Fracture Damage Damage Area ys (mm) (mm) (mm) (mm) (mm’) P1058 2 1.28 72 18 90 64.2 P1109 3 1.03 118 16 134 264.9 PIHE018 4 1.05 135 25 160 463.1 P1125 5 1.15 39 20 59 41.0 P1129 5 1.18 120 0 120 155.3 PIHE017 5 1.14 50 0 50 326.0 P1113 7 0.99 80 45 125 53.7 PIHE006 7 1.33 140 0 140 443.3 PIHE008 7 1.15 65 0 65 256.2 PIHE009 7 1.12 135 0 135 478.8 P1HE001 8 - 85 20 105 207.3 PIHE002 8 - 73 7 80 146.3 PIHE005 8 1.47 12 0 12 276.6 P1HE007 8 1 .43 45 0 45 451 .8 P1110 8 1.06 90 55 145 225.6 P1126 9 1.42 15 0 15 103.6 PIHE015 9 1.99 95 0 95 339.0 PIHE016 9 1.27 130 0 130 251.5 P1130 1 1 0.95 30 25 55 NO PF P1116 12 1.64 123 0 123 349.2 P1002 13 1.84 95 21 116 158.6 P1006 13 1.02 190 0 190 120.4 PIHE003 14 - 80 22 102 224.0 PIHE004 14 1.73 80 0 80 155.3 PIHE011 15 2.00 155 0 155 622.9 PIHE012 15 1.93 20 0 20 492.0 PIHE013 15 1.08 115 37 152 235.6 P1111 16 2.01 95 27 122 278.8 PIHE010 16 2.28 185 0 185 549.1 P1055 17 1.53 40 0 40 248.6 P1070 19 1.41 220 0 220 61.9 PIHE014 19 1.78 35 0 35 520.3 PIHE022 20 2.28 105 0 105 580.8 PIHE023 20 2.73 225 15 240 695.2 PIHE026 24 2.91 90 0 90 628.2 PIHE027 24 2.72 165 0 165 706.9 P1HE030 28 2.59 65 0 65 584.9 PIHE031 28 3.92 55 0 55 810.1 74 vde 08.3 88.0 ..wNN _ 5.39. 32. ON? N 381?. me. For Nmmd 030.0 00.3 _ 3462 m 5.. oNF wN NmomIE 2.1m? $93 38.0 8.: _ 3.32 33 ON N 38.15 .Iemm Emu :36 0. can Stone. :3 .895 - 80.80. 8.. Eco 628..ng .33 E00 32. oN VN mNomIE $69. Na. 3 $86 8.2. New: m 5.. ON VN wNomI_ d oN.mNN owmd 9.8.0 vNNv 86wa 39 cm 0N mNomIE 8N? mNNd Nmood 3.9 Nm.wNN_. 32. co 0N «NomIE mNde on: 2.86 86 Eva: m5? co 2. veomIE N53: ommd 9.8.0 8.3 m Sb: 89. ON me 882.. Ned? Ne mm 2.86 8.9 5. Km 82. ON ..F Nmom; a ohmw ommd mvood w _..m_. 0N. _.oo_. 32. cm or meow T: a 1emmEmu :33 9 can Stone. :5 Same: - 8282 So: Emu 62.2.59: .23 :80 wow? cm or FNomIE 56m comm 38.0 mm: 5N8 mom: 2. m 3815 NmN: NNN.m 38.0 3.3 3.3.0 moor ow N NvomIE m 3m Name mNood N v.3 no. 5N moor ov o mmom I_ n. mm. NF 08.. N Sod 86 woNoN moor o... m 381E NN.mw momé 88.0 no.3 wodmm meme 9. m 88.15 8.8? 056 wNood N3; odeN 83 o... v 2.815 863 N94. 9.86 3.9. oodmN 89. ow m m Pom—IE :5. 2:25 3 3 2:5 28.. as: .2. 83. a. can: 292.. .233 oa< 55.8% «caesium >925 22.5.. c2350 32:... «a 22.53320 82:... 9...?“— .35 .Soemfim coerce: Sewage—.8 3.28 .33 Bob 8% 26M .m.< 033-. 75 Table A.4. Thickness, fracture length, and contact area measurements for high energy compliant interface impacts. Age Average Bone Suture Total Contact Specimen (da ) Thickness Fracture Damage Damage Area ’3 (mm) (mm) (mm) (mm) (mm’) PIHE019 3 1.18 90 10 100 756.2 PIHE040 4 1 .46 45 20 65 464.5 PIHE020 5 1 .08 55 15 70 796.3 PIHE041 5 1.23 95 20 115 842.7 PIHE039 6 1.31 30 25 55 472.9 PIHE042 7 1 .02 105 15 120 724.3 PIHE038 8 1.64 45 0 45 621 .9 PIHE021 10 1.66 175 0 175 480.3 PIHE043 10 1.58 100 0 100 906.1 PIHE037 11 2.42 5 0 5 844.0 PIHE036 13 2.31 10 0 10 580.5 PIHE044 18 1.38 145 