368
THE ROLE OF APOLIPOPROTEIN E IN RECOVERY FROM TRAUMATIC BRAIN INJURY AND DEVELOPMENT OF CHIMERA: A NOVEL CLOSED-HEAD IMPACT MODEL OF ENGINEERED ROTATIONAL ACCELERATION by Dhananjay Rajaram Namjoshi M.Sc., The University of British Columbia, 2008 M.Pharm., The University of Mumbai, 2003 B.Pharm., The University of Mumbai, 1999 A THESIS SUBMITTED IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in The Faculty of Graduate and Postdoctoral Studies (Neuroscience) THE UNIVERSITY OF BRITISH COLUMBIA (Vancouver) February 2015 © Dhananjay Rajaram Namjoshi, 2015

the role of apolipoprotein e in recovery from traumatic brain

Embed Size (px)

Citation preview

THE ROLE OF APOLIPOPROTEIN E IN RECOVERY FROM TRAUMATIC BRAIN

INJURY AND DEVELOPMENT OF CHIMERA: A NOVEL CLOSED-HEAD IMPACT

MODEL OF ENGINEERED ROTATIONAL ACCELERATION

by

Dhananjay Rajaram Namjoshi

M.Sc., The University of British Columbia, 2008

M.Pharm., The University of Mumbai, 2003

B.Pharm., The University of Mumbai, 1999

A THESIS SUBMITTED IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

The Faculty of Graduate and Postdoctoral Studies

(Neuroscience)

THE UNIVERSITY OF BRITISH COLUMBIA

(Vancouver)

February 2015

© Dhananjay Rajaram Namjoshi, 2015

Abstract

Traumatic brain injury (TBI) is a “silent epidemic” that currently lacks any effective

treatment. While a major health care problem in itself, TBI also increases Alzheimer’s

disease (AD) risk and leads to the deposition of neurofibrillary tangles and amyloid

deposits similar to those found in AD. Agonists of Liver X receptors (LXRs), which

regulate the expression of many genes involved in lipid homeostasis and inflammation,

improve cognition and reduce neuropathology in AD mice. One pathway by which LXR

agonists exert their beneficial effects is through ATP-binding cassette transporter A1-

mediated lipid transport onto apolipoprotein E (apoE). In the first part of this thesis, I

show that a short-term treatment with synthetic LXR agonist GW3965 improves post-

injury outcomes in mice subjected to closed-head, mild, repetitive weight drop TBI

(mrTBI). My results suggest that both apoE-dependent and apoE-independent

pathways contribute to the ability of GW3965 to promote recovery from mrTBI. While

many drugs have shown promising outcomes in preclinical TBI models, clinical drug

trials for TBI so far have failed, suggesting that the translational potential of TBI models

may require further improvement. As most human TBIs result from impact to an intact

skull, closed head injury (CHI) rodent models are highly relevant. Traditional CHI

models like weight drop however suffer from large experimental variability that may be

due to poor control over biomechanical inputs. To address this caveat we developed a

novel CHI model called CHIMERA (Closed-Head Impact Model of Engineered

Rotational Acceleration) that fully integrates biomechanical, behavioral, and

neuropathological analyses. CHIMERA is distinct from existing neurotrauma model

systems in that it uses a completely non-surgical procedure to precisely deliver impacts

ii

of prescribed dynamic characteristics to a closed skull while enabling kinematic analysis

of unconstrained head movement. Here I show that repeated TBI in mice using

CHIMERA mimics many features of the human TBI including neurological, motor, and

cognitive deficits along with persistent neuroinflammation and diffuse axonal injury, and

increased endogenous tau phosphorylation up to 14 days with a reliable biomechanical

response of the head. This makes CHIMERA well suited to investigate the

pathophysiology of TBI and for drug development programs.

iii

Preface

Chapter 1.

Portions of the introductory text are reproduced from the published review article:

Namjoshi DR, Good C, Cheng WH, Panenka W, Richards D, Cripton PA, Wellington

CL (2013) Towards clinical management of traumatic brain injury: a review of models

and mechanisms from a biomechanical perspective. Disease Models & Mechanisms 6:

1325-1338.

Contributions:

Conception and outline of the article were designed by Dhananjay R. Namjoshi and

Cheryl L. Wellingtion. All authors contributed towards the preparation of the manuscript

with major contribution from Dhananjay R. Namjoshi.

Portions of the introductory text are reproduced from the published review article:

Namjoshi D, Stukas S, Wellington CL (2011) ABCA1, apoE and apoA-I as potential

therapeutic targets for treating Alzheimer’s disease. Neurodegenerative Disease

Management 1: 245-259.

Contributions:

Conception and outline of the article were designed by Dhananjay R. Namjoshi, Sophie

S. Stukas and Cheryl L. Wellingtion. Manuscript was prepared by Dhananjay R.

Namjoshi, Sophie Stukas and Cheryl L. Wellington with equal contribution from

Dhananjay Namjoshi and Sophie Stukas.

iv

Chapter 2.

A version is published in:

Namjoshi DR, Martin G, Donkin J, Wilkinson A, Stukas S, Fan J, Carr M, Tabarestani

S, Wuerth K, Hancock REW, Wellington CL (2013) The Liver X receptor agonist

GW3965 improves recovery from mild repetitive traumatic brain injury in mice partly

thorough apolipoprotein E. PLoS ONE 8(1): e53529.

All the animal experiments were carried out at UBC. All the biochemical and histology

experiments, except the cytokine assays described in this chapter were carried out in

the Wellington Laboratory at UBC. Cytokine assays were carried out by Kelli Wuerth in

the Hancock Laboratory at UBC.

Contributions:

All the experiments were designed by Dhananjay R Namjoshi and Cheryl L. Wellington.

James Donkin contributed towards the initial stages of experimental design. The

majority of experiments and data analyses were performed by Dhananjay R. Namjoshi.

Support for histological experiments and analysis was provided by Georgina Martin,

Michael Carr, and Sepediah Tabarestani and that for biochemical assays was provided

by Anna Wilkinson, Sophie Stukas, and Jianjia Fan. Manuscript was written by

Dhananjay R. Namjoshi and Cheryl L. Wellington.

v

Chapter 3.

A version is published in:

Namjoshi DR, Cheng WH, McInnes K, Martens KM, Carr M, Wilkinson A, Fan J,

Roberts J, Hayat A, Cripton PA, Wellington CL (2014) Merging Pathology with

Biomechanics using CHIMERA (Closed-Head Impact Model of Engineered Rotational

Acceleration): A Novel, Surgery-Free Model of Traumatic Brain Injury. Molecular

Neurodegeneration 9:55.

CHIMERA impactor was constructed by Kurt McInnes in the Cripton Laboratory at UBC.

All the animal experiments described in this Chapter were carried out at UBC. All the

biochemical and histology experiments described in this Chapter were carried out in the

Wellington Laboratory at UBC.

Contributions:

The CHIMERA impactor design was conceived by Dhananjay R. Namjoshi, Wai Hang

Cheng, Kurt McInnes, Peter A. Cripton, and Cheryl L. Wellington. Dhananjay R.

Namjoshi and Wai Hang Cheng designed and conducted experiments and analyzed

data. Support for head injury procedure was provided by Michael Carr and Kris M.

Martens. Behavioral assays were conducted by Dhananjay R. Namjoshi, Wai Hang

Cheng and Kris M. Martens. Head kinematics analysis was performed by Wai Hang

Cheng. Kurt McInnes and Peter A. Cripton provided expert advice on kinematic

analysis. Biochemical assays were carried out by Anna Wilkinson and Jianjia Fan.

Histology experiments were conducted by Dhananjay R. Namjoshi and Wai Hang

vi

Cheng. Support for histology was provided by Michael Carr, Jerome Roberts and Arooj

Hayat. Manuscript was prepared by Dhananjay R. Namjoshi, Wai Hang Cheng, and

Cheryl L. Wellington.

Study Approval by UBC Ethics Boards:

All the animal experiments described in this thesis were conducted according to the

protocols approved by the UBC Animal Care Committee as follows:

Protocol # A07-0706: ABCA1 and apoE function in traumatic brain injury.

Protocol # A11-0225: ABCA1 and apoE function in traumatic brain injury (New

Protocol).

The candidate’s animal care and training certification record:

3937-09: Online animal care training program

RBH-661-09: Rodent biology and husbandry

RA-343-09: Rodent anesthesia

RSHX-268-09: Rodent surgery

vii

Table of Contents

Abstract ............................................................................................................................ii

Preface ............................................................................................................................iv

Table of Contents .......................................................................................................... viii

List of Tables ..................................................................................................................xv

List of Figures ................................................................................................................ xvi

List of Abbreviations ...................................................................................................... xix

Acknowledgements ..................................................................................................... xxiii

Dedication .................................................................................................................... xxv

Chapter 1: Introduction .................................................................................................... 1

1.1 Traumatic Brain Injury (TBI): Definition ............................................................. 1

1.2 TBI: Epidemiology ............................................................................................. 4

1.3 TBI: Classification ............................................................................................. 7

1.3.1 TBI Classification Based on Clinical Severity .......................................... 7

1.3.2 TBI Classification Based on Pathophysiology ....................................... 10

1.3.3 TBI Classification Based on Physical Mechanism ................................. 11

1.4 TBI: Biomechanical Principles ........................................................................ 12

1.4.1 Head Motion during Impact TBI ............................................................. 15

1.4.2 Brain Motion during TBI ......................................................................... 17

1.4.3 The Clinical Picture: Human Tolerance and Related Biomechanical

Studies of Human TBI .................................................................................... 19

1.5 TBI: Pathophysiology ...................................................................................... 22

1.5.1 Primary Brain Injury ............................................................................... 23

viii

1.5.1.1 Focal Brain Injury ...................................................................... 24

1.5.1.1.1 Cerebral Contusion and Laceration .............................. 25

1.5.1.1.2 Intracranial Hemorrhage ............................................... 27

1.5.1.2 Diffuse Brain Injury .................................................................... 30

1.5.1.2.1 Diffuse Axonal Injury ..................................................... 30

1.5.1.2.2 Diffuse Vascular Injury .................................................. 34

1.5.2 Secondary Brain Injury .......................................................................... 34

1.5.2.1 Cerebrovascular Response to TBI ............................................ 35

1.5.2.2 Neurometabolic Changes Following TBI ................................... 36

1.5.2.3 Neuroinflammation Following TBI .............................................. 37

1.6 TBI, Alzheimer Disease, and Apolipoprotein E ............................................... 38

1.6.1 TBI and Alzheimer Disease ................................................................... 38

1.6.2 Alzheimer’s Disease: Pathology ............................................................ 39

1.6.2.1 Amyloid Beta Formation in the CNS .......................................... 40

1.6.2.2 Tauopathy in Alzheimer Disease ............................................... 42

1.6.3 TBI and Alzheimer’s Pathology: Similarities and Differences ................ 45

1.6.3.1 Amyloid Pathology after Moderate to Severe Single TBI ........... 45

1.6.3.2 Consequences of Repetitive TBI: Chronic Traumatic

Encephalopathy ..................................................................................... 50

1.6.4 Apolipoprotein E (ApoE) at the Nexus of TBI and AD Pathology ........... 55

1.6.4.1 ApoE Synthesis in the CNS ....................................................... 56

1.6.4.2 Liver X Receptors (LXR) ........................................................... 58

1.6.4.3 LXR: Molecular Mechanism of Action ........................................ 59

ix

1.6.4.4 LXR: Regulation of Cholesterol Metabolism .............................. 60

1.6.4.5 ABCA1: The Principal LXR Target ............................................ 62

1.6.4.6 The Role of LXRs in the CNS .................................................... 67

1.6.4.8 ApoE and Amyloid Beta (Aβ) Metabolism ................................. 75

1.6.4.8.1 ApoE and Aβ Interaction ............................................... 76

1.6.4.8.2 ApoE and Enzymatic Degradation of Aβ ....................... 77

1.6.4.8.3 Cellular Uptake and Degradation of Aβ ......................... 78

1.6.4.8.4 Aβ Clearance through the Cerebrovasculature ............. 79

1.6.4.9 ApoE, Synaptic Function, and Synaptogenesis ......................... 82

1.6.4.10 Role of ApoE in Inflammation .................................................. 84

1.6.5 ApoE and TBI ........................................................................................ 85

1.7 TBI: An Overview of Animal Models................................................................ 87

1.7.1 Open Head Injury Models ...................................................................... 89

1.7.1.1 Fluid Percussion (FP) ................................................................ 90

1.7.1.2 Controlled Cortical Impact (CCI) ................................................ 91

1.7.2 Closed Head Injury Models.................................................................... 92

1.7.2.1 CHI by Impact ............................................................................ 93

1.7.2.2 Non-Impact CHI by Blast Waves ............................................... 96

1.7.2.3 Non-Impact CHI by Inertial Loading .......................................... 97

1.7.3 Challenges Associated With Current TBI Models .................................. 98

1.8 Pharmacological and Non-pharmacological Management of TBI: An Overview

of Randomized Clinical Trials ............................................................................. 103

1.9 Study Rationale, Hypothesis, and Objectives ............................................... 106

x

Chapter 2: The Liver X Receptor Agonist GW3965 Improves Recovery from Mild

Repetitive Traumatic Brain Injury in Mice Partly Through Apolipoprotein E ................ 110

2.1 Summary ...................................................................................................... 110

2.2 Introduction ................................................................................................... 111

2.3 Materials and Methods .................................................................................. 115

2.3.1 Animals ................................................................................................ 115

2.3.2 Mild Repetitive Traumatic Brain Injury (mrTBI) .................................... 115

2.3.3 GW3965 Treatment ............................................................................. 119

2.3.4 Cognitive Assessment by Novel Object Recognition ........................... 120

2.3.5 Motor Function Assessment by Accelerating Rotarod ......................... 122

2.3.6 Biochemical Analyses .......................................................................... 122

2.3.6.1 Tissue Collection ..................................................................... 123

2.3.6.2 Protein Extraction .................................................................... 123

2.3.6.3 Western Blot Analyses ............................................................ 124

2.3.6.4 Quantitative Assessment of Endogenous Aβ by ELISA .......... 125

2.3.6.5 Quantitative Assessment of IL-6, TNFα, and MCP-1 by ELISA

............................................................................................................. 125

2.3.7 Histology .............................................................................................. 126

2.3.7.1 Assessment of Axonal Injury by Silver Staining ....................... 126

2.3.7.2 Assessment of Microglial Activation by Iba1

Immunohistochemistry ......................................................................... 127

2.3.8 Statistical Analyses .............................................................................. 128

2.4 Results .......................................................................................................... 129

xi

2.4.1 ApoE is Required for GW3965 to Improve Cognitive Function After

mrTBI ............................................................................................................ 129

2.4.2 Spontaneous Recovery of Motor Dysfunction After mrTBI is Unaffected

by GW3965 .................................................................................................. 133

2.4.3 GW3965 Prevents mrTBI-Induced Elevation of Endogenous Aβ Levels

..................................................................................................................... 135

2.4.4 GW3965 Enhances ABCA1 Induction Aafter mrTBI: ........................... 139

2.4.5 ApoE is Required for GW3965-Mediated Suppression of Axonal Damage

After mrTBI ................................................................................................... 142

2.4.6 Weight Drop TBI Model Produces Negligible Neuroinflammation........ 148

2.4.7 Retrospective Power Analysis ............................................................. 152

2.5 Discussion .................................................................................................... 154

Chapter 3: Merging Pathology and Biomechanics Using CHIMERA: A Novel, Surgery-

Free, Closed-Head Impact Model of Engineered Rotational Acceleration................... 161

3.1 Summary ...................................................................................................... 161

3.2 Introduction ................................................................................................... 162

3.3 Materials and Methods .................................................................................. 165

3.3.1 CHIMERA Impactor ............................................................................. 165

3.3.1.1 Animal Holding Platform .......................................................... 165

3.3.1.2 Pneumatic Impactor System.................................................... 166

3.3.1.3 Impact Piston and Barrel System ............................................ 166

3.3.1.4 CHIMERA Calibration .............................................................. 168

3.3.2 CHIMERA Repetitive TBI (rTBI) Procedure ......................................... 170

xii

3.3.3 High-Speed Videography and Kinematic Analyses ............................. 172

3.3.4 Behavioral Analyses ............................................................................ 174

3.3.5 Tissue Collection and Processing ........................................................ 178

3.3.6 Iba1 Immunohistochemistry and Silver Staining .................................. 179

3.3.7 Biochemical Analyses .......................................................................... 180

3.3.7.1 Tissue Processing ................................................................... 180

3.3.7.2 Assessment of TNF-α and IL-1β by ELISA .............................. 180

3.3.7.3 Quantitative Assessment of Phosphorylated and Total Tau by

Simple Western Analysis ..................................................................... 180

3.3.8 Statistical Analyses .............................................................................. 182

3.4 Results .......................................................................................................... 183

3.4.1 Head Kinematics Following CHIMERA rTBI ........................................ 183

3.4.2 CHIMERA rTBI Induces Behavioral Deficits ........................................ 189

3.4.3 CHIMERA rTBI Induces Widespread Diffuse Axonal Injury ................. 193

3.4.4 CHIMERA rTBI Induces Widespread Microglial Activation .................. 196

3.4.5 CHIMERA rTBI Increases Proinflammatory Cytokine Levels .............. 202

3.4.6 CHIMERA rTBI Increases Endogenous Tau Phosphorylation ............. 203

3.5 Discussion .................................................................................................... 205

Chapter 4: Conclusions and Future Directions ............................................................ 213

4.1 Chapter 2: Conclusions ................................................................................. 214

4.2 Chapter 2: Study Limitations and Future Directions ...................................... 219

4.3 Chapter 3: Conclusions ................................................................................. 223

4.4 Chapter 3: Study Limitations and Future Directions ...................................... 227

xiii

Bibliography ................................................................................................................ 230

xiv

List of Tables

Table 1.1 TBI incidence (in %) by cause across the globe. ............................................. 6

Table 1.2: Glasgow coma scale. ..................................................................................... 8

Table 1.3: North American classification of TBI according to severity of clinical signs. ... 8

Table 1.4: TBI classification according to European Federation of Neurological Societies

guidelines. ..................................................................................................................... 10

Table 1.5: Differences in the pathological features of CTE and AD. ............................. 53

Table 1.6: Pathological and clinical stages of CTE. ...................................................... 54

Table 1.7: Summary of physical and physiological properties of apoE alleles. ............. 75

Table 2.1: Details of primary antibodies used for Western blotting. ............................ 124

Table 2.2: Details of primary and secondary antibodies used for cytokine ELISA. ...... 126

Table 2.3: Summary of retrospective power calculations ............................................ 154

Table 3.1: Neurological severity score (NSS) tasks. ................................................... 178

Table 3.2: Details of monoclonal antibodies for probing p-tau and total tau. ............... 181

Table 3.3: Summary of peak values of head kinematic parameters. ........................... 187

Table 3.4: Comparison of head kinematic parameters between rodent TBI models and

human TBI. .................................................................................................................. 188

Table 4.1: Comparison of weight drop-rTBI and CHIMERA-rTBI models……………...226

xv

List of Figures

Figure 1.1: Annual incidence of TBI in people under 40 years of age ............................. 5

Figure 1.2: Biomechanics of head impact and brain tissue deformation. ...................... 14

Figure 1.3: Wayne State Tolerance Curve (WSTC). ..................................................... 20

Figure 1.4: Primary TBI. ................................................................................................ 23

Figure 1.5: Proposed molecular mechanism of diffuse axonal injury and generation of

Aβ. ................................................................................................................................. 33

Figure 1.6: Amyloidogenic and anti-amyloidogenic processing of APP. ........................ 42

Figure 1.7: Regulation of ApoE Lipidation by LXR. ....................................................... 62

Figure 1.8: ApoE structure, polymorphism and domain interaction. .............................. 74

Figure 1.9: Animal models of TBI. ................................................................................. 89

Figure 1.10: Fluid percussion injury model. ................................................................... 91

Figure 1.11: Controlled cortical impact model. .............................................................. 92

Figure 1.12: Blast Wave TBI Model. .............................................................................. 97

Figure 1.13: Variability in input parameters and outcomes in weight-drop TBI studies

reported in the literature. ............................................................................................. 100

Figure 2.1: Weight drop CHI model and impact location. ............................................ 118

Figure 2.2: Average animal body weight and drop height across all study groups. ..... 119

Figure 2.3: Novel object recognition phases. .............................................................. 122

Figure 2.4: Summary of mrTBI, Gw3965 treatment and post-injury outcomes. ........... 128

Figure 2.5: ApoE is required for GW3965 to improve NOR performance after mrTBI. 131

Figure 2.6: NOR performance was not affected by motor impairment. ....................... 132

xvi

Figure 2.7: mrTBI-induced motor impairment recovers spontaneously independent of

GW3965 and apoE. ..................................................................................................... 134

Figure 2.8: GW3965 prevents mrTBI-induced accumulation of endogenous Aβ in WT

and apoE-/- mice. ........................................................................................................ 137

Figure 2.9: APP and APP-CTFα levels remain unchanged following mrTBI. .............. 138

Figure 2. 10: GW3965 augments ABCA1 levels in WT and apoE-/- mice following

mrTBI. ......................................................................................................................... 140

Figure 2.11: ApoE and LDLR levels are unaffected by mrTBI or GW3965. ................. 141

Figure 2.12: Loss of apoE exacerbates axonal injury after mrTBI. .............................. 143

Figure 2.13: mrTBI leads to mild axonal damage in WT mice. .................................... 144

Figure 2.14: Loss of apoE exacerbates axonal damage after mrTBI. ......................... 145

Figure 2.15: ApoE is required for GW3965 to suppress axonal damage after mrTBI. . 147

Figure 2.16: mrTBI does not induce microglial activation in the cortex. ...................... 149

Figure 2.17: mrTBI induces negligible hippocampal microglial activation.................... 150

Figure 2.18: Pronounced microglial activation is localized only around contused areas.

.................................................................................................................................... 151

Figure 2.19: Pronounced microglial activation is localized only around contused areas.

.................................................................................................................................... 152

Figure 3.1: CHIMERA impactor and impact pistons. ................................................... 168

Figure 3.2: Piston energy-air pressure calibration curve. ............................................ 169

Figure 3.3: Mouse head and impact position. .............................................................. 171

Figure 3.4: External markers used for mouse head tracking. ...................................... 174

Figure 3.5: Open field thigmotaxis. .............................................................................. 177

xvii

Figure 3.6: CHIMERA-rTBI procedure and post-rTBI endpoints ................................. 182

Figure 3.7: Head kinematics following CHIMERA rTBI. ............................................... 185

Figure 3.8: CHIMERA allows unrestricted head motion during TBI. ............................ 186

Figure 3.9: CHIMERA rTBI induces behavioral deficits. .............................................. 192

Figure 3.10: CHIMERA rTBI does not significantly affect general mobility. ................. 193

Figure 3.11: CHIMERA rTBI induces diffuse axonal injury. ......................................... 195

Figure 3.12: CHIMERA rTBI induces sustained axonal injury ..................................... 196

Figure 3.13: CHIMERA rTBI induces widespread microglial activation. ...................... 199

Figure 3.14: Quantitative analysis of microglial response to rTBI. ............................... 201

Figure 3.15: CHIMERA rTBI increases proinflammatory cytokine levels. .................... 202

Figure 3.16: CHIMERA rTBI increases endogenous tau phosphorylation. .................. 204

xviii

List of Abbreviations

ABCA1: Adenosine triphosphate binding cassette transporter A1

Aβ: Amyloid beta peptide

ACRM: American Congress of Rehabilitation Medicine

AICD: Amyloid intracellular domain

ApoE/apoE: Apolipoprotein E

ApoA-I/apoA-I: Apolipoprotein A-I

APP: Amyloid precursor protein

AD: Alzheimer disease

ANOVA: Analysis of variance

BASE-1: Beta site APP cleaving enzyme-1

BBB: Blood-brain barrier

CAA: Cerebral amyloid angiopathy

CCI: Controlled cortical impact

CDC: Centers for disease control and prevention

CG: Center of gravity

CHI: Closed-head injury

CHIMERA: Closed-head impact model of engineered rotational acceleration

CNS: Central nervous system

CSF: Cerebrospinal fluid

CT: Computed tomography

CTE: Chronic traumatic encephalopathy

CTF: APP C-terminal fragment

xix

DAI: Diffuse axonal injury

DI: Discrimination index

DMSO: Dimethyl sulfoxide

DTI: Diffuse tensor imaging

DVI: Diffuse vascular injury

EDH: Epidural hemorrhage/hematoma

EFNS: European federation of neurological societies

FP: Fluid percussion

GCS: Glasgow coma scale

HDL: High-density lipoprotein

HIC: Head injury criterion

ICH: Intracranial hemorrhage/hematoma

ICP: Intracranial pressure

IDE: Insulin degrading enzyme

IL-1β: Interleukin 1 beta

IL-6: Interleukin 6

i.p.: Intraperitoneal

ISF: Interstitial fluid

LDL: Low-density lipoprotein

LDLR: LDL receptor

LRP1: LDLR related protein 1

LOC: Loss of consciousness

LRR: Loss of righting reflex

xx

LTP: Long-term potentiation

LXR: Liver X receptor

LXRE: LXR response element

MAPT: Microtubule associated protein tau

MCP-1: Monocyte chemotactic protein-1

MRI: Magnetic resonance imaging

mTBI: Mild traumatic brain injury

mrTBI: Mild, repetitive traumatic brain injury

MTBR: Microtubule binding repeat

MVA: Motor vehicle accident

NEP: Neprilysin

NFL: National football league

NFT: Neurofibrillary tangles

NMDA: N-methyl-d-aspartate

NOR: Novel object recognition

NSS: Neurological severity score

OHI: Open head injury

PBS: Phosphate-buffered saline

PS-1/2: Presenilin 1/2

PTA: Post-traumatic amnesia

RCT: Reverse cholesterol transport

RIPA: Radioimmunoprecipitation assay

rTBI: Repetitive traumatic brain injury

xxi

RXR: Retinoid X receptor

TBI: Traumatic brain injury

SDS: Sodium dodecyl sulfate

s.c.: Subcutaneous

SCI: Spinal cord injury

SDH: Subdural hemorrhage/hematoma

TNF-α: Tumor necrosis factor alpha

VLDL: Very low-density lipoprotein

WHO: World health organization

WSTC: Wayne state tolerance curve

WT: Wild-type (C57Bl/6) mice

xxii

Acknowledgements

As I come to this milestone in the pursuit of knowledge, I cannot help but reflect back

and acknowledge all those who have contributed towards completion of this work in one

or the other way.

First of all, I offer my deepest gratitude to Almighty GOD for showering ITS infinite

bounties, graces and mercies on me and giving me strength in the most difficult times.

Without ITS wishes and blessings, this work could have remained a dream only.

I would like to dedicate this work to my parents for their unconditional love, constant

encouragement, to my wife Archana for her immense love, patience, and unstinting

support, who stood with me in every good and bad moment and to my parents-in-law for

their constant motivation and support.

I express my deepest gratitude for my mentor Dr. Cheryl Wellington. Thank you Cheryl

for your guidance, support, constant encouragement, constructive criticism, and

valuable suggestions throughout my work. It has been an enriching experience and

privilege working in your laboratory.

I sincerely thank my supervisory committee members, Drs. Brian MacVicar, Wolfram

Tetzlaff, and Gordon Francis for providing critical guidance throughout this work.

xxiii

I thank all the current and past Wellington Lab members, especially, Anna Wilkinson,

Jeniffer Chan, Dr. James Donkin, Dr. Veronica Hirsh-Reinshagen, Dr. Kris Martens, Dr.

Jianjia Fan, Dr. Sophie Stukas, Tom Cheng, Mike Carr, Georgina Martin, and Arooj

Hayat for all the help and support throughout this work. Special thanks to Tom for all the

intellectual talks and exchange of ideas and who has a lion’s share in the development

of CHIMERA.

Special thanks to Dr. Peter Cripton and Kurt McInnes for helping me understand the

biomechanics of traumatic brain injury as well as for the wonderful collaboration that

resulted in the development of CHIMERA.

I express my gratitude towards Alzheimer Society of Canada and its patrons for

supporting me through Alzheimer Society Research Program doctoral award. This work

was also supported by the grants from Canadian Institutes of Health Research to Dr.

Cheryl Wellington.

I am grateful to Child and Family Research Institute, Centre for Disease Modeling, as

well as Djavad Mowafaghian Centre for Brain Health for allowing me to use their core

facilities and space to conduct this research work.

Last but not the least I would like to acknowledge the unsung heroes of this work, the

laboratory mice without whom this work would not be possible.

xxiv

Dedication

To my parents and parents-in-law.

To my wife, Archana

xxv

Chapter 1: Introduction

1.1 Traumatic Brain Injury (TBI): Definition

Traumatic brain injury (TBI) occurs when an external mechanical force traumatically

injures the brain resulting in altered mental state. TBI is a type of acquired brain injury

that is caused by events after birth as opposed to genetic or congenital defects. As

discussed further in this Chapter, TBI pathology is heterogeneous and each TBI is

unique and hence defining TBI has been challenging. The definition as well as the

classification of TBI has evolved through several attempts by different groups. As

discussed further about 75% of the human TBI are mild and therefore in most of the

attempts of defining TBI have considered only mild TBI (mTBI). The most commonly

used outcomes defining mTBI are Glasgow Coma Scale (GCS) score, loss of

consciousness (LOC) and post-traumatic amnesia (PTA). For example, the American

Congress of Rehabilitation Medicine (ACRM) defined mTBI as “traumatically induced

physiological disruption of brain function as manifested by at least one of the following:

1) any period of LOC, 2) any loss of memory events immediately before or after the

accident, 3) any alteration in mental state (e.g., dizziness, disorientation, or confusion)

at the time of the accident, and 4) focal neurological deficits which may or may not be

transient; but where the severity of injury does not exceed the following: LOC ≤ 30 min,

GCS score of 13-15 after 30 min, and PTA < 24 h” (Kay et al., 1993). The U.S. Centers

for Disease Control and Prevention (CDC) proposed a general definition of mTBI as “ an

injury to the head as a result of blunt trauma or acceleration forces that result in: 1)

transient confusion, disorientation, or impaired consciousness or 2) dysfunction of

memory around the time of injury or 3) LOC < 30 min or 4) neurological or

1

neuropsychological dysfunction such as seizures following injury (Gerberding and

Binder, 2003). The CDC also recommends developing case-specific definition of mTBI

based on this general definition. Later, the World Health Organization (WHO)

Collaborating Centre for Neurotrauma Task Force on Mild Traumatic Brain Injury

published another definition for mTBI as “an acute brain injury resulting from mechanical

energy to the head from external physical forces” manifesting into confusion or

disorientation, LOC for 30 min or PTA for < 24 h and GCS score of 13-15 after 30 min

(Carroll et al., 2004). The WHO definition further emphasizes that mTBI manifestations

must not be due to drugs, alcohol, or injuries caused by other mechanisms, e.g.,

systemic injuries, facial injuries or penetrating craniocerebral injury.

Thus, the ACRM, CDC, and WHO definitions address only the mild spectrum of TBI and

not moderate and severe TBIs, including penetrating brain injuries. Moreover, these

definitions consider only neurological signs while ignoring any evidence of the resulting

brain pathology.

Most recently, the Demographics and Clinical Assessment Working Group of the

International and Interagency Initiative toward Common Data Elements for Research on

Traumatic Brain Injury and Psychological Health proposed a broader definition of TBI as

“an alteration in brain function, or other evidence of brain pathology caused by an

external force” (Menon et al., 2010).

2

The alteration of brain function is assessed by clinical examination and includes any

one of the following: 1) any period of LOC, 2) any loss of memory immediately before

(retrograde amnesia) or after (PTA) TBI, 3) neurological deficits including weakness,

loss of balance, sensory loss, aphasia (problem with speaking, listening, reading and

writing), and paralysis, and 4) any alteration in mental state (confusion, disorientation) at

the time of injury (Menon et al., 2010). With the advances in modern imaging techniques

such as Magnetic Resonance Imaging (MRI), Diffusion Tensor Imaging (DTI) and

Computed Tomography (CT), it is now possible to visualize brain damage in some

cases. The neurological examination of TBI can therefore be supplemented with these

techniques (in addition to potential biomarkers of great interest for future use) to obtain

evidence of brain pathology. Lastly, TBI is caused by an external mechanical force,

which distinguishes TBI from other types of acquired brain injuries such as

cerebrovascular accident (stroke), brain tumors, brain infection, and brain injury caused

by substance abuse. The external mechanical force can be an impact (e.g., a moving

object striking stationary head such as during an assault or a moving head striking the

stationary object such as during a fall) or non-impact (e.g., rapid head movement during

whiplash or head exposed to blast waves) or a penetrating force (e.g., a bullet

penetrating the skull).

The new definition also encourages using the more appropriate term “traumatic brain

injury” rather than “head injury”, which may be limited to injury to the face and the scalp

and may not result in neurological deficits.

3

1.2 TBI: Epidemiology

TBI is a leading cause of death and disability throughout the world. The global annual

incidence of TBI is estimated to be ~200 per 100,000 people (Reilly, 2007). The annual

incidence of TBI in the North America is greater than the combined incidence of breast

cancer, HIV/AIDS, multiple sclerosis, and spinal cord injuries (Figure 1.1A). In the

United States, the overall incidence of TBI is estimated to be 538 per 100,000

population, which represents at least 1.7 million new cases per year since 2003

(Gerberding and Binder, 2003; Langlois et al., 2006; Faul et al., 2010). Annually

approximately 50,000 Canadians sustain a TBI with more than 11,000 deaths occurring

each year (Brain Injury Society of Brain Injury Society of Toronto, 2014). The rate of TBI

is reportedly lower in Europe (235 per 100,000) and in Australia (322 per 100,000)

(Cassidy et al., 2004; Tagliaferri et al., 2006) although the emerging evidence indicates

the otherwise (see below).

Estimating the true burden of TBI is challenging due to several factors. Most of the

epidemiological studies report data based on hospitalization and government records

that lack systematic epidemiological monitoring. Compounding this is the growing

awareness that more than 75% of TBI are mild (mTBI, a term synonymous to

concussion) that do not necessarily need hospitalization and therefore are not always

reported. As will be noted below, underdiagnosis of mTBI may nonetheless pose

challenges to interpretation of potential long-term consequences. Finally, most of the

epidemiological studies are retrospective, include small patient number and suffer from

recall bias. These limitations have been highlighted by a recent study in which a

4

substantially higher TBI incidence (749 per 100,000) was observed than previously

appreciated (Feigin et al., 2013). In this study, the authors collected data in an urban

and rural New Zealand population by both prospective and retrospective surveillance

systems. This study suggested that TBI incidence may be far greater than reported in

early epidemiological studies that mostly used retrospective data.

Furthermore, although most of the epidemiological data on TBI comes from the

developed countries, it is estimated that the incidence of TBI in developing countries is

rising and posing a significant health problem (Figure 1.1B). With the increasing

worldwide incidence rate as well as high number undiagnosed/underdiagnosed cases, it

is not surprising that TBI is often called a “silent epidemic”.

Figure 1.1: Annual incidence of TBI in people under 40 years of age (A) TBI has higher incidence rate compared to the combined incidence of other major health

issues including multiple sclerosis (MS), spinal cord injury (SCI), HIV/AIDS and breast cancer

combined. (B) Estimated annual incidence of TBI across the globe. For data sources, please

refer Table 1.1.

5

Falls and motor vehicle accidents (MVA) are the two most common causes of severe

TBI, with assaults/violence often being the third most-common cause (Table 1.1)

(Cassidy et al., 2004; Hyder et al., 2007; Faul et al., 2010). TBI resulting from high-

contact/collision sports such as boxing, American football, ice hockey, soccer, and

rugby account for almost 21% of all head injuries among children and adolescents

(Centers for Disease Control and Prevention 2007; American Association of

Neurological Surgeons, 2011). TBI is also considered a “signature injury” in modern

warfare, as approximately 20% of veterans from the Iraq or Afghanistan wars have

experienced a TBI, 80% of which involve blast injury (Taber et al., 2006; Hoge et al.,

2008; Elder and Cristian, 2009).

Country/ Region Falls MVA War Assault/

Violence Other/

Undefined Source

Canada 45 23 - 9 10

Canadian Institute for Health Canadian Institute for Health Information (2006)

USA 35 17 - 10 21 Faul et al. (2010)

Europe 39 39 - 7.2 19 Tagliaferri et al. (2006)

Australia 21 40 - 8 25 Reilly (2007) New

Zealand 38 20 - 17 21 Feigin et al. (2013)

China 17.5 46.3 - 23.8 21 Zhao and Wang (2001); Wu et al.

(2008) India 50 22 - 10 22 Gururaj (2002) Latin

America 24 41 4 25 8 Puvanachandra and Hyder (2008)

Africa 5 13 53 33 16 Hyder et al. (2007) Global 8 62 2 24 4 Hyder et al. (2007)

Table 1.1 TBI incidence (in %) by cause across the globe.

6

MVA: motor vehicle accident.

1.3 TBI: Classification

Traditionally TBI is classified based on three systems: 1. clinical severity, 2.

pathophysiology and 3. physical mechanism (Saatman et al., 2008).