25 170 995.2 PIHE024 20 2.23 100 0 100 824.6 PIHE025 20 2.37 35 0 35 918.9 PIHE028 24 2.18 35 0 35 687.5 PIHE029 24 2.27 115 0 115 807.2 PIHE034 24 2.51 0 0 0 1133.1 PIHE032 28 2.83 70 5 75 1311.6 PIHE033 28 2.62 0 0 0 1394.9 76 APPENDIX B RAW DATA FROM CHAPTER 3 77 888 88.8 88.8 8.8 8.88 88 8.82 3 3888 8.88 «88 8 188 8.8 8.83 88 8.88 3 3888 8.82 88.8 88.8 8. : 8.88 88 8 : 8 m 88 S 8.02 88.8 888 8.8 2.88 88 8.8: m? 28.88 2.88 888 S88 8.8 8.8: 88 3.2 8 88. 8 8. SF 888 888 8.8 8.88 88 8.3: 8 8858 88: 82. :88 8.8 8.88 88 N82 : $888 8.3: 82.8 8 88 8.8 8.88 88 8.82 : $8.88 :8: 88.8 :88 8.2 :88 88 8.82 2 28:8 8.8 88.8 888 8.: :88 88 8.82 S 8 E88 8.88 88.8 288 8.8 8.88 88 8.8: m 88.88 8.8 88.8 288 8.2 2.88 88 8.8: m 8888 8.88 88.8 28.8 8.8 8.88 88 83 8 888.8 888 88.8 888 8.8 8.8 E 88 39 m 88: 8 8.8 83 888 8.3 8.88 88 8.8? 8 88:8 8.8: 888 8.8.8 8.8 8.8: 88 8.88 8 8888 8.8: ~88 88.8 8.8 88 E 88 :2 o 888 8 8.8. 888 288 8.8 8.88 88 :2 8 88:8 8.8 88 88.8 8. : .8. 8 88 8.82 8 8888 8.82 88.8 888 8.2 8. 8 88 8.89 m 88.88 8.8 8: 88.8 8.8 8.8 8 88 m. 8 m 88_ 8 8.88 88 88.8 8.8 8.88 88 we: 8 88.88 888 88 88.8 8. 3 8.88 88 N8: 8 8888 8.88 88.8 :88 8.8 8.88 88 3.: m «888 88 F F 88.8 888 8.8 8E 88 8.8: m 88. 8 8.82 88.8 888 8.8 8.88 88 8.8 N 88.88 .82.: S 3 8:5 .83 .8: 22. 88". a. a8: .53 883 222.. 58325 30525 >Eocw 95:3. acacia «03...: «a EeEeowEeE 33E. new... no.5 o2 .888 8888 28 :8 88 88 88:8 88 38 .3 28.8 78 m ... BF 88.0 omood om. Pm "odour mac 8. EL 5 n {ammo coda «82¢ mmood 5.8 8.3:. on» 5va or 0 Kim 3 mad? Rad ..wood 56w 8.03: on» ~89 S. 0 Kim 3 wmdmv 89m wooed o P. S. 8.89 com 88: 3 m FEV 8 mod? «mod owood mm. E 3.38. com 9va 2 m Fin 5 Eu .22... .3 ... .52. .23.. z... .z. 83.. a. .3: . . E3. 222.. :25“.on 23525 .535 95:3. 20.3.59 «03:: «a 22.33.33 «can»... .52.. no.5 02 62.5880 .—.m flank. 79 Table 8.2. Fracture length and contact area measurements for free fall high energy rigid interface impacts. Age Bone Suture Total Contact Specimen (da ) Fracture Damage Damage Area ’3 (mm) (mm) (mm) (m_m’) OO3FF02 2 10 0 10 13.828 021 FF03 3 40 20 60 44.004 022FF03 3 17 15 32 190.561 024FF04 4 32 0 32 61 .469 025FF04 4 52 0 52 47.698 001 FF05 5 5 30 35 197.919 002FF05 . 5 65 10 75 58.892 005FF05 5 0 0 0 61 .240 017FF06 6 0 0 0 144.123 018FF06 6 55 10 65 129.408 026FF07 7 20 10 30 30.863 027FF07 7 15 20 35 98.058 011FF08 8 30 45 75 91.559 012FF08 8 10 15 25 167.027 004FF09 9 15 0 15 52.708 008FF09 9 15 0 15 153.543 006FF10 10 6 20 26 66.049 007FF10 10 0 0 0 144.925 028FF11 11 5 0 5 129.064 029FF11 11 30 0 30 188.786 030FF12 12 15 0 15 71.489 031FF12 12 10 0 10 77.874 009FF13 13 5 25 30 183.146 010FF13 13 15 10 25 112.144 019FF14 14 20 0 20 328.214 020FF14 14 17 22 39 260.619 013FF15 15 15 0 15 253.061 014FF15 15 0 0 0 205.134 015FF16 16 0 0 0 238.860 016FF16 16 15 0 15 133.072 023FF17 17 30 0 30 140.602 80 MICHIGAN STATE UNIVERSITY LIBRARIES ILIIIIIIIH 3 1293 (13 63 6800