1.3.1 TBI Classification Based on Clinical Severity

Conventional clinical TBI taxonomy divides severity of injury into three categories; mild,

moderate and severe. The most commonly accepted method to determine injury

severity is the GCS score (Table 1.2), which measures the patient’s level of

consciousness based on verbal, motor, and eye opening responses after injury. A

patient with a GCS score of 3-8 (out of 15) is considered to have sustained a severe

traumatic brain injury, 9-12 is moderate, and >12 mild (Table 1.3) (Teasdale and

Jennett, 1974). While the GCS is useful for the clinical management of TBI, it does not

provide specific information about the neuropathological mechanisms involved. The

GCS is also less useful in pediatric TBI. The prognostic ability of this system thus is

limited and as such other descriptors have been added. In the most recent iteration of

the widely adopted Departments of Defense and Veteran Affairs classification of

severity of brain injury, mTBI is further denoted by PTA lasting < 24 h, and a LOC of <

30 min. Similarly, to meet moderate TBI criteria, GCS must be between 9-12, PTA must

not exceed one week, and LOC must not last longer than 24 h (U.S Departments of

Defense and Veterans United Sates Departments of Defense and Veterans Affairs,

2008) (Table 1.3).

7

Score Eye Opening Response Verbal Response Motor Response

6 N/A N/A Obeys commands for movement

5 N/A Oriented Purposeful movements to painful stimulus

4 Spontaneous Confused, but able to answer questions

Withdraws in response to pain

3 To verbal stimuli Inappropriate words

Flexion in response to pain (decorticate posturing)

2 To pain only (not applied to face)

Incomprehensible speech

Extension in response to pain (decerebrate posturing)

1 No response No response No response

Table 1.2: Glasgow coma scale.

Glasgow coma scale determines a patient’s neurological state based on eye opening,

verbal and motor response. Adapted from (Teasdale and Jennett, 1974).

Injury Severity GCS Score (Out of 15)* LOC Duration PTA Duration Mild > 12 < 30 min < 24 h

Moderate 9-12 30 min to 24 h 1 to 7 days Severe 3-8 > 24 h > 7 days

Table 1.3: North American classification of TBI according to severity of clinical signs.

* Best available score within 24 h of injury. GCS: Glasgow Coma Score, LOC: Loss of

Consciousness, PTA: Posttraumatic Amnesia.

The European Federation of Neurological Societies (EFNS) guidelines have taken a

slightly different approach and categorized TBI into four levels of severity: mild,

moderate, severe, and critical (Table 1.4) (Vos et al., 2002). The heterogeneity inherent

8

to mTBI has been addressed with various refinements to this category. For example,

the EFNS guidelines further classify mTBI into four categories, 0, 1, 2, and 3. In addition

to the parameters used by the North American guidelines (i.e., GCS score, duration of

LOC and PTA), the EFNS guidelines also include risk factors for intracranial

complications in the grading of mTBI severity. The risk factors for intracranial

complications include unclear or ambiguous accident history, continued PTA (GCS

verbal score of 4), retrograde amnesia for > 30 min, trauma above clavicles including

clinical signs of skull fracture, severe headache, vomiting, focal neurological deficits,

seizure, age of < 2 or > 60 years, coagulation disorders, and alcohol/drug intoxication

(Vos et al., 2002). In this system, the subcategories of mTBI dictate post-TBI

management. Thus, a person with mTBI category 0 is considered to have no TBI and is

immediately discharged, mTBI category 1 recommends CT scanning while CT scanning

is mandatory for persons with mTBI categories 2 and 3 (Vos et al., 2002). The finer

distinction of mTBI may be of questionable value since in a recent major revision of

evidence-based guidelines for evaluation and management of sports-related

concussions, endorsed by multiple sporting bodies and physician groups, these finer

distinctions have now been dropped due to a lack of prognostic utility (Giza et al., 2013).

9

TBI Class GCS Score

LOC Duration

PTA Duration

Risk Factors for Intracranial

Complications Mild GCS = 13-15 mild Category 0 15 No No Not present mild Category 1 15 < 30 min < 1 h Not present mild Category 2 GCS = 15 and risk factors present mild Category 3 13-14 < 30 min < 1 h With or without risk factors Moderate GCS = 9-12 Severe GCS ≤ 8

Critical GCS = 3-4 with loss of pupillary reactions and absent or decerebrate motor reactions

Table 1.4: TBI classification according to European Federation of Neurological Societies guidelines.

GCS: Glasgow Coma Score, LOC: Loss of Consciousness, PTA: Posttraumatic

Amnesia. Adapted from (Vos et al., 2002).

1.3.2 TBI Classification Based on Pathophysiology

The action of external forces on the head leads to a cascade of pathological processes.

These processes are broadly classified into primary and secondary brain injuries

(discussed in section 1.5). Primary brain injury occurs at the time of mechanical loading

on the head leading to structural damage to the brain parenchyma and

cerebrovasculature. Primary injury results in laceration or contusion of the cortical

surface, diffuse axonal injury (DAI), and hematoma. Damage caused by primary injury is

considered to be irreversible and may not respond to any pharmacological intervention.

Primary brain injury initiates a plethora of secondary processes that result in complex

cellular, inflammatory, neurochemical, and metabolic alterations (McIntosh et al., 1996;

Blumbergs, 1997; Davis, 2000; Giza and Hovda, 2001; Werner and Engelhard, 2007;

10

McAllister, 2011). The secondary injury pathways may be more amenable to

pharmacological treatment.

1.3.3 TBI Classification Based on Physical Mechanism

TBI is caused by external mechanical forces, which can be static or dynamic. Static

stress (e.g., a head caught between elevator doors) does not cause head acceleration

and does not lead to brain injury unless the crushing forces exceed the threshold

required for skull fracture causing substantial tissue damage (Gennarelli, 1993). The

majority of TBIs result from dynamic forces including contact (impact) or non-contact

(inertial) forces. Dynamic stresses induce rapid acceleration and/or deceleration of the

head followed by distribution of inertial forces to the underlying brain tissue causing

tissue damage. Contact injuries are induced by a moving object striking a stationary

head or a moving head striking a stationary object and usually involve intense

mechanical loading of short duration (Davis, 2000). The basis of classifying TBI on the

basis of physical mechanism lies in the observed relation between the type of

mechanical force and the resulting neuropathological characteristics. Thus, brain

injuries as a result of impact forces are thought to induce localized (focal) damage

including contusions, skull fracture, and epidural hematoma while inertial forces tend to

cause more widespread (diffuse) brain damage such as diffuse axonal injury and

subdural hematoma (McLean and Anderson, 1997; King, 2000; Saatman et al., 2008).

11

1.4 TBI: Biomechanical Principles

Biomechanics involves the study of the motion (such as head acceleration) of a person,

animal or anthropomorphic test device (crash test device) and mechanical loads

sustained or applied (such as head impact forces). The work of Gurdjian and Lissner

(Gurdjian and Lissner, 1945, 1946, 1947; Gurdjian et al., 1947; Gurdjian and Webster,

1947) and Holbourn (Holbourn, 1943, 1945) pioneered the biomechanical evaluation of

TBI to understand the relationship between mechanical loading (force/stress) and the

physical response of the head and brain (deformation/strain) and resulting pathology.

Besides understanding the mechanism of brain injury, biomechanical studies also help

to predict head injury tolerance levels, which are used for developing and standardizing

vehicle safety parameters as well as in helmet design to provide improved protection for

the head during impact (Versace; Henn, 1998; Eppinger et al., 2000). Although, the

mechanisms by which mechanical forces induce brain injury are a subject of

considerable debate (Hardy et al., 1994; Drew and Drew, 2004), brain deformation or

strain resulting from external mechanical loading (i.e., force/stress) is the currently

accepted biomechanical theory of TBI (McLean and Anderson, 1997; King et al., 2003;

King et al., 2004; LaPlaca et al., 2007). Brain deformation increases with increasing

stress in a non-linear fashion (LaPlaca et al., 2007). In a head impact, dynamic

mechanical forces act on the skull and brain to cause both linear and rotational

movement of the head and skull. This in turn leads to deformation and structural

damage of brain and cerebrovascular tissues and triggers many secondary injury

pathways.

12

Human impact TBI can occur under many conditions. One example involves the slowly

moving or stationary head being impacted by a rapidly-moving object, such as when a

vehicle strikes a pedestrian’s head. Another example involves a head moving at a high

rate of speed impacting a stationary object, such as when a hockey player slides head

first into the boards. These impacts cause intense mechanical loading that lasts for only

a fraction of a second (< 50 ms) (Gurdjian et al., 1966; Ono et al., 1980; Pellman et al.,

2003b) and causes pressure gradients and mechanical strain (i.e., local areas of

stretching or compression of the brain tissue) to be induced within the brain tissue as

shown in Figure 1.2 (Meaney and Smith, 2011). Head impact can result in pure linear

(straight line) motion of the head or combined linear and rotational motion in response

to impact. Combined linear and rotational motion is more common than pure linear

motion because the head is coupled to the body by the neck and so almost any impact

to the head results in a combination of linear and rotational motion of the head due to

restraint forces on the head applied by the neck (Greaves et al., 2009). In pure linear

motion of the head, the pressure gradients and tissue strains described above will

occur. In combined rotational and translational motion of the head the pressure

gradients and tissue strains described above will have much larger tissue strains, which

arise from the rotation of the head, superimposed on the pressure gradient-related

strains (Meaney and Smith, 2011).

13

Figure 1.2: Biomechanics of head impact and brain tissue deformation.

(A) Impact to the back of the skull with the skull moving downwards (red arrow) causes

momentary skull deformation (red outline). (B) The skull stops suddenly (in ~ 50 ms)

and the momentum of the brain keeps it moving causing relative motion of the brain with

respect to the skull (orange outline). (C) This motion of the brain sets up positive

pressure at the impact site (i.e., coup) and negative pressure opposite the coup site at

the contrecoup site. Reproduced with permission from Namjoshi et al. (2013a)

Impact is by far the most common cause of TBi in civilian populations. In the military,

however, improvised explosive devices (IEDs) can lead to blast forces capable of

causing TBI. In contrast to impact TBI, pure blast TBI is non-contact and involves

dynamic forces with very short durations on the order of a few microseconds (µs)

(Goldstein et al., 2012). Other non-contact TBI mechanisms are hypothesized to occur

as a result of “inertial” loading of the head where torso motion, for example when a

football player’s torso is impacted by another player during a tackle or during “shaken

baby” syndrome, causes the head to move even when no direct impact forces are

subjected to it. This mechanism of TBI is a matter of some controversy and many

investigators have concluded that this does not occur in real world human TBI (Lau et

al., 1989; McLean, 1995; Yoganandan et al., 2009; Meaney and Smith, 2011; Wright et

14

al., 2013). The duration of inertial head loading in these situations varies from 50 to 200

ms (Gennarelli, 1993). Most epidemiological studies also conclude that these injuries

cannot feasibly occur in adult humans (Lau et al., 1989; McLean, 1995).

TBI can also result from static or near static loads that essentially crush the brain and

skull resulting in direct compression of the brain or contusion injury via bone fragments

(Denny-Brown and Russell, 1940; Lopez-Guerrero et al., 2012; Mattei et al., 2012). In

these crushing or nutcracker injuries, the head is generally not subjected to rapid linear

or rotational movements that occur during impact injuries.

1.4.1 Head Motion during Impact TBI

Head/skull acceleration that occurs during a head impact can be described using three-

dimensional linear (translational) and rotational (angular) accelerations. Linear

acceleration is defined as the change in velocity over a given time through translational

coordinates of the head’s center of gravity (CG) e.g., x-y-z or its resultant, and is usually

expressed in units of “g” (one g is the acceleration due to gravity on earth) or m/s2.

Rotational acceleration is the change in rotational velocity of the head over a given time

and is expressed in units of radians/s2 (rad/s2) or degrees/s2 (°/s2). One revolution is

equal to 360° or 2π rad. Acceleration can be measured using devices called

accelerometers.

The relative amounts of linear and rotational head acceleration that result from a

particular head impact depend on several factors including the type of impact force, the

15

direction of the force, the location of force on the skull, and the material properties of the

skull and brain. An impulsive contact force applied to the head is a vector with

magnitude and direction. A force that passes through the CG of the head (i.e., aligned

with the maxilla) will initiate primarily linear motion of the head during the impact. A

force that does not pass through the CG (e.g., an impact to the high forehead) will

produce an impulsive moment (conceptually a “twisting force”) about the CG and will

initiate both linear and rotational acceleration.

There is considerable debate about whether linear (Gurdjian et al., 1955; Haddad et al.,

1955; Gurdjian et al., 1961) or rotational (Holbourn, 1943; Gennarelli et al., 1981;

Gennarelli and Thibault, 1982; Gennarelli et al., 1982) acceleration is a better predictor

of brain injury. Advocates of rotational acceleration argue that pure linear impact is rare

in the clinical setting and angular acceleration is the principle mechanism underlying

brain injury (Holbourn, 1943; Hardy et al., 1994). Notably, the Head Injury Criterion

(HIC), which is a currently incorporated in vehicle safety standards around the world,

takes only linear acceleration into account. The HIC is calculated as a function of the

acceleration magnitude and duration of acceleration so that high accelerations acting for

long time intervals result in high HIC while lower acceleration values or shorter

exposure times result in lower HIC. In the context of automotive safety testing, the HIC

has been credited with considerably reducing the incidence of MVA-related head

injuries for over three decades (King et al., 2004). Moreover, recent studies by King and

colleagues contend that helmets significantly reduce linear acceleration without

changing rotational acceleration (King et al., 2003), leading these authors to propose

16

that the response of the brain itself (i.e., deformation of the structures of the brain) to

mechanical loading may be a better predictor of brain injury than linear or rotational

acceleration of the skull (Hardy et al., 2001; King et al., 2003; King et al., 2004; King et

al., 2011). Many authors have proposed metrics that are a combination of both linear

and rotational acceleration (and other factors such as HIC and impact force location on

the skull) (Gurdjian, 1975; Ono et al., 1980; Pellman et al., 2003b; Greenwald et al.,

2008; Rowson and Duma, 2013) as the most predictive mechanisms of brain injury.

1.4.2 Brain Motion during TBI

Like most soft tissue, the brain has viscoelastic properties with non-linear mechanical

stress-strain responses (LaPlaca et al., 2007). Importantly, it is shear strain, rather than

tissue compression or pressure gradients, which is believed to be the major mechanism

underlying most concussion pathology (Meaney and Smith, 2011). Brain motion inside

the skull during impact has been studied using a variety of techniques. In the earlier

studies, brain motion during impact was directly recorded through a cranial window

created by surgically replacing the skull cap of Macaque monkeys with a cap made up

of acrylic (called the “Lucite Calvarium”) (Shelden et al., 1944; Pudenz and Shelden,

1946) or polycarbonate (the “Lexan Calvarium”) (Ommaya et al., 1969; Gosch et al.,

1970). Brain motion following sub-concussive impacts in the frontal, temporal, and

parietal regions was recorded using high-speed film. In these studies, it was found that

the regardless the direction of impact, the principal brain displacement was in parieto-

occipital region with minimum displacement in the frontal region, which is likely due to

the anterior cranial fossa that restricts movement of frontal lobes. This study showed

17

relative movements of brain and the skull, concluding that the relative displacement of

the brain depended on the degree of skull movement. Thus, an immobile skull following

impact caused little to no brain displacement while a freely moving skull resulted in

considerable brain movement. With the advances in high-speed X-ray videography in

1960’s it was possible to record brain movement with minimal surgical manipulation.

Using intravascular contrast media and lead targets, brain motion was captured after

impact with high speed X-ray videography in dogs (Hodgson et al., 1966) and primates

(Shatsky et al., 1974). Hardy and colleagues improved on this technique by using

neutral density particles to record brain displacement in human cadavers (Hardy et al.,

2001; Hardy et al., 2007). The studies in human cadavers reported brain displacement

of ± 5 mm relative to the skull (King et al., 2011). These studies also reported a lag in

the brain displacement relative to the motion of the skull. Moreover, these studies also

show that the radio-opaque particles return to their original position following

displacement suggesting elastic properties of the brain tissue, which may be important

for studying brain displacement over repeat impacts. The effects of post-mortem

changes in the brain tissue on the brain displacement however are not known. Recently

brain motion was studied in human volunteers using tagged MRI. Bayly and colleagues

conducted a series of studies in which the head of human volunteers was subjected to

deceleration with occipital impact of 2-3 g (Bayly et al., 2005) or angular acceleration of

250-300 rad/s2 (Sabet et al., 2008) and the resulting strains were assessed using

tagged MRI (tagged MRI enables tracking the motions of tissue by tagging specific

tissue points with a sequence of radio-frequency pulses before imaging) although

without isolating relative movements of brain and the skull. By tracking the points along

18

tag lines imposed on MRI images, the authors reported that mechanical inputs that were

10-15% of those experienced by soccer players resulted in strains in the range of 0.02-

0.06 (a 0.05 strain corresponds to a 5% change in the dimension of local tissue).

1.4.3 The Clinical Picture: Human Tolerance and Related Biomechanical Studies

of Human TBI

The traumatic injury threshold for humans has been investigated using animal models,

physical brain surrogate models, analytical models and finite element models and

studies in athletes (football and hockey) who experience frequent head impacts and

concussion.

Using animal models in the 1950s, researchers at the Wayne State University

demonstrated that the duration of intracranial pressure was an important exposure

variable in injury tolerance; higher exposure generated a more severe injury in a shorter

amount of time (Lissner et al., 1960). Increasingly sophisticated instrumentation allowed

head acceleration to be measured at the occiput (posterior part of the skull) along with

changes in intracranial pressure in response to forehead impacts for whole and partial

cadavers on automotive instrument panels, windshields, and non-yielding surfaces

(Evans et al., 1958; Lissner et al., 1960). These experiments led to the preliminary

Wayne State Tolerance Curve (WSTC) for head injury. The initial tolerance curve

predicted whether head injury would occur as a function of the head impact duration

and the average linear acceleration measured at the occiput. This initial curve was

further refined through additional cadaver testing and in human volunteers (Eiband,

19

1959; Gurdjian et al., 1961; Patrick et al., 1963). The revised WSTC, as shown in Figure

1.3, assumed that the underlying experimental impacts that caused a linear skull

fracture also caused a moderate to severe concussion (Gurdjian et al., 1966). The first

experimental and biomechanics-based, quantitative human brain injury criterion, was

based on linear head acceleration. Rotational based human injury criteria would not

come until later (Newman, 2002). The WSTC data provides the basis for several widely

used injury metrics currently in use such as the Gadd Severity Index (GSI) (Gadd,

1966), and the HIC (Versace, 1971).

Figure 1.3: Wayne State Tolerance Curve (WSTC).

The WSTC describes the relationship between linear head acceleration, duration of

acceleration, and onset of concussion. The WSTC suggests that the head can

withstand very high acceleration for a very short duration. Conversely, any increase in

20

the duration of impact for the same intensity of acceleration is likely to cause head

injury. Reproduced with permission from Namjoshi et al. (2013a).

Several groups have generated head injury risk curves using logistic regression models

based on linear acceleration as a measure of exposure (Prasad and Mertz, 1985; Hertz,

1993; Kuppa, 2004). Prasad and Mertz suggest that a HIC15 (measurement of impact

over 15 ms) value of 700 represents a less than 5% risk of life-threatening brain injury.

More recently, Zhang et al. (2004) used a validated finite element human head model

and predicted the maximum resultant linear acceleration at the CG of the head to be 66,

82, and 106 g for a 25%, 50%, and 80% probability of mTBI, respectively, whereas the

maximum resultant rotational accelerations for a 25%, 50%, and 80% probability of

sustaining a mTBI are estimated at 4600, 5900, and 7900 rad/s2, respectively. Funk et

al. (2007) estimated a 10% risk of mTBI at 165 g, a HIC of 400, and an angular head

acceleration of 9000 rad/s2.

To put injury threshold experiments into context with mTBI, many groups have directly

measured or reconstructed concussive impacts during sporting events. Using helmets

fitted with triaxial accelerometers, Pellman et al. (2003b) reconstructed impacts

involving NFL players where players sustained concussions or significant head impacts

and determined that peak head accelerations in concussed players averaged 98 g and

in uninjured players averaged 60 g. The lowest measured acceleration where a player

sustained a concussion was 48 g. The peak angular acceleration in concussed and

uninjured players averaged 6432 rad/s2 and 4235 rad/s2, respectively. The lowest

angular acceleration where a player sustained a concussion was 2615 rad/s2. These

21

data are consistent with the predictions made from finite element modeling (FEM)

studies.

Researchers at Virginia Tech used instrumented helmets (helmets designed with

embedded accelerometers) to record tens of thousands of head impacts in football

players including 57 diagnosed concussions. Linear acceleration of 171 g, 192 g, and

214 g were identified to result in a 25%, 50%, and 75% risk of mTBI (Rowson and

Duma, 2011). Rotational accelerations of 5821 rad/s2, 6383 rad/s2, and 6945 rad/s2

were associated with a 25%, 50%, and 75% chance of mTBI (Rowson et al., 2012).

The most recent study investigated mTBI injury risk as a function of linear acceleration

alone, rotational acceleration alone and a combination of both linear and rotational

acceleration using logistic regression (Rowson and Duma, 2013). All three models were

found to be good predictors of mTBI outcome. Although the combined model is

preferred, it was statistically equivalent to the model using linear acceleration alone.

1.5 TBI: Pathophysiology

The TBI event can be divided into four stages: 1) mechanical forces (usually dynamic),

act on the skull and brain to cause 2) head motion (rotation and/or translation of the

head and skull) and 3) brain motion/deformation causing structural damage of brain and

vascular tissues (primary injury), which leads to 4) a delayed biological response of the

brain (i.e., deleterious biochemical and cellular processes that cause many TBI patients

to gradually deteriorate in the hours and days after injury) referred to as secondary

injury.

22

1.5.1 Primary Brain Injury

The primary brain injury is caused by mechanical forces resulting into structural damage

to the brain and the cerebrovasculature. The primary brain injury can be focal or diffuse

(Figure 1.4). Most clinical TBIs result in both focal and diffuse brain damage.

Figure 1.4: Primary TBI. Primary TBI results from structural damage caused to the brain parenchyma and

cerebrovasculature from mechanical forces. Focal injury results in contusion, laceration

and/or vascular injury

Primary TBI

Diffuse

Diffuse axonal injury

Diffuse vascular

injury

Focal

Skull Fracture Contusion Laceration Vascular

injury

Subdural hemorrhage

Epidural hemorrhage

Intracerebral hemorrhage

Intraventricular hemorrhage

23

1.5.1.1 Focal Brain Injury

Focal brain injury is caused with or without skull fracture by impact forces such as a

moving object striking against a stationary head (e.g., during physical assault) or a

moving head striking against a stationary object (e.g., during fall). Focal injuries are

thought to account for about two-thirds of TBI-related deaths (Davis, 2000). Whether or

not the head is stationary may result in characteristic focal damage. Thus, a moving

object striking a stationary head initiates head acceleration resulting in focal injury at the

side of the impact called “coup injury”. On the other hand, a moving head striking

stationary object results in sudden head deceleration and increases the probability of

focal damage to the side opposite to the impact site called “contrecoup injury” (Dawson

et al., 1980). Counterintuitively, the contrecoup injuries tend to be more severe than

coup injuries as reported in several clinical observations (Dawson et al., 1980). The

mechanisms of coup-contrecoup injuries are not well understood and are subject to

controversies and as such several theories are proposed. According to the positive

pressure theory, as the skull moves before impact, the brain lags and is pressed

against the lagging surface of the skull (i.e., side opposite to the forthcoming impact

site) (Lindenberg and Freytag, 1960; Drew and Drew, 2004). Also, as the brain lags, the

cerebrospinal fluid (CSF) is displaced in the space in the coup site created by the

lagging brain. The impact generates pressure waves that further increase compression

of the brain against the lagging surface of the skull. Accumulation of CSF near the

impact site further cushions against the coup injury causing more contrecoup damage.

The negative pressure theory (cavitation theory) argues that upon impact the head

suddenly stops while the brain still continues to move towards the impact site (Russel,

24

1932; Drew and Drew, 2004). This creates negative pressure (cavitation) at the

contrecoup site pulling brain towards it. Dawson et al proposed the angular

acceleration theory which is similar to positive pressure theory (Dawson et al., 1980).

This theory is based on the principle of rotation of a tethered object around its center;

the more distant the object is from the center of rotation or more massive the object is

than to which it is tethered, slower will it move. By applying this principle, when the

accelerating head strikes a stationary object, the brain lags resulting injury at the

lagging side (i.e., contrecoup side). Using the principles of Newtonian mechanics and

based on the difference between the relative densities of the brain (0.96 g/L) and the

CSF (1 g/L) Drew and Drew recently proposed an alternate theory. According to this

theory, when the moving head suddenly stops by impact, the CSF being denser than

the brain, continues to move in the direction of the head movement and accumulates at

the site of the impact pushing the brain towards the contrecoup side (Drew and Drew,

2004). This initial displacement of the brain results in more severe contrecoup injury

compared to the coup injury.

1.5.1.1.1 Cerebral Contusion and Laceration

The most common result of coup-contrecoup injury is cerebral contusion, which is

observed in about 20-30% of severe TBI (Khoshyomn and Tranmer, 2004). Cerebral

contusion usually involves the cortex and is associated with damage to the gyri and

small blood vessels often resulting in microhemorrhages (Yokobori and Bullock, 2013).

The pia mater remains intact in contusion differentiating it from laceration in which the

pia mater is torn (see below). The most common areas susceptible to contusion include

25

the orbitofrontal cortex, anterior temporal lobes, and posterior portion of superior

temporal gyrus and are thought to be due to the irregular surface of the bony floor

formed by the frontal and middle cranial fossa on which these structures lie

(Khoshyomn and Tranmer, 2004). The acute phase of contusion is characterized by

inflammatory cascades including vascular migration of polymorphonuclear leukocytes

seen within 24 h of injury, microglial activation seen over 3-5 days leading to release of

proinflammatory cytokines such as tumor necrosis factor–α (TNF-α) and interleukin-1β

(IL-1β) (Holmin et al., 1998; Holmin and Hojeberg, 2004; Clausen et al., 2007),

apoptosis and necrosis (Raghupathi, 2004). This contributes to damage of the blood

brain barrier (BBB) and vasogenic edema resulting in swelling of the areas adjacent to

the contusion site, which is seen in about 30% of all contusions (Khoshyomn and

Tranmer, 2004). Extensive contusions may lead to subdural hematoma and together

they are called a “burst lobe” (Gennarelli and Graham, 2005). In addition to cortical

contusions, shearing forces (often in the sagittal plane) within the white matter can lead

to gliding contusions that occur in the subcortical regions causing damage at the gray-

white matter junction. Gliding contusions tend to be bilateral and are often associated

with DAI caused by differential motion of subcortical structures and acute subdural

hematomas caused by tearing of parasagittal veins (Adams et al., 1986; Sganzerla et

al., 1989).

Laceration is the most severe form of primary TBI resulting in disruption of the brain

parenchyma and involves torn pia-arachnoid membranes. Laceration requires greater

mechanical force compared to cerebral contusions. Penetrating brain injury or

26

depressing skull fractures lead to direct laceration, while parenchymal disruption

secondary to tissue deformation caused by mechanical stress cause indirect lacerations

(Blumbergs, 1997).

1.5.1.1.2 Intracranial Hemorrhage

Intracranial bleeding is a life-threatening complication of TBI and is the most common

cause of TBI fatalities. Intracranial bleeding starts at the time of trauma and can expand

to hematoma in 30-60% of patients with post-TBI coma (Gennarelli and Graham, 2005).

Intracranial bleeding is categorized into four types: epidural hematoma, subdural

hematoma, subarachnoid hemorrhage and intracerebral hemorrhage.

Epidural hematoma/hemorrhage (EDH) is accumulation of blood in the epidural space

(i.e., between the skull and dura) caused by rupture of interposed meningeal vessels

due to shearing stress. Usually EDH occurs due to coup injury and results in separation

of the dura from periosteum. About two-thirds of EDHs are associated with skull fracture

(McKissock et al., 1960). EDH most commonly occurs in the temporoparietal region and

usually involves damage to the middle meningeal artery (Blumbergs, 1997). The

incidence of EDH is about 2-4% of overall TBI and 15% of severe TBI (Bullock et al.,

2006). Compared to adults, EDH is less common in pediatric TBI as the skull is more

pliable reducing occurrence of skull fracture, the dura is firmly attached to the skull and

the meningeal vessels are not yet embedded in the developing skull of infants (Yokobori

and Bullock, 2013). In many cases, patients with EDH have a brief LOC, followed by a

lucid interval and may deteriorate if the symptoms are not readily recognized/treated

27

(‘talk and die’ phenomenon). The common symptoms of EDH include headache,

nausea/vomiting, seizures, and focal neurological deficits including aphagia, weakness,

and numbness. EDH treated with neurosurgical intervention have better prognosis.

Subdural hematoma/hemorrhage (SDH) is bleeding between the dura and arachnoid

membrane resulting from the rupture of bridging veins that cross the subdural space

connecting the cerebral surface with the superior sagittal sinus. SDH occurs in about

5% of TBI and is associated with 65% of cases with prolonged LOC. While EDH are

localized, SDH tend to spread as the fluid can freely move within the subdural space

(Yokobori and Bullock, 2013). SDH is one of the classical triads of shaken baby

syndrome, the other two being retinal hemorrhage and cerebral edema. SDH are

classified into acute, subacute, and chronic depending on the duration of onset and

composition (clotted blood, fluid blood or mixture of clotted and fluid blood) (Blumbergs,

1997; Gennarelli and Graham, 2005). Acute SDH is composed of clotted blood and

forms within first 48 h of trauma and usually is caused by intense mechanical forces.

Acute SDH can be associated with cerebral contusion/laceration and intracerebral

hemorrhage together forming ‘burst lobes’ that cause delayed neurological deterioration

(Blumbergs, 1997). Acute SDH is life-threatening with 30 to 90% mortality rate unless

the patients are operated within the ‘golden hours’ (usually within 4 h) of injury (Seelig et

al., 1981). The poor neurological outcome following acute SDH is associated with

increased intracranial pressure (ICP) that decreases cerebral perfusion pressure (CPP),

which in turn leads to cerebral ischemia (Graham et al., 1978). Subacute SDH is

composed of both clotted and fluid blood that develops over 3-21 days while chronic

28

SDH is composed of fluid blood and develops for > 21 days. Chronic SDH is common in

elderly and alcoholic patients who are more prone to cerebral atrophy and/or

coagulopathy (Blumbergs, 1997). Subacute and chronic SDH have better prognosis

than acute SDH.

Subarachnoid hemorrhage (SAH) is bleeding between the arachnoid membrane and

pia mater and is the most common intracranial hemorrhage. SAH is observed in 40-65%

of moderate-severe TBI and is an independent predictor of poor post-TBI prognosis

(Eisenberg et al., 1990; Kakarieka et al., 1994; Taneda et al., 1996; Servadei et al.,

2002). In most cases, SAH is mild but the severity of SAH increases when associated

with cerebral contusions/lacerations (Blumbergs, 1997).

Intracerebral hemorrhage (ICH) is a hematoma 2 cm or greater in size, not in contact

with the surface of brain, that is caused by rupture of intraparenchymal blood vessels at

the time of impact. ICH is found in 15-20% of autopsy cases of severe TBI (Blumbergs,

1997; Yokobori and Bullock, 2013). Rupture of multiple vessels result in multiple ICH of

which about 28% are associated with SDH and 10% with EDH (Soloniuk et al., 1986).

Small ICH can occur at multiple sites including corpus callosum, the walls of third

ventricle, and basal ganglia. Intraventricular hemorrhage is thought to be caused by

deformation and rupture of subependymal veins due to sudden dilation of ventricular

system during mechanical impact (Zuccarello et al., 1981) or extension of hematomas in

the adjacent regions (Fujitsu et al., 1988).

29

1.5.1.2 Diffuse Brain Injury

1.5.1.2.1 Diffuse Axonal Injury

Diffuse axonal injury (DAI) is the widespread damage to the white matter and is the

most common, most important, and yet the most difficult-to-diagnose TBI pathology.

DAI is the major cause of TBI-induced unconsciousness/coma and vegetative state and

associated with long-term disabilities (Gusmao and Pittella, 2003; Fork et al., 2005).

Although the diagnosis of DAI with traditional imaging techniques such as CT and MRI

has been challenging, recent advances in MRI techniques, such as DTI and

susceptibility weighed imaging may enable the clinicians to capture this ‘stealth

pathology’ even in mild cases (Hunter et al., 2012). DAI was first described by Strich as

“diffuse degeneration of the white matter” observed in post-TBI human brain autopsies

(Strich, 1956).

The principal biomechanical mechanism of DAI is thought to be rapid inertial loading

(e.g., car crash) that induces dynamic shear, tensile and compressive stain in the brain

tissue (Gennarelli et al., 1982; Smith and Meaney, 2000; Smith et al., 2003a). The

axonal damage during TBI is affected by the unique material properties, size, and mass

of the brain tissue, the presence of intracranial folds of dura mater (e.g., falx cerebri)

and the rate and magnitude of strain (Smith and Meaney, 2000; Smith et al., 2003a).

Several studies have shown that nervous tissue is a nonlinear, viscoelastic material

(Donnelly and Medige, 1997; Darvish and Crandall, 2001; Takhounts et al., 2003). In

addition, brain is highly inhomogeneous and anisotropic, thus showing regional

differences in the biomechanical response to the mechanical loading. Thus, the gray

30

matter is isotropic while highly-organized white matter shows anisotropy with the

greatest anisotropy in the corpus callosum (Prange et al., 2000; Prange and Margulies,

2002). When axons are subjected to slowly applied external load (e.g., during normal

head rotation), they behave like a compliant and ductile material and return to their

original geometry after removal of the stress (Franze et al., 2009; van Dommelen et al.,

2009). On the other hand, under rapid and severe external loading (e.g., during car

crash), axons behave like a stiffer and brittle material resulting structural failure. This

property of axons along with their high anisotropic nature is thought to make white

matter more vulnerable to the mechanical trauma (Geddes et al., 1997; Geddes et al.,

2000). The direction of inertial loading and presence of intracranial dural folds my also

affect DAI characteristics. For example, mechanical loading in the temporal region

causes head rotation in the coronal plane. During such rotation, the falx cerebri (the

dural fold that descends vertically between the two cerebral hemispheres) impedes

motion of the following cerebral hemisphere while the leading cerebral hemisphere

continues moving along the direction of motion. This creates excessive shearing forces

along the midline between the two hemispheres. This, along with highly anisotropic

nature of the corpus callosum, may lead to greater axonal damage in this region (Smith

and Meaney, 2000; Smith et al., 2003a).

In extreme but rare TBI cases mechanical forces can cause direct axonal breakage,

called ‘primary axotomy’. In most cases, however axonal pathology develops over time

following mechanical insult (Buki and Povlishock, 2006). Smith and colleagues

conducted a series of elegant studies to understand the mechanism of axonal injury at

31

the cellular level. Using an in vitro axonal stretch injury model, Smith and colleagues

showed that immediately following mechanical loading the axons become temporarily

undulated at multiple sites, which is attributed to partial breakage of microtubules

leading to cytoskeletal damage (Smith et al., 1999a; Tang-Schomer et al., 2010; Tang-

Schomer et al., 2012). The initial axonal stretch results in massive Na+ influx through

mechanosensitive voltage-gated Na+ channels causing membrane depolarization.

Increased intracellular Na+ levels trigger Ca2+ influx, principally via opening of voltage-

gated calcium channels and to a modest level, by reversal of Na+-Ca2+ exchanger (Wolf

et al., 2001; Iwata et al., 2004). The abnormal increase in intracellular Ca2+ leads to

calpain-mediated structural proteolysis causing delayed damage to the axonal

cytoskeleton and ion channels that ensues over days to weeks (Buki et al., 1999a; Buki

et al., 1999b; Iwata et al., 2004; Huh et al., 2006; McGinn et al., 2009). Structural

damage to the axonal cytoskeleton along multiple sites leads to disruption of axonal

transport and subsequent accumulation of axonal proteins causing characteristic axonal

swellings (axonal varicosities or beads) (Tang-Schomer et al., 2012). Two important

proteins that accumulate at the site of axonal swellings are amyloid precursor protein

(APP) and neurofilament (NF), which are transported by fast and slow transport

mechanisms, respectively, and as such are the most commonly-used diagnostic

markers of DAI (Gentleman et al., 1993; Grady et al., 1993; Sherriff et al., 1994a;

McKenzie et al., 1996; Pierce et al., 1996). Accumulation of APP in axonal varicosities

is thought to act as a nidus for the enhanced amyloid beta (Aβ) production, which has

an important implication in the increased risk of Alzheimer disease in TBI survivors

(discussed further). The proposed molecular mechanism of DAI and subsequent Aβ

32

accumulation are depicted in Figure 1.5. The progressive disruption of the cytoskeleton

and subsequent protein accumulation ultimately lead to axonal disconnection, called

‘secondary axotomy’ characterized by formation of axonal bulbs at the terminal end of

axon (Smith and Meaney, 2000; Smith et al., 2003a; Johnson et al., 2013).

Figure 1.5: Proposed molecular mechanism of diffuse axonal injury and generation of Aβ.

Following rapid mechanical stretch (A) the axon becomes temporarily undulated (B) and

relaxes back (C). This causes partial microtubule breakage at multiple sites (B & C).

Axonal stretch also activates mechanosensitive Na+ channels causing massive Na+

influx, which in turn activate voltage gated Ca2+ channels. The excessive Ca2+ influx is

33

thought to exacerbate calpain-mediated disruption of the cytoskeleton (red arrows in E).

Microtubule interrupts protein transport along the axon causing accumulation of

proteins, including APP along with components of APP-cleaving enzymes, BACE1 and

PS1 resulting in axonal varicosities (F). The progressive disruption of cytoskeleton

results in axonal disconnection with formation of axonal bulb (G). The cleavage of APP

accumulated in the axonal varicosities/bulbs by BACE1 and PS1 is thought to lead to a

burst of Aβ production (H). APP: Amyloid Precursor Protein, BACE1: Beta Site APP

Cleaving Enzyme 1, PS1: Presenilin-1, VGCC: Voltage-Gated Ca2+ Channel, VGNC:

Voltage-Gated Na+ Channel. Adapted from Tang-Schomer et al. (2012) and Johnson et

al. (2010)

1.5.1.2.2 Diffuse Vascular Injury

Diffuse vascular injury (DVI) is characterized by multiple macroscopic and/or

microscopic hemorrhagic brain lesions (Tomlinson, 1970; Adams, 1980). Diffuse

vascular injury is almost always fatal with short survival duration (< 24 h) post-trauma

(Pittella and Gusmao, 2003). DVI and DAI are suggested to be associated with each

other and share the same biomechanical mechanism (Pittella and Gusmao, 2003). One

major difference between the two pathologies however, is that while DAI can occur at

low acceleration, DVI is almost exclusively seen in motor vehicle accidents that involve

high-rate inertial forces (Gennarelli and Thibault, 1982; Pittella and Gusmao, 2003).

1.5.2 Secondary Brain Injury

The primary TBI triggers downstream cellular and biochemical cascades that ensue for

days to months, called secondary brain injury, which lead to long-term disabilities. In

most cases the damage caused by the primary brain injury is irreparable however

secondary brain injury is amenable to therapeutic and supportive interventions (Park et

34

al., 2008). The consequences of secondary injury include disruption of

cerebrovasculature including damage to the BBB, alteration in the cerebral blood flow

(CBF), edema, neurometabolic changes including disruption of brain energy

metabolism, inflammation, free radical formation, necrosis and apoptosis. While a

complete discussion of secondary injury pathways is beyond the scope of this thesis,

key secondary injury mechanisms are briefly discussed below.

1.5.2.1 Cerebrovascular Response to TBI

Brain injury causes changes in cerebrovascular response through alteration in BBB

function and CBF that lead to edema, increased intracranial pressure (ICP), cerebral

ischemia and hypoxia. The BBB consists of specialized endothelial cells that form tight

junctions with each other and are surrounded by the basal lamina matrix, pericytes and

astrocyte end feet. This gliovascular unit tightly regulates transport of metabolites,

nutrients and xenobiotics between the brain and the periphery (Abbott et al., 2006). TBI

is often accompanied by disruption of BBB via a variety of mechanisms including,

increased BBB permeability and proinflammatory cytokine activity, loss and/or

redistribution of tight junction proteins and changes in the activity of BBB transporters

(Zink et al., 2010). Changes in BBB function in response to injury are thought to

influence the evolution of injury and response to therapeutic interventions. A

compromised BBB allows entry of plasma-derived, osmotically-active molecules such

as albumin and fibrinogen into the brain interstitial space leading to development of

vasogenic edema (Chodobski et al., 2011).

35

The brain receives about 20% of the cardiac output. CBF is tightly controlled through

constant modulation of vascular resistance and BBB (Zink et al., 2010). TBI decreases

CBF as shown by CT and PET scanning (Bouma et al., 1992; Scalfani et al., 2012). The

reduction in CBF is thought to occur due to dysfunction of both macro- and

microvasculature leading to focal or in severe cases global ischemia (Bouzat et al.,

2013).

1.5.2.2 Neurometabolic Changes Following TBI

TBI leads to perturbations of cell membrane leading to redistribution of ions and

neurotransmitters. During the acute phase of injury, stretching of axonal membrane

opens voltage-dependent K+ channels leading to massive K+ efflux. This has been

shown with increased extracellular K+ concentration measured by in vivo microdialysis

in various animal models of TBI (Takahashi et al., 1981; Hubschmann and Kornhauser,

1983; Katayama et al., 1990). The duration of K+ increase is shown to be dependent on

the injury dose with shorter duration following mTBI whereas moderate-severe TBI

cause prolonged K+ increase (Katayama et al., 1990). The former is associated with

increased neuronal firing while the latter coincides with indiscriminate glutamate release

(Katayama et al., 1990). In addition to K+, excessive glutamate also leads to

uncontrolled shifts in Na+ and Ca2+ homeostasis (Bullock, 1994; Bullock et al., 1998). To

compensate for these ionic alterations, the ion pumps, especially Na+/K+-ATPase go

into the overdrive mode further increasing metabolic demand setting a vicious cycle of

CBF-metabolism uncoupling to the cell (Werner and Engelhard, 2007).

36

1.5.2.3 Neuroinflammation Following TBI

Neuroinflammation is an important secondary injury mechanism that contributes to

ongoing impairment following TBI. Microglia, the resident macrophages in the central

nervous system (CNS) act as the “first responders” mediating the immune response to

the brain injury. Microglial activation has been demonstrated as early as 72 hours after

human TBI and can persist for years after injury (Engel et al., 2000; Beschorner et al.,

2002; Gentleman et al., 2004; Ramlackhansingh et al., 2011). This activation profile is

also mirrored in animal models of TBI (Csuka et al., 2000; Koshinaga et al., 2000;

Maeda et al., 2007). Studies have shown that under physiological and pathological

conditions microglia assume different activation states that have distinct morphological

characteristics (Boche et al., 2013). At resting state, microglia show ramified

morphology characterized by short, fine processes and are thought to constantly

monitor their local environment. In response to an acute insult, such as trauma, resting

microglia are activated with dramatic changes in morphology and behavior. The cells

retract their processes and assume amoeboid morphology resembling macrophages

that proliferate and migrate towards the site of injury (Loane and Byrnes, 2010; Boche

et al., 2013). In certain chronic infections, microglia can assume rod morphology

characterized by markedly elongated nuclei, scanty cytoplasm and few processes.

Recently, rod microglia were observed in the brain regions that have an associated

sensory sensitivity following diffuse brain injury in rats (Ziebell et al., 2012) although

their significance in TBI pathology is currently unknown. Classically activated microglia,

also called M1 microglia release proinflammatory mediators, including cytokines,

chemokines, nitric oxide and superoxide free radicals that contribute to neuronal

37

dysfunction and cell death in response to injury (Loane and Byrnes, 2010). Both human

and animal studies have shown rapid elevation of proinflammatory cytokine IL-1β within

hours following TBI (Woodroofe et al., 1991; Fan et al., 1995; Winter et al., 2002), which

activates proinflammatory pathways such as TNF-α (Rothwell, 2003) through interleukin

1 receptor type 1 expressed in microglia and neurons (Pinteaux et al., 2002; Lu et al.,

2005). Recent studies have identified an alternatively activated state of microglia, called

M2 microglia (Colton, 2009; Kigerl et al., 2009) that are considered to have anti-

inflammatory actions and secrete anti-inflammatory cytokines including IL-10 and

neurotrophic factors such as nerve growth factor and transforming growth factor β

following injury (Csuka et al., 1999).

1.6 TBI, Alzheimer Disease, and Apolipoprotein E

1.6.1 TBI and Alzheimer Disease

Research over last 30 years indicates an increased risk of dementia, particularly of

Alzheimer’s disease (AD) in TBI survivors (Van Den Heuvel et al., 2007; May et al.,

2011). Moreover, besides age, antecedent TBI is now considered as one of the most

important environmental risk factors for dementia. Several epidemiological studies have

indicated increased dementia risk in TBI victims (French et al., 1985; Mortimer et al.,

1985; Mortimer et al., 1991; van Duijn et al., 1992; Rasmusson et al., 1995; Guo et al.,

2000; Plassman et al., 2000; Fleminger et al., 2003; Barnes et al., 2014), although this

association is not always observed (Katzman et al., 1989; Li et al., 1992; Fratiglioni et

al., 1993). One potential reason for the disagreement between the epidemiological

studies is that the data collection in most of these studies was retrospective, based on

38

low sample number and thus may have suffered from recall bias (May et al., 2011).

Nonetheless, a high-powered EURODEM meta-analysis pooled the findings of 11 case-

control studies and reported a significant relative risk of 1.82 (95% CI: 1.26-2.67) for

head trauma associated with AD (Mortimer et al., 1991). Based on the available

epidemiological data, the Committee on Gulf War and Health: Brain Injury in Veterans

and Long-Term Health Outcomes recently concluded that “there is sufficient evidence of

association between moderate to severe TBI and AD, limited/suggestive evidence of

association between mild TBI with LOC and AD and inadequate evidence of association

between mild TBI without LOC and AD” (Rutherford et al., 2008). TBI and AD show

striking similarities in their pathologies, which may shed light into their association. In

the following sub-sections I have discussed similarities and differences between TBI

and AD pathologies.

1.6.2 Alzheimer’s Disease: Pathology

AD is defined by two neuropathological hallmarks: 1) extraneuronal amyloid plaques

mainly composed by aggregated amyloid beta (Aβ) peptides and 2) intraneuronal

neurofibrillary tangles (NFT) formed by hyperphosphorylated tau protein. Central to the

AD pathology is the widely accepted, yet controversial amyloid cascade hypothesis

according to which the deposition of Aβ in the brain parenchyma is a crucial step that

triggers formation of intraneuronal tau aggregates and ultimately leads to AD (Hardy

and Higgins, 1992; Selkoe, 1994; Hardy and Selkoe, 2002; Karran et al., 2011). The

amyloid cascade hypothesis was developed on the basis of the discovery of autosomal

dominant mutations in APP, the parent protein of Aβ peptides and Presenilin-1 (PS1)

39

and Presinilin-2 (PS2), the components of γ-secretase (an enzyme involved in cleavage

of APP that leads to Aβ generation) that cause Familial/Early Onset AD (EOAD)

(Chartier-Harlin et al., 1991; Goate et al., 1991; Citron et al., 1992; Levy-Lahad et al.,

1995; Rogaev et al., 1995; Scheuner et al., 1996). On the other hand, apolipoprotein E

(apoE), specifically the apoE ε4 allele is an established genetic risk factor for the

sporadic or Late Onset AD (LOAD) and is thought to affect several aspects of AD

pathology, including Aβ aggregation and clearance, tau phosphorylation, inflammation,

neuronal repair and synaptic plasticity (discussed further).

1.6.2.1 Amyloid Beta Formation in the CNS

Aβ peptides are physiologically produced by the sequential enzymatic processing of the

highly-conserved type I transmembrane glycoprotein, APP. APP is a single

transmembrane domain protein that consists of a large extracellular domain, a

hydrophobic transmembrane domain and a short intracellular C terminal tail (Suh and

Checler, 2002). The juxtamembrane region of APP contains the Aβ sequence

composed of 43 amino acids (amino acids 597-639 of APP). APP is highly expressed in

neurons (Slunt et al., 1994; Lorent et al., 1995), undergoes fast anterograde axonal

transport (Koo et al., 1990; Buxbaum et al., 1998) and is processed via two mutually-

exclusive pathways: anti-amyloidogenic and amyloidogenic (Figure 1.6). The majority of

APP is processed through the anti-amyloidogenic pathway that involves enzymatic

cleavage, first by α-secretase followed by γ-secretase (Suh and Checler, 2002). Alpha-

secretase, a membrane bound, zinc metalloproteinase of the ADAM family, first cleaves

the APP within the Aβ sequence between residues Lys16 and Leu17 releasing a large

40

secreted α APP (sAPPα) domain and a truncated, 83-residue carboxyl terminal

fragment (CTFα or C83) that lacks the N-terminal portion of Aβ (Naslund et al., 1994). In

the amyloidogenic pathway instead of α-secretase APP is cleaved by β-secretase,

which is the rate-limiting step in the formation of Aβ species (Vassar, 2004). The

neuronal form of β-secretase, also known as β-Site APP Cleaving Enzyme 1 (BASE1),

a membrane-bound aspartyl protease cleaves the extracellular domain of APP

generating sAPPβ and a 99-residue CTFβ or C99 that contains the N-terminal domain

of Aβ. Both CTFα and CTFβ are substrates of γ-secretase, which is a complex of four

subunits: PS-1, PS-2, nicastrin, APH-1 and PEN-2 (Iwatsubo, 2004). Gamma-secretase

cleaves CTFα in the transmembrane domain releasing a 3 kDa, N-terminal-truncated

Aβ peptide, called P3 in the extracellular fluid and a cytosolic amyloid intracellular

domain (AICD). Cleavage of CTFβ by γ-secretase results in the release of 4 kDa Aβ in

the extracellular fluid and cytosolic AICD. Gamma-secretase can cut CTFβ at multiple

sites; first in the ε site generating 48 or 49-residue Aβ and subsequently every 3-4

amino acid residues generating 39-43 residue Aβ species (Kakuda et al., 2006; Selkoe

and Wolfe, 2007). The longer residue Aβ species tend to aggregate more readily and

are toxic. Under physiological conditions, γ-secretase cleavage generates less-toxic

Aβ40 and more-toxic and fibrillogenic Aβ42 species that account for 90% and <10% of

secreted Aβ species, respectively (Thinakaran and Koo, 2008). While Aβ42 is the major

component of amyloid plaques, Aβ40 accounts for most of the amyloid deposited in

cerebral vessels, which is called cerebral amyloid angiopathy (CAA) and is observed in

80% of AD patients (Biffi and Greenberg, 2011; Masters and Selkoe, 2012).

41

Figure 1.6: Amyloidogenic and anti-amyloidogenic processing of APP. The anti-amyloidogenic processing of APP involves sequential cleavage of APP by first

α-secretase that generates sAPPα and CTFα followed by cleavage of CTFα by γ-

secretase generating AICD and truncated Aβ fragment, P3. The amyloidogenic pathway

involves cleavage of APP by first β-secretase that generates CTFβ and sAPPβ followed

by cleavage of CTFβ generating Aβ and AICD.

1.6.2.2 Tauopathy in Alzheimer Disease

The second pathological hallmark of AD is intraneuronal neurofibrillary tangles (NFT),

which are composed of abnormally hyperphosphorylated tau (p-tau) protein that forms

paired helical filaments (PHF) (Iqbal et al., 2010). According to the amyloid hypothesis,

while amyloid plaque formation is the initiating event of AD pathology, formation of NFT

is a downstream event and executes neuronal death (Hardy and Selkoe, 2002; Karran

42

et al., 2011). Tau is the major microtubule associated protein (MAP), expressed almost

exclusively in neurons and is mostly localized in axons (Binder et al., 1985).

Microtubules are non-covalent cytoskeletal polymers composed of α- and β-tubulins that

play an important role in trafficking of organelles and other substances along axons

(Conde and Caceres, 2009). Humans express six tau isoforms including three of each

of three-repeat (0N3R, 1N3R, and 2N3R) and four-repeat (0N4R, 1N4R, and 2N4R) tau,

which are generated by alternative splicing of the MAP tau (MAPT) gene on

chromosome 17q21.31 (Neve et al., 1986; Goedert et al., 1989; Andreadis et al., 1992).

In contrast to humans, the adult mouse brain expresses only four-repeat isoforms

(0N4R, 1N4R, and 2N4R) (Gotz et al., 2010). The two important known functions of tau

are 1) facilitation of microtubule assembly and 2) microtubule stabilization by reducing

microtubule dynamics through its interaction with tubulin (Weingarten et al., 1975; Bré

and Karsenti, 1990). The interaction of tau with tubulin occurs at the C-terminus of tau

protein that contains microtubule binding repeat (MTBR) domain composed of three or

four tandem repeat motifs (Lee et al., 1989). The binding of tau to tubulin is regulated

and is inversely correlated to the degree of tau phosphorylation, especially at Ser262

and Thr231 residues within and flanking the MTBR domain (Wang et al., 2007). Thus,

phosphorylation of Ser262 within the flanking region of MTBR significantly reduces

binding of tau with microtubules (Biernat et al., 1993). Phosphorylated tau under

physiological conditions contains 2-3 moles of phosphate per tau molecule, while

hyperphosphorylated tau associated with AD is defined to contain > 6 moles of

phosphate per tau molecule (Iqbal et al., 2010). Abnormally p-tau is incapable of binding

to tubulin, thus remains free and polymerizes into PHF, which negatively affects

43

microtubule assembly and stabilization (Alonso et al., 2006). While tau can be

phosphorylated by a variety of kinases (e.g., glycogen synthase kinase, cdk5, MAP

kinase, JNK, and cdc2), glycogen synthase kinase 3β (GSK3β) plays the most

important role in regulating tau phosphorylation under physiological and pathological

conditions (Avila et al., 2004). The principle enzyme that controls tau dephosphorylation

is protein phosphatase 2A (PP2A). The phosphorylation state of tau is regulated by the

balance between the activities of protein kinases and phosphatases. For example,

PP2A levels in AD brains are significantly downregulated (Sontag et al., 2004a; Sontag

et al., 2004b). On the other hand, Pei et al found approximately 50% increase although

without any increase in the activity of GSK3β levels in AD brains (Pei et al., 1997).

While Aβ plaque pathology is seen exclusively in AD, tauopathy besides AD is also

associated with other neurodegenerative disorders such as Frontotemporal Dementia

and Parkinson’s Disease linked to chromosome 17, Pick’s Disease, Corticobasal

Degeneration, Progressive Supranuclear Palsy, Argyrophilic Grain Disease,

Amyotrophic Lateral Sclerosis and Parkinsonism-Dementia Complex (Spillantini and

Goedert, 2013) and Chronic Traumatic Encephalopathy (CTE) (McKee et al., 2009). A

distinct feature of tauopathy seen in AD and CTE compared to other tauopathy-related

neurodegenerative disorders is that the NFT in AD and CTE are composed of all six

abnormally p-tau isoforms (Goedert et al., 1992; Schmidt et al., 2001). On the other

hand, the tauopathy in other neurodegenerative disorders involves abnormal

hyperphosphorylation of either four-repeat (e.g., Progressive Supranuclear Palsy,

Corticobasal Degeneration and Argyrophilic Grain Disease) or three-repeat (e.g., Pick’s

44

Disease) tau (Flament et al., 1991; Ksiezak-Reding et al., 1994; Delacourte et al., 1996;

Spillantini et al., 1997; Tolnay et al., 2002).

1.6.3 TBI and Alzheimer’s Pathology: Similarities and Differences

1.6.3.1 Amyloid Pathology after Moderate to Severe Single TBI

One of the first studies linking TBI to dementia was by Rudelli et al who reported

classical AD pathology seen in the brain autopsy of a 38-year old man who died 16

years following recovery after a single severe head trauma (Rudelli et al., 1982). In

subsequent brain autopsy studies it was reported that about 30% of fatal TBI cases had

cortical Aβ deposition, strikingly similar to AD plaques (Roberts et al., 1991; Roberts et

al., 1994; Gentleman et al., 1997). Moreover, the Aβ deposits seen in severe TBI cases

were predominantly composed of insoluble Aβ42 (Gentleman et al., 1997; DeKosky et

al., 2007). There are however, important differences in the plaque pathology found in

the severe head trauma cases and AD. First, Aβ deposition in these acute TBI cases

was observed irrespective of the patient’s age, i.e., even in children, which is in contrast

to AD in which the plaque pathology developed in advanced age. Second, severe head

trauma triggered rapid Aβ deposition, seen as early as 2 h of TBI, as opposed to AD

pathology that develops over years (Ikonomovic et al., 2004). Third, TBI-associated Aβ

deposits are not restricted to the gray matter but have also been reported in the white

matter (Smith et al., 2003b). Finally, the plaques formed following TBI tend to be diffuse

compared to solid, dense-core plaques seen in AD brains. In contrast to the above

studies, Adle-Biasette et al found no evidence of Aβ pathology in 23 cases of age

45

between 17 and 63 years and 17 cases of age between 69 and 79 years who died

between 0-76 days after head trauma (Adle-Biassette et al., 1996).

The mechanisms of rapid Aβ deposition following TBI are unclear. A potential source of

Aβ is postulated to be the axonal swellings (varicosities) formed by DAI (Johnson et al.,

2010). In vitro studies have shown that axonal stretch at high shear rate and intensity

induces microtubule breakage at multiple sites along the length of the axon resulting in

interruption of axonal transport of various proteins, including APP (as discussed in

section 1.5.1.2.1). Intra-axonal accumulation of APP following TBI has been reported in

humans (Gentleman et al., 1993; Sherriff et al., 1994a; Gorrie et al., 2002) and animal

models (Lewen et al., 1995; Pierce et al., 1996; Smith et al., 1999b). Interestingly, the

components of APP cleaving enzymes, BASE1 and PS1 have also been shown to co-

accumulate along with APP in the damaged axons in humans (Chen et al., 2004; Uryu

et al., 2007) and animals (Chen et al., 2004). The axonal transport of these proteins is

shown to be mediated by the axonal transport protein, kinesin-I (Kamal et al., 2000) and

APP is thought to act as kinesin-I receptor facilitating transport of BACE-1 and PS1

(Kamal et al., 2001). Based on these observations it is proposed that the co-

accumulation of APP, BASE-1 and PS1 within the axonal varicosities form a nidus for

Aβ production (Johnson et al., 2010). Supporting this hypothesis, intra-axonal

accumulation of Aβ has been reported following TBI (Chen et al., 2004; Uryu et al.,

2007; Chen et al., 2009). Thus, DAI seems to create a unique environment bringing

together all the machinery necessary for Aβ production (i.e., APP, BACE-1 and PS1),

which is thought to lead to a burst of Aβ production that can deposit as amyloid plaques

46

(Johnson et al., 2010). To counteract the increased Aβ production, the traumatized

brain may also upregulate Aβ clearance mechanisms. For example, Chen et al reported

increased intra-axonal accumulation of neprilysin (NEP), a major Aβ degrading enzyme,

following TBI (Chen et al., 2009). In support of this observation, Johnson et al reported

that a genetic polymorphism in NEP that resulted into longer GT repeats in the enzyme

was associated with increased risk of Aβ deposition after TBI while shorter GT repeats

NEP decreased the incidence of Aβ deposition (Johnson et al., 2009). Thus, a balance

between Aβ production and clearance after TBI may be crucial in preventing Aβ –

related pathologies in TBI. In this regard, apoE may also play an important role as it is

shown to modulate Aβ clearance mechanisms (discussed further).

While the majority of clinical studies indicate rapid plaque pathology in the acute phase

of TBI, the Aβ plaque dynamics in the long-term TBI survivors seems controversial.

Chen and colleagues studied 23 cases with post-TBI survival up to 3 years (mean

survival 245 days) and found widespread axonal pathology along with intra-axonal

accumulation of APP, BASE1, PS1 and Aβ over the long term. The authors also

reported NEP immunoreactivity in axons for up to 8 month post-injury. Surprisingly, the

authors did not find any evidence of Aβ plaques in the same cases (Chen et al., 2009).

The authors suggested that while Aβ plaques are seen in short-term post-TBI survivors,

plaques may undergo NEP-mediated regression over months-to-years post-injury. In

contrast to the above study, in a recent study by the same group, widespread tau and

amyloid pathology was found in 39 single moderate to severe TBI cases with survival up

to 47 years (Johnson et al., 2012). It should be noted however that Chen et al have not

47

reported on any tauopathy as reported by Johnson et al in the absence of Aβ

deposition. This may be important for two reasons. First, while amyloid hypothesis is

widely accepted as the basis of AD pathology, it is not without controversies. For

example, few studies have found no correlation between amyloid plaque load and the

degree of cognitive impairment, i.e., some subjects with substantial amyloid plaques do

not exhibit clinical signs of AD (Armstrong, 1994; Giannakopoulos et al., 2003). Second,

in most of the studies mentioned above, Aβ plaque deposits were found only in one-

third of the cases. Thus, as suggested for AD pathogenesis (Hardy and Selkoe, 2002;

Karran et al., 2011), while Aβ plaque pathology may be a triggering event, tauopathy

may be the executioner of the synaptic and neuronal loss and long-term cognitive

deficits in TBI survivors. This is underlined by tauopathy which is the principal pathology

of CTE, a progressive neurodegenerative disorder thought to be caused by repeated

head trauma (see the next section). Moreover, in a recent study it was shown that

inhibition of APP processing by treating 3XTg-AD mice with a γ-secretase inhibitor

blocked Aβ accumulation without affecting tauopathy following controlled cortical impact

injury suggesting that Aβ accumulation and tauopathy in TBI may be independent

pathways (Tran et al., 2011a).

While the data from post-mortem brain analyses indicate increased Aβ production

following TBI, Aβ and tau levels have also been measured in CSF and recently in brain

interstitial fluid (ISF) of TBI patients and have revealed intriguing changes in Aβ

dynamics. For example, two studies reported an initial increase the CSF Aβ42 levels

peaking at ~ 5 days after severe TBI (GCS <8) followed by decline over several weeks

48

(Raby et al., 1998; Olsson et al., 2004). In both these studies, CSF tau levels were not

measured. In contrast to these findings, Franz et al found a significant decrease in CSF

Aβ42 measured over 1-30 days in severe TBI cases (Franz et al., 2003). Tau levels in

the same patients were significantly increased peaking between 5-15 days followed by

a decline over 30 days (Franz et al., 2003). Marklund et al assessed total tau and Aβ42

levels in brain ISF collected by real-time in situ microdialysis from eight patients with

either focal/mixed TBI or DAI over 8 days following TBI and found a non-significant

increase in Aβ42 levels in patients with DAI compared to patients with focal/mixed TBI.

Brody et al measured Aβ levels in ISF sample obtained from 18 TBI patients from 12-48

h and observed some very intriguing trends in Aβ dynamics (Brody et al., 2008). The

authors found that the ISF Aβ levels correlated positively with the patient’s neurological

status as determined by GCS such that ISF Aβ levels declined in patients with worse

neurological status, increased as the patient’s neurological status improved and

remained stable in clinically stable patients (Brody et al., 2008). In their follow-up study,

Brody and colleagues found > 3-fold increase in ISF tau levels at 1-12 h followed by a

decline over 61-72 h following TBI (Magnoni et al., 2012). The authors further found that

the high initial tau levels were correlated with worse neurological outcomes and

inversely correlated with initial low ISF Aβ levels. Based on these surprising findings,

the authors hypothesized that low Aβ and high tau levels during the acute post-TBI

phase may be associated with reduced synaptic activity (Magnoni and Brody, 2010) as

ISF Aβ levels are directly regulated by synaptic activity (Cirrito et al., 2005). Brain ISF

Aβ dynamics was also replicated in a controlled cortical impact (CCI) model in AD as

well as wild-type mice wherein both human as well as murine ISF Aβ levels were

49

significantly reduced following TBI (Schwetye et al., 2010). The authors further found

that in concordance with ISF, Aβ levels in phosphate buffered saline (PBS)-soluble

hippocampal and cortical lysates were also reduced. The pathophysiological

implications of differential Aβ dynamics following brain trauma in the short-term vs long-

term effects of TBI are unclear and require further investigation.

1.6.3.2 Consequences of Repetitive TBI: Chronic Traumatic Encephalopathy

While the data on long-term consequences of moderate to severe single TBI is limited, it

is long known that mild, repetitive head trauma can increase the risk of late-life

dementia. In 1928, Harrison Martland first described “Punch Drunk” syndrome among

boxers who sustained multiple head blows with early symptoms characterized by gait

deficits with slow muscular movements, hand tremors and speech hesitancy (Martland,

1928). In the later stages, the fighters with severe head blows suffered from

Parkinsonian-like facial tremors, and vertigo ultimately resulting in mental deterioration.

Millispaugh coined the term ‘dementia pugilistica’ (pugilist: a professional boxer) to the

symptoms described by Martland in boxers (Millspaugh, 1937). The first detailed

pathological features of dementia pugilistica were described by Brandenburg and

Hallervorden (Brandenburg and Hallervorden, 1954) and later by Corsellis et al

(Corsellis et al., 1973). While the earlier studies on the aftermath of repetitive head

impacts were focused on boxing, studies in the last decade showed that the clinical

symptoms and neuropathology described as dementia pugilistica were also seen in

other situations that involved increased risk of repetitive head impacts including contact

sports such as American football, wrestling, hockey and association football (soccer)

50

(Cantu, 2007; McKee et al., 2009) and in military veterans exposed to blast injuries

(Omalu et al., 2011; Goldstein et al., 2012). Thus the currently-accepted term for

dementia pugilistica is chronic traumatic encephalopathy (CTE).

Presently no biomarkers are available for the diagnosis of CTE. The current

neuropathological data comes from the post-mortem analysis of athlete brains, while the

clinical features of CTE are mostly based on the interviews of friends and family

members. The clinical manifestation of CTE occurs several years after exposure to

repetitive TBI and involves cognitive impairments, neuropsychiatric alteration, and motor

deficits that are thought to occur over three stages (Chin et al., 2011). Stage 1 is

characterized by affective disturbances and psychotic symptoms (paranoia, agitation)

and changes in mood and behavior (depression, aggression or violence, irritability,

suicidal tendency). Stage 2 involves social instability and memory deficits, with initial

symptoms of Parkinsonism. In the last stage the patient manifests general cognitive

dysfunction and progression of dementia, speech or gait deficits or full-blown

Parkinsonism.

The macroscopic neuropathology of CTE is characterized by marked atrophy of

cerebral regions (frontal and temporal regions) (Neubuerger et al., 1959; Corsellis et al.,

1973; McKee et al., 2009) and general atrophy of the cerebellum (Williams and

Tannenberg, 1996). Another gross neuropathological feature of CTE is cavum septum

pellucidum (separation of leaflets of septum pellucidum) with a fenestrated septum

(Omalu et al., 2006; McKee et al., 2009; Stern et al., 2011; McKee et al., 2013).

51

Although many of the clinical features of CTE overlap with those of AD, the microscopic

pathology of CTE shows some features that distinguish CTE from AD (Table 1.5). Thus,

the primary histological feature of CTE is extensive deposition of p-tau as NFT such that

CTE is now often considered as a tauopathy (McKee et al., 2009; McKee et al., 2013).

A second distinctive histopathologic feature of CTE is presence of TAR DNA-binding

protein 43 (TDP-43) with TDP-43 positive neurites in the cortex, brain stem, and

temporal lobe seen in early stages and in frontal subcortical white matter and fornix,

brainstem and medial temporal lobes in later stages (McKee et al., 2010; McKee et al.,

2013). TDP-43 pathology is seen in about 80% of CTE cases (Baugh et al., 2014) as

opposed to a lower incidence (25-30%) in AD (Wilson et al., 2011). Lastly, while amyloid

deposition is the neuropathological signature of AD, the presence of Aβ deposition is an

inconsistent feature of CTE and is reported in only 25-50% of CTE cases (McKee et al.,

2009; McKee et al., 2013).

52

Chronic Traumatic Encephalopathy Alzheimer Disease Prominent presence of tau tangles in the astrocytes

Astrocyte tau tangles absent

Perivascular deposition of NFT Perivascular NFT absent Prominent NFT deposition at depths of cerebral sulci

No NFT deposition at depths of cerebral sulci

Prominent subpial astrocytic tangles Absence of subpial astrocytic tangles

Periventricular astrocytic tangles Absence of periventricular astrocytic tangles

Focal NFT seen in the frontal lobe cerebral cortex in early stages

NFT seen in entorhinal cortex, amygdala and hippocampus

Patchy and irregular NFT distribution in advanced stages

Uniform NFT distribution in advanced stages

Tau pathology seen in white matter Virtual absence of tau pathology in white matter

Amyloid deposits seen only in 25-50% of CTE cases

Amyloid deposits is the signature pathology

High incidence (~ 80%) of TDP-43 pathology

TDP-43 pathology seen in 25-30% of cases

Table 1.5: Differences in the pathological features of CTE and AD. Adapted from (McKee et al., 2013).

Recently McKee and colleagues proposed four stages of CTE neuropathology based on

the severity of pathological and clinical features recorded from 68 cases with confirmed

CTE (McKee et al., 2013). The four stages of CTE are briefly described in the Table 1.6.

53

CTE Stage

Neuropathological Features Associated Clinical Features

I

• Mild lateral ventricle enlargement seen in some cases

• Focal perivascular NFT at depth of cortical sulci

• Headache • Loss of attention and

concentration

II

• Mild enlargement of frontal horn of lateral and 3rd ventricles

• Small cavum septum pellucidum • NFT in superficial cortical layers

adjacent to focal epicenters and in nucleus basalis of Meynert and locus ceruleus

• Depression and mood swings • Explosivity • Headache • Loss of attention and

concentration • Short-term memory loss

III

• Mild cerebral atrophy, ventricular dilation, septal abnormalities, a sharply concave contour of 3rd ventricle, depigmentation of locus ceruleus and substantia nigra

• Dense p-tau pathology in hippocampus, entorhinal cortex, amygdala and widespread in cortex, diencephalon, brainstem and spinal cord

• Cognitive impairment • Memory loss • Executive dysfunction • Depression • Explosivity • Loss of attention and

concentration • Visuospatial abnormalities

IV

• Atrophy of medial temporal lobe, cerebrum, hypothalamus, thalamus, mammillary body, septal abnormalities, ventricular dilation and pallor of locus ceruleus and substantial nigra

• Widespread p-tau pathology including white matter

• Prominent neuronal loss and gliosis in cortex

• Hippocampal sclerosis

• Dementia with profound short-term memory loss

• Executive dysfunction • Loss of attention and

concentration • Explosivity and aggression • Paranoia, depression, impulsivity • Visuospatial abnormalities

Table 1.6: Pathological and clinical stages of CTE.

Adapted from McKee et al. (2013).

54

1.6.4 Apolipoprotein E (ApoE) at the Nexus of TBI and AD Pathology

Familial or early onset AD is caused by mutations in APP, PS1 and PS2 genes and

represents less than 10% of total AD cases. The majority of the AD cases has no known

causative genetic mutations and is thus called sporadic or late onset AD (LOAD). While

the exact cause of LOAD is currently unknown it is believed that the development and

progression of AD is affected by multiple environmental and genetic risk factors. For

example, advanced aging is considered the greatest environmental risk factor. On the

other hand, apolipoprotein E (apoE), particularly the ε4 (apoE4) allele is the most

established genetic risk factor for LOAD (Corder et al., 1993; Saunders et al., 1993).

Interestingly, apoE4 is also thought to affect recovery from TBI (discussed further).

The CNS is the most lipid-rich tissue in the body, containing over 25% of total body

cholesterol in only 2% of total body weight (Dietschy and Turley, 2001; Vance et al.,

2005). Almost all of the CNS cholesterol is in the unesterified form and about 70% of

this form is present in oligodendrocytes that make myelin sheaths, which accounts for ~

50% of the white matter (Bjorkhem and Meaney, 2004). Being insoluble in water,

cholesterol is transported on lipoprotein particles that consist of apolipoproteins that

surround, stabilize, and solubilize their fatty cargo. The CNS lipid metabolism has some

unique features that separate it from the peripheral lipid metabolism. For example,

because of the presence of the BBB, there is virtually no exchange of cholesterol and

lipoproteins between the CNS and periphery and virtually all cholesterol is synthesized

in situ (Dietschy and Turley, 2001). This observation is supported by many studies that

showed virtual absence of radioactivity in the CNS following peripheral administration of

55

radiolabeled cholesterol (Chobanian and Hollander, 1962; Plotz et al., 1968; Meaney et

al., 2001). All brain cells synthesize cholesterol during embryonic development, but

mature neurons do not produce sufficient cholesterol to meet their lifelong ongoing

requirements for membrane synthesis and repair (Dietschy and Turley, 2004). Adult

neurons therefore depend upon lipids transported on glial-derived lipoproteins for

optimal survival and maintenance. The CNS lipoprotein metabolism is based entirely

upon particles that resemble plasma high-density lipoproteins (HDL) in density, size,

and composition (LaDu et al., 2000b). A major difference between brain and plasma

HDL is that apoE is the major apoprotein in brain HDL whereas apoA-I is the primary

apoprotein of plasma HDL (LaDu et al., 2000b).

1.6.4.1 ApoE Synthesis in the CNS

Under physiological conditions, apoE in the CNS is synthesized by primarily by

astrocytes (DeMattos et al., 2001) and to a lesser extent by microglia (Gong et al.,

2002). ApoE mRNA expression has been reported in cortical and hippocampal neurons

but not in the cerebellar neurons (Xu et al., 1999). Some studies suggest that neurons

can express apoE under certain conditions like excitotoxic stress (Xu et al., 2006) and is

controlled by astrocytes through Erk kinase pathway (Harris et al., 2004). Compared to

the peripheral lipoprotein synthesis and metabolism, little is known about CNS

lipoprotein metabolism. In the periphery, liver secretes lipid-poor apoA-I containing

lipoprotein particles that acquire free cholesterol and phospholipid from the peripheral

tissue (Yu et al., 2011). This initial transfer of free cholesterol and phospholipid is

mediated by the membrane-bound cholesterol transporter, Adenosine Triphosphate

56

Binding Cassette A1 (ABCA1) (Oram and Vaughan, 2000). ABCA1-mediated transport

of cholesterol is the critical and rate-limiting step in the biogenesis of HDL (Bodzioch et

al., 1999; Brooks-Wilson et al., 1999). The partially-lipidated apoA-I particles further

acquire more lipids from another ABC transporter, ABCG1 forming nascent HDL3

(Nakamura et al., 2004; Kennedy et al., 2005). The HDL3 particles are matured by

further acquiring phospholipids, free cholesterol, and triglycerides via lecithin-cholesterol

acyltransferase (LCAT), an enzyme that esterifies free cholesterol resulting in spherical

HDL2.

Based on the pathway of peripheral HDL biosynthesis, it is proposed that glia secrete

apoE-containing lipoproteins as lipid-poor nascent discoid particles that are composed

of phospholipids and free cholesterol but lack a cholesterol ester/triglyceride core (LaDu

et al., 1998; Fagan et al., 1999). The nascent apoE particles are remodeled into mature

lipoproteins in the brain parenchyma by acquiring cholesterol and phospholipids (LaDu

et al., 1998; Koch et al., 2001). We and others have shown that the cholesterol

transporter, ABCA1 is expressed by both glia and neurons (Wellington et al., 2002;

Hirsch-Reinshagen et al., 2004; Tachikawa et al., 2005) and is required for the lipidation

of apoE in the brain (Hirsch-Reinshagen et al., 2004; Wahrle et al., 2004). The ABCA1-

mediated lipidation of apoE is regulated by nuclear transcription factors, liver X

receptors (LXR).

57

1.6.4.2 Liver X Receptors (LXR)

LXR have two isoforms in mammals: LXRα and LXRβ, which share 80% sequence

homology and have differential expression patterns (Wojcicka et al., 2007). LXRα is

expressed principally in the liver, intestine, kidney, spleen, and adipose tissue (Auboeuf

et al., 1997) whereas LXRβ is ubiquitously expressed albeit at lower levels (Song et al.,

1994). LXRs were initially identified as orphan receptors, i.e., there were no known

endogenous ligands at the time of their discovery. Later studies identified oxysterols as

their endogenous ligands (Janowski et al., 1996; Janowski et al., 1999). Oxysterols are

monooxygeneted derivatives of cholesterol that are generated as intermediates in the

cholesterol biosynthetic pathway [24(S),25-epoxycholesterol], steroid hormone

biosynthesis [22(R)-hydroxycholesterol], and cholesterol metabolism pathways [24(S)-

hydroxycholesterol, 25-hydroxycholesterol, and 27-hydroxycholesterol] (Schroepfer,

2000). 24(S)-hydroxycholesterol (cerebrosterol) is the principal cholesterol metabolite

and mechanism through which cholesterol is removed from the CNS (Lund et al., 1999)

whereas 27-hydroxycholesterol is the most abundant cholesterol metabolite in the

periphery (Wojcicka et al., 2007). Most oxysterols have similar affinity towards both LXR

isoforms. Various synthetic LXR agonists have been developed; the two most

commonly used LXR agonists in experimental studies are T0901317 and GW3965.

T0901317 activates both LXR isoforms with almost equal affinity (Schultz et al., 2000)

as well as pregnane X receptor (Mitro et al., 2007) and farnesoid X receptor (Houck et

al., 2004). GW3965 (2-[3-[3-[[2-chloro-3-(trifluoromethyl)phenyl]methyl-[2,2-

di(phenyl)ethyl]amino]propoxy]phenyl]acetic acid) is a potent, orally-active synthetic

LXR agonist with slightly higher affinity for LXRβ (EC50: 30 nM) than LXRα (EC50: 190

58

nM) (Collins et al., 2002). Unlike T0901317, GW3965 has higher selectivity for LXR over

other nuclear receptors (Collins et al., 2002). The natural LXR antagonists include

geranylgeranylpyrophosphate (Forman et al., 1997), polyunsaturated fatty acids

(arachidonic acid, eicosapentaenoic acid, docosahexaenoic acid and linoleic acid) (Ou

et al., 2001; Yoshikawa et al., 2002) and 5α,6α-epoxycholesterol-3-sulfonate (Song et

al., 2001).

1.6.4.3 LXR: Molecular Mechanism of Action

LXRs regulate gene expression by forming obligate heterodimers with another nuclear

receptor, the retinoid X receptor (RXR). The LXR-RXR complex can regulate gene

expression via two mechanisms: direct activation and transrepression (Gabbi et al.,

2014). In the absence of agonist, the transcriptional activity of LXR/RXR is suppressed

by corepressors [e.g., Nuclear Receptor Corepressor (NCoR), Silencing Mediator of

Retinoic Acid and Thyroid Hormone Receptor (SMRT)] (Chen and Evans, 1995; Horlein

et al., 1995). During direct activation, agonist binding induces conformational change

in the LXR causing dissociation of the corepressor from the LXR/RXR dimer and results

in moderate level of transcription. Full transcriptional activity is gained by recruitment of

coactivators [e.g., Peroxisome Proliferator Activated Receptor-γ (PPARγ) coativator-1a

(PGC-1a), Steroid Receptor Coactivator-1 (SRC-1), Activating Signal Cointegrator-2

(ASC-2)] (Oberkofler et al., 2003; Huuskonen et al., 2004; Lee et al., 2008). The

activated LXR/RXR binds to the LXR-response element (LXRE) that consists of a direct

repeat of the core sequence 50-AGGTCA-30 separated by 4 nucleotides (DR4) of the

target gene (Wiebel and Gustafsson, 1997). LXRs, particularly LXRβ inhibit transcription

59

of NF-κB-regulated proinflammatory genes that lack LXRE by an alternate mechanism

called transrepression. In this case, upon agonist binding LXRβ undergoes a specific

SUMOylation by SUMO-2/3 that promotes interaction with the corepressor, NCoR and

its dissociation from the LXR, which in turn blocks transcription of proinflammatory

genes (Ghisletti et al., 2007; Venteclef et al., 2010).

1.6.4.4 LXR: Regulation of Cholesterol Metabolism

Much of the current understanding of the physiological role of LXRs comes from the

studies of their actions in peripheral tissues. It is now well established that LXRs act as

the “master regulators” of cholesterol metabolism. LXRs act as cholesterol sensors and

regulate intracellular cholesterol levels via three processes: 1) by promoting the

transport of excessive cholesterol in the peripheral tissues to the liver for excretion via

process called reverse cholesterol transport (RCT), 2) by inhibiting intestinal cholesterol

absorption and, to a lesser extent, 3) by inhibiting cellular cholesterol synthesis

(Wojcicka et al., 2007). LXRs regulate the RCT by controlling the expression of several

target genes. Excessive intracellular cholesterol leads to accumulation of oxysterols,

which activate LXRs that in turn increase expression of their principle target proteins,

ABCA1 and ABCG1, thus leading to enhanced intracellular cholesterol efflux (Figure

1.7). This pathway has been demonstrated in a wide variety of cell types (Costet et al.,

2000; Venkateswaran et al., 2000; Kennedy et al., 2001). Both ABCA1 and ABCG1 are

highly expressed in macrophages and prevent excessive accumulation of intracellular

cholesterol and formation of foam cells (Oram and Lawn, 2001). The cholesterol

pumped out by ABC transporters is accepted by apoA-I, the principle apolipoprotein in

60

peripheral HDL, which transports cholesterol to liver for excretion. In addition to ABC

transporters, LXRs also increase the expression of Niemann-Pick C1 (NPC1) and

Niemann-Pick C (NPC2), which are intracellular cholesterol carriers that redistribute

cholesterol from the endosomal compartment to the plasma membrane (Rigamonti et

al., 2005). LXRs also increase expression of ABCG5 and ABCG8 in hepatocytes that

transport cholesterol to bile canaliculi for its ultimate secretion into bile (Yu et al., 2002;

Yu et al., 2003). Finally, in rats and mice synthetic LXR agonists upregulate cholesterol

7α-hydroxylase (CYP7a), a rate limiting enzyme that converts cholesterol to bile acids

(Lehmann et al., 1997; Menke et al., 2002). This mechanism is unique in rats and mice

as the human CYP7a gene lacks an LXRE and therefore cannot be upregulated by LXR

agonists (Chiang et al., 2001; Goodwin et al., 2003) and may partly explain the

susceptibility of humans to dietary-induced hypercholesterolemia (Menke et al., 2002).

In addition to induction of RCT, LXR agonists have also shown to reduce intestinal

cholesterol absorption by upregulating ABCG5 and ABCG8 expression in the

enterocytes that increase cholesterol recycling into the intestinal lumen (Repa et al.,

2002; Yu et al., 2003).

61

Figure 1.7: Regulation of ApoE Lipidation by LXR. (1) In the absence of agonist LXR/RXR are bound to corepressors (CoR) which inhibit

their transcription activity. (2) Agonist (A) binding induces dissociation of corepressors

and association with coactivators (CoA) turning on transcriptional activity of LXR/RXR.

(3) Activated LXR activate the membrane-bound cholesterol transporter, ABCA1 that

pumps free cholesterol and phospholipids out of the cell. (4) Unlipidated ApoE act as

cholesterol acceptor and mature to lipidated apoE.

1.6.4.5 ABCA1: The Principal LXR Target

The principal target gene controlled by LXRs is ABCA1, a 2261 amino acid integral

membrane protein. ABCA1 is almost ubiquitously expressed with high levels of

expression in the macrophages, liver, intestine, heart, adrenal gland, endothelial cells,

testes, and in placental trophoblasts (Langmann et al., 1999; Lawn et al., 2001; Liao et

62

al., 2002; Wellington et al., 2002). High ABCA1 expression levels have been also

reported in mouse (Wellington et al., 2002) and rat (Fukumoto et al., 2002) brains with

highest levels in neuronal layer of the cerebellum. ABCA1 belongs to subfamily A of

large superfamily of ABC transporters. Although exact domain structure and topology of

ABCA1 are not fully elucidated, several models have been predicted based on

hydropathy analysis (analysis of hydrophobic and hydrophilic amino acids) that predict

ABCA1 structure (Schmitz and Langmann, 2001; Cavelier et al., 2006). According to

these models, ABCA1 is a full ABC transporter and consists of two transmembrane

domains, each consisting of six membrane-spanning helices followed by two

cytoplasmic nucleotide (ATP) binding domains. In addition ABCA1 also has a regulatory

domain between the two halves, a feature unique to ABCA1 among other ABC

transporters. The ATP-binding domains contain two conserved peptide motifs, Walker A

and Walker B and a family signature region. In addition, ABCA1 has two large

extracellular domains linked with two disulfide bonds and are essential for apoA-I

binding and HDL formation (Tanaka et al., 2001; Nagao et al., 2012).

Cellular expression of ABCA1 is tightly regulated at both the transcriptional and post-

transcriptional levels and its regulation is integrated with mechanisms that sense

intracellular cholesterol content (Tall, 2008). ABCA1 transcription is controlled

primarily by members of the nuclear receptor family of ligand-activated transcriptional

factors, specifically LXR-RXRs and PPARs (Schmitz and Langmann, 2005). ABCA1

expression is increased by natural oxysterols along with 9-cis retinoic acid, an RXR

agonist, as shown in various cell lines including macrophages, liver, intestine, and

63

neuronal cells in vitro (Costet et al., 2000; Schwartz et al., 2000; Kennedy et al., 2001;

Murthy et al., 2002; Koldamova et al., 2003). In addition, PPARγ agonists also induce

ABCA1 and associated apolipoprotein-mediated lipid efflux, although this induction is

likely secondary to activation of LXRs by activated PPARγ (Chawla et al., 2001; Ogata

et al., 2009). The in vivo efficacy and relevance of PPARγ-mediated activation of

ABCA1 via LXRs remains to be determined.

Cellular levels of ABCA1 are also controlled by a fine balance of post-transcriptional

modifications that either stabilize or degrade the ABCA1 protein, which has a half-life

of less than 1h (Oram et al., 2000; Wang and Tall, 2003). For example, interaction of

apoA-I with ABCA1 stabilizes ABCA1 protein (Wang et al., 2000; Wang et al., 2003).

ABCA1 contains a PEST sequence – proline (P), glutamate (E), serine (S) and

threonine (T) - in the cytoplasmic loop of ABCA1, the protein kinase A-dependent

phosphorylation of which induces degradation of ABCA1 by calpain (Martinez et al.,

2003; Wang et al., 2003). Interaction of ABCA1 with apoA-I prevents the

phosphorylation of this PEST sequence, thus stabilizing ABCA1. ABCA1 also

undergoes ubiquitin-proteosome-mediated proteolytic degradation (Feng and Tabas,

2002). Recently, a small non-coding microRNA, miR-33, has been shown to play an

important role in ABCA1 expression (Moore et al., 2010). MiR-33 inhibits ABCA1

expression in mouse hepatocytes and macrophages, blocking cholesterol efflux and

reducing plasma HDL levels. Conversely, inhibiting miR-33 by anti-sense

oligonucleotides increases ABCA1 expression and promotes lipid efflux both in vitro in

murine macrophages as well as in human THP-1, IMR-90 and HepG2 cell lines and in

64

vivo (Moore et al., 2010; Rayner et al., 2010). Whether CNS ABCA1 levels can also be

regulated by microRNAs remains an important question to address.

ABCA1 plays the pivotal role in controlling apoA-I and apoE metabolism. The critical

role of ABCA1 in HDL metabolism was first highlighted by loss-of-function mutations of

chromosome 9q31 leading to defective ABCA1 that cause Tangier disease, which is

characterized by extremely low plasma HDL levels (Hobbs and Rader, 1999; Rader and

deGoma, 2012). ABCA1 mediates transport of free cholesterol and phoshpolipids to

lipid-free apoA-I generating pre-β HDL, a rate-limiting step in HDL biogenesis (Oram

and Vaughan, 2000). Although the role of ABCA1 in HDL biogenesis is well established,

the exact mechanism of lipid transport to apoA-I by ABCA1 is still unclear and as such

various models have been proposed. Vedhachalam et al. (2007) proposed a three step

model of ABCA1-mediated lipidation of apoA-I in murine macrophages: in the first step,

apoA-I directly binds to ABCA1 causing ABCA1 upregulation that results in enhanced

translocation of phospholipids to the outer leaflet of the cell membrane. The net

increase and reduction in phospholipid content leads to later compression of

phospholipids in the outer half and expansion of phospholipids in the cytoplasmic leaflet

of the lipid bilayer, respectively. This asymmetric packing in the phospholipids causes

strain in the local cell membrane region. In the second step, the cell membrane relieves

the localized strain by bulging into the extracellular space forming an exovesiculated

domain to which apoA-I binds with high affinity. In the final step, binding of apoA-I

initiates spontaneous microsolubilization of the membrane phospholipids and

65

cholesterol and creation of discoidal HDL particles containing two, three or four apoA-I

molecules/particle.

A second model proposes that apoA-I binds to ABCA1 at the plasma membrane

followed by internalization of apoA-I-ABCA1 complex and subsequent delivery to the

late endosomes where apoA-I acquires lipids (Takahashi and Smith, 1999; Chen et al.,

2001). The lipidated apoA-I is then resecreted from the cell by exocytosis.

Using state of the art single-molecule fluorescence tracking and diffusion analysis

technique in HeLa cell line, Nagata et al recently proposed a novel mechanism of apoA-

I lipidation by ABCA1 (Nagata et al., 2013). According to this model, ABCA1 monomers

freely diffuse and translocate lipids on the plasma membrane, even in the absence of

apoA-I in an ATP-dependent manner. Upon lipid acquisition, presumably in its

extracellular domain or near its vicinity in the plasma membrane, ABCA1 monomer

undergoes conformational changes and forms dimers. The dimers are stabilized to the

plasma membrane by their interaction with the membrane-skeleton actin filaments.

Lipid-free apoA-I directly binds to the extracellular domains of ABCA1 dimers but not the

monomers and accepts lipids reserved by ABCA1. Upon lipid transfer to apoA-I, the

ABCA1 dimers dissociate into monomers, which are released from actin filament and

resumes its function to translocate lipids in an ATP-dependent manner.

Although the above models of apoA-I lipidation have been proposed based on the

studies in macrophages, it is proposed that ABCA1 may mediate apoE lipidation in the

CNS through similar mechanisms (Koldamova et al., 2010).

66

1.6.4.6 The Role of LXRs in the CNS

Although LXR function in the CNS is poorly understood, it is conceivable that like in the

periphery, LXRs regulate cholesterol metabolism in the CNS. The net movement of

cholesterol out of the brain is largely controlled by its conversion to the principal

metabolite, 24(S)-hydroxycholesterol (Dietschy and Turley, 2004). The second route of

intracellular cholesterol removal is via ABCA1 and apoE-mediated transport. Both

24(S)-hydroxycholesterol and synthetic LXR agonists increase ABCA1 expression in

neurons and astrocytes in vitro as well as in vivo along with increased cholesterol efflux

(Whitney et al., 2002; Liang et al., 2004; Abildayeva et al., 2006). Similarly, LXRs also

increase apoE expression in astrocytes in vivo (Liang et al., 2004; Abildayeva et al.,

2006). Cholesterol carried on apoE in the brain is thought to play minor role in the net

cholesterol removal from the CNS but may be rather important in redistribution of

cholesterol required for synaptogenesis, especially during development and following

recovery from brain injury.

As discussed further apoE plays crucial role in Aβ clearance and thus in turn in AD

pathophysiology. It is thus imperative to determine whether enhancing apoE function

through LXR activation may be beneficial in AD. Indeed, the in vivo activation of LXRs

has shown to decrease brain Aβ levels as well as improve cognition in mouse models of

AD. Koldamova et al reported decreased soluble Aβ40 and Aβ42 levels following a six-

day treatment with T0901317 in four-month old APP-23 mice (Koldamova et al., 2005b).

These findings were replicated by Riddell et al in another AD mouse model, Tg2527 in

67

which a six-day treatment with T0901317 in five-month old mice selectively lowered

Aβ42 but not Aβ40 in the hippocampus and reversed contextual memory deficits

(Riddell et al., 2007). It should be noted that in the above two studies the LXR agonist

response was assessed in young mice (4-5 months) and not in old mice (> 10 months)

that that develop extensive Aβ deposits. Jiang et al treated one-year old Tg2527 mice

with another LXR agonist, GW3965 for four months and observed ~50% reduction in

Aβ40 and Aβ42 levels with corresponding ~67% reduction in the plaque load (Jiang et

al., 2008a). The authors also reported significant decrease in the number of plaque-

associated activated microglia as well as IL-6 mRNA levels in GW3965-treated mice,

supporting anti-inflammatory effects of LXRs in the brain (Zelcer et al., 2007). Recently

Fitz et al treated 9 months old APP-23 mice with T0901317 for four months and

reported a significant decrease in the amyloid plaque load, soluble and insoluble Aβ40

and Aβ42 as well as A11-positive Aβ oligomers in cortex and hippocampus along with

reversal of Morris water memory deficits (Fitz et al., 2010). In all of the studies

mentioned above, treatment with LXR agonists did not affect APP processing, indicating

that LXR activation may affect Aβ metabolism but not its synthesis.

The beneficial effects of LXRs in Aβ metabolism are dependent on ABCA1 as shown by

several studies. We and other have shown that deletion of ABCA1 in mice results in

poorly-lipidated apoE and reduced apoE levels (Hirsch-Reinshagen et al., 2004; Wahrle

et al., 2004), which is associated with increased amyloid burden in several mouse

models of AD (Hirsch-Reinshagen et al., 2005; Koldamova et al., 2005a; Wahrle et al.,

2005). Moreover, in a recent study, lack of even one copy of ABCA1 resulted increased

68

amyloid deposition and exacerbated memory deficits in AD mice expressing human

apoE4 but not in mice expressing human apoE3 (Fitz et al., 2012). Conversely,

selective overexpression of ABCA1 in PDAPP mice has been shown to increase apoE

lipidation and prevent amyloid deposition (Wahrle et al., 2008). We also showed that

ABCA1 is required for the beneficial effects of GW3965 in AD mice in increasing CSF

apoE levels, reducing amyloid burden, and restoring novel object recognition (NOR)

memory, suggesting that lipidated apoE particles may underlie the neuroprotective

effects of LXR agonists (Donkin et al., 2010).

While the in vivo studies indicate beneficial effects of LXR activation in AD mice, the

results from in vitro studies are conflicting. Fukumoto et al showed increased secreted

levels of both Aβ40 and Aβ42 from neuroblastoma cell lines treated with T0901317, an

effect antagonized by blocking ABCA1 expression by RNAi (Fukumoto et al., 2002). On

the other hand, in two other studies treatment with the same LXR agonist resulted in

decreased Aβ secretion from neuroblastoma (Sun et al., 2003) and neuroglioma

(Koldamova et al., 2005b) cell lines. The reason for this discrepancy is not clear.

The role of LXRs in the CNS is further highlighted by gross anatomical abnormalities as

well as excessive lipid deposits in the brain and spinal cord seen in LXR knock-out

mice. Specifically, in LXRα/β double knockout mice, Wang et al reported closed lateral

ventricles that were lined with lipid-laden cells, enlargement of blood vessels particularly

in the substantia nigra pars reticulata and globus pallidus, excessive lipid deposits in the

brain parenchyma, neuronal loss, astrocyte proliferation and myelin sheath defects

69

(Wang et al., 2002). The astrocytes surrounding blood vessels also accumulated

excessive lipids in aged LXRα/β null mice (Wang et al., 2002). Recently, Anderson et al

reported that deletion of LXRβ in mice resulted in spinal-cord specific deficits, including

motor neuron loss, axonal atrophy, astrogliosis and lipid accumulation, and impaired

motor coordination; some of these features akin to that of amyotrophic lateral sclerosis

(Andersson et al., 2005).

1.6.4.7 ApoE Structure and Polymorphisms and their Significance in AD

Pathology

ApoE is a 299 amino acid glycoprotein with the molecular mass of 34 kDa and is

composed of N- and C- terminal domains that are connected by a flexible hinge domain

such that each domain can fold independently (Figure 1.8 A) (Aggerbeck et al., 1988;

Wetterau et al., 1988). The N-terminal domain is an elongated four-helix bundle and

contains receptor-binding site (22 kDa, amino acid residues 136-150) located in helix 4

(Wilson et al., 1991). The C-terminal domain contains an amphipathic α-helical lipid-

binding site (10kDa, amino acid residues 244-272) that can switch between lipid-bound

and lipid-free states (Hatters et al., 2006; Zhong and Weisgraber, 2009). The receptor-

binding site is rich in lysine and arginine residues that are thought to interact with the

ligand-binding domain of the apoE receptors, i.e. members of the LDL receptor family

(Hatters et al., 2006). Lipid-free apoE has poor affinity for binding to LDL receptor

(LDLR), while lipid-bound apoE binds with LDLR with high affinity.

70

Apolipoprotein E is encoded by the APOE gene located on chromosome 19, which is

highly polymorphic. In humans, APOE gene has three alleles, ε2, ε3, and ε4 that

encode three apoE isoforms, apoE2, apoE3 and apoE4, respectively (Weisgraber,

1994; Mahley, 1998). This apoE polymorphism leads to expression of six phenotypes:

E2/E2, E3/E3, E4/E4, E2/E3, E2/E4 and E3/E4. The three apoE isoforms differ by

single amino acid changes at residues 112 and 158: apoE 2 (Cys112, Cys158), apoE3

(Cys112, Arg158) and apoE4 (Arg112, Arg158) (Figure 1.8 B).

The single amino acid change has profound influence on the physical and physiological

properties as well as stability of the three alleles (Table 1.7), which may help to

understand their role in pathological conditions (Weisgraber, 1994). In apoE3 and

apoE4, the Arg at position 158 forms a salt bridge with Asp154. In apoE2 the Cys at 158

cannot form such bridge with Asp154. Instead, Asp154 forms salt bridge with Arg150

which lies in the receptor binding domain of apoE2, causing conformational changes

and interfering with receptor binding. As a result, apoE2 poorly binds with LDLR, while

apoE3 and apoE4 have ~50-fold higher affinity for LDLR (Weisgraber et al., 1982).

Two physical properties of apoE4, viz. domain interaction and poor stability have

been thought to alter its function and contribute to increased AD risk. X-ray

crystallography, site-directed mutagenesis, and fluorescent resonance energy transfer

studies have revealed that the amino acid at position 112 is the key to domain

interaction. Thus, Arg112 in apoE4 induces conformational changes in Arg61 causing

its side chain away from the four-helix bundle promoting interaction with Glu255 in the

71

C-terminal domain (Figure 1.8 C) (Dong et al., 1994; Dong and Weisgraber, 1996;

Hatters et al., 2005). On the other hand Cys at position 112 in apoE2 and apoE3 tucks

Arg61 side chain into the two-helix bundles preventing its interaction with Glu255

(Figure 1.8 C). All non-human species have threonine instead of arginine at the position

61, which does not engage in domain interaction. The mouse apoE, for example is

structurally similar to human apoE4 (Arg112, Arg158) (Figure 1.8 B) but functionally

behaves like apoE3 due to the absence of domain interaction (Figure 1.8 C).

Interestingly, replacing Thr61 with Arg in mouse apoE gene introduces domain

interaction causing the mutated apoE behave like human apoE4 resulting in lower apoE

levels (Raffai et al., 2001) along with cognitive deficits (Adeosun et al., 2014).

It is hypothesized that domain interaction alters apoE4 properties in two ways that may

contribute to the negative effects of apoE4 in AD. First, domain interaction is thought to

make the C-terminus of apoE4 more susceptible to cleavage by an unidentified

neuronal protease yielding C-terminal truncated apoE fractions that are thought to be

neurotoxic (Zhong and Weisgraber, 2009). This is supported by the observation that

neuron-specific expression of apoE4 in mice leads to intraneuronal accumulation of C-

terminal truncated apoE fragments, neuronal loss, increased tau phosphorylation, and

cognitive deficits (Harris et al., 2003; Brecht et al., 2004). Moreover, truncated apoE4

fragments have been reported in AD brains (Huang et al., 2001). Second, domain

interaction is also thought to influence the binding preference of apoE to lipid particles;

apoE2 and apoE3 preferentially bind to HDL while apoE4 is principally associated with

VLDL and LDL (Weisgraber, 1990; Dong and Weisgraber, 1996). The preference of

72

apoE4 to triglyceride-rich LDL and VLDL is thought to make it inefficient in lipid transport

required for synaptogenesis and neuronal repair (Zhong and Weisgraber, 2009).

The other physical property of apoE that is postulated to contribute to its influence in AD

pathology is the differences in the stability of the three isoforms. The N-terminal

domain of apoE2 is the most stable, that of apoE4 the least stable while apoE3 stability

is in between the two (Mahley and Huang, 2006; Zhong and Weisgraber, 2009). As a

result, apoE4 acquires a unique molten globule state in which the four-helix bundle is

partially unfolded exposing the proteolytic cleavage site (Morrow et al., 2002). A likely

effect of molten globule formation is apoE4-mediated potentiation of Aβ42-induced

lysosomal leakage and apoptosis, which has been shown in Neuro-2a cells transfected

with apoE4 (Ji et al., 2002). It is proposed that the acidic pH inside lysosomes favors

molten globule state of apoE4 that has increased affinity to phospholipids and

membranes, which in turn can result in destabilization of the lysosomal membrane and

ultimately lysosomal leakage (Mahley and Huang, 2006).

73

Figure 1.8: ApoE structure, polymorphism and domain interaction. (A) Schematic of the basic structure of apoE depicting the receptor and ligand binding

regions. (B) Human apoE isoforms with amino acids at key positions. Mouse apoE is

shown at the bottom for comparison. (C) Domain interaction in apoE4. In human apoE2

and 3 (blue) Cys112 in the N terminal domain changes conformation of Arg61 such that

it cannot engage in domain interaction with Glu225 in the C terminus. Mouse apoE

(yellow) although has arginine at position 112, the Thr61 cannot engage in domain

interaction with Glu225. In human apoE4, Arg112 changes conformation of Arg61 such

that it engages in domain interaction with Glu225.

74

Parameter ApoE2 ApoE3 ApoE4 General Allelic Frequency (%) 8.4 77.9 13.7

Frequency in AD Population (%)

3.9 59.4 36.7

Amino Acid at Position 112 Cys Cys Arg

Amino Acid at Position 158 Cys Arg Arg

Stability Highest Intermediate Least Folding Intermediates No Yes Yes

Domain Interaction No No Yes

Resistance to Molten Globule Formation

Highest Intermediate Least

LDLR Binding Affinity

Poor (associated with type III

hyperlipoproteinemia)

Bind with ~ 50-fold greater affinity than apoE2

Lipoprotein Binding Preference

Preferential binding to HDL Preferential

binding to LDL and VLDL

Table 1.7: Summary of physical and physiological properties of apoE alleles.

1.6.4.8 ApoE and Amyloid Beta (Aβ) Metabolism

The genetic association of apoE with AD pathologies stems from its well-documented

role in Aβ metabolism including clearance and BBB integrity. Data from these studies

indicate that the allele, lipidation status, and levels of apoE influence three aspects of

Aβ and amyloid metabolism: interaction with Aβ, Aβ degradation and Aβ clearance via

cerebrovasculature. While Aβ is continually generated at a rate of 7.6%/h, it is also

efficiently cleared from the CNS at an equivalent rate of 8.3%/h (Bateman et al., 2006).

75

Thus, the net steady state level of Aβ represents a delicate balance between its

formation and degradation. Aβ is cleared through three major pathways: 1) through

intracellular and extracellular enzymatic degradation including lysosomal degradation,

2) along interstitial fluid drainage pathways, and 3) through direct transport across

cerebrovasculature.

1.6.4.8.1 ApoE and Aβ Interaction

The first clues on the association between apoE and Aβ came from apoE

immunoreactivity found to be co-localized with parenchymal and cerebrovascular

amyloid deposits in AD patients (Namba et al., 1991; Wisniewski and Frangione, 1992)

and in AD mouse models (Burns et al., 2003). The lipid binding site, especially residues

244-272 located in the C terminal domain of apoE interact with residues 12-28 of Aβ

(Strittmatter et al., 1993). A later epitope mapping study revealed that residues 144-148

located in the receptor binding site on the N-terminal domain of apoE can also interact

with residues 13-17 of Aβ (Winkler et al., 1999). Several in vitro studies have shown that

apoE can bind to Aβ in an isoform-specific manner with apoE3 forming a more stable

complex with Aβ than apoE4 (LaDu et al., 1994; Tokuda et al., 2000; Petrlova et al.,

2011; Tai et al., 2013). It should, however, be noted that the findings of these studies

are influenced not only by the apoE isoform but also by its lipidation status and

experimental conditions. For example, purified and lipid-free recombinant apoE4 was

found to bind with Aβ with faster kinetics than apoE3 in vitro (Strittmatter et al., 1993).

Contrary to this finding, Tokuda et al reported that delipidation of apoE caused a 5-10-

fold decrease in Aβ binding without any influence of isoform (Tokuda et al., 2000). On

76

the other hand studies using HEK23 cell-secreted or native plasma apoE showed that

apoE3 formed more abundant sodium dodecyl sulfate (SDS)-stable complexes with Aβ

than apoE4 (LaDu et al., 1994; LaDu et al., 1995; Tokuda et al., 2000). To address this

discrepancy, LaDu et al compared Aβ binding with purified, lipid-free apoE vs native,

lipidated apoE and reported that delipidation of apoE during purification process causes

conformational changes in apoE resulting in diminished interactions between apoE3

and Aβ (LaDu et al., 1995).

1.6.4.8.2 ApoE and Enzymatic Degradation of Aβ

Several metalloproteases can degrade monomeric Aβ, the most important of which are

neprilysin (NEP) and insulin degrading enzyme (IDE) (Saido and Leissring, 2012). NEP,

a type-II membrane-associated zinc metalloprotease is the most potent Aβ degrading

enzyme in the CNS (Shirotani et al., 2001) and can degrade both monomeric and

synthetic oligomeric Aβ (Kanemitsu et al., 2003). NEP is highly expressed in microglia

(Iwata et al., 2000). Partial loss or decreased activity of NEP induced by intracerebral

injection of a NEP inhibitor in rats (Iwata et al., 2000) or in NEP knock-out mice (Iwata et

al., 2001; Rogers et al., 2002) increases Aβ levels by approximately two-fold, especially

in the hippocampus, suggesting important role of NEP in Aβ catabolism. Cortical and

hippocampal expression of NEP and activity decreases with aging in both rodents (Apelt

et al., 2003) and humans with AD pathology (Yasojima et al., 2001b; Yasojima et al.,

2001a; Russo et al., 2005; Wang et al., 2005) as well as in the cerebrovasculature in

AD cases with CAA (Carpentier et al., 2002; Miners et al., 2006). Another important

enzyme involved in Aβ catabolism is IDE, which is both membrane-bound as well as

77

secreted by microglia and mediates extracellular degradation exclusively of monomeric

Aβ (Qiu et al., 1998; Vekrellis et al., 2000). The Aβ degrading activity (Perez et al.,

2000) as well as expression levels (Cook et al., 2003) of IDE are significantly decreased

in AD compared to control cases. Complete IDE knockdown causes > 50% reduction in

Aβ degradation in mice suggesting an important role of IDE in regulation of Aβ levels

(Farris et al., 2003; Miller et al., 2003). Conversely, overexpression of NEP or IDE in AD

mice significantly decreases Aβ levels and retards plaque pathology suggesting that

these two enzymes may be potential therapeutic targets (Leissring et al., 2003).

ApoE facilitates both intracellular and extracellular proteolytic degradation of Aβ in vitro

(Jiang et al., 2008a). Further, the ability of apoE to facilitate enzymatic degradation of

Aβ is dependent on the lipidation status of apoE; poorly-lipidated apoE is inefficient

while lipidated apoE generated through LXR activation is efficient in stimulating

proteolytic degradation (Jiang et al., 2008a). Moreover, loss of ABCA1 impairs while

enhancement of ABCA1 expression through LXR activation promotes the ability of

microglia to degrade Aβ. This study also showed that human apoE2 promotes maximal

Aβ degradation while apoE4 is the least efficient, consistent with the isoform-dependent

effect of apoE in human AD pathology.

1.6.4.8.3 Cellular Uptake and Degradation of Aβ

Another route of Aβ degradation is through lipoprotein receptor-mediated uptake and

subsequent lysosomal degradation by the brain parenchyma and cerebrovasculature

(Kanekiyo et al.; Bu, 2009). The major lipoprotein receptors involved in Aβ uptake are

78

low density lipoprotein (LDL) receptor and LDL receptor related protein 1 (LRP1), which

are expressed in microglia, astrocytes, neurons, and smooth muscle cells at the BBB

and cerebral arteries (Fan et al., 2001). Aβ binds indirectly to LDLR and LRP1 through a

chaperone, mainly apoE, which can affect Aβ clearance. Recent in vitro studies have

shown that Aβ can also directly bind to both LDLR (Basak et al., 2012) as well as LRP1

(Deane et al., 2004b) indicating an apoE-independent receptor-mediated clearance

pathway. Overexpression of LDLR enhances cellular uptake of Aβ in vitro (Basak et al.,

2012) and reduces aggregation as well as enhances clearance of Aβ in APP/PS1

transgenic mice (Kim et al., 2009), suggesting LDLR modulation as a therapeutic target

for AD. Following endocytosis, the majority of endocytosed Aβ is transferred through

early and late endosomes before degradation in the lysosomes (Hu et al., 2009; Li et

al., 2012), while a small fraction is recycled through Ras-related protein, Rab-11-

mediated mechanism (Li et al., 2012). These studies have also shown that increasing

intracellular Aβ concentration through inhibition of Aβ degradation by lysosome

inhibitors or by excessive transport of Aβ to lysosomes by Rab-5 and Rab-7 can be

detrimental due to enhanced Aβ aggregation (Hu et al., 2009; Li et al., 2012) indicating

importance of lysosomal degradation in intracellular Aβ degradation.

1.6.4.8.4 Aβ Clearance through the Cerebrovasculature

Aβ is also cleared by the cerebrovasculature through two important routes. The first

clearance route is through the perivascular drainage pathway in which brain solutes,

including Aβ and ISF diffuse along the basement membrane between smooth muscle

cells of cerebral arterioles and ultimately drain into the CSF and lymphatic system

79

(Weller et al., 2008). The movement of Aβ and ISF along the perivascular space is in

the opposite direction of the blood flow and is thought to be driven by the blood vessel

pulsations (Schley et al., 2006). Aging as well as pathologies such as CAA and

cardiovascular disease can stiffen the vessel walls, reducing vessel pulsation and

slowing Aβ and ISF movement along the perivascular space, ultimately promoting Aβ

deposition within the capillary walls (Weller et al., 2002; Schley et al., 2006). The

second vascular clearance route involves receptor-mediated trafficking of Aβ from the

brain parenchyma into the peripheral circulation across the BBB. The two most

important receptors of this system are LRP1 and receptor for advanced glycation end

products (RAGE), which are shown to have opposite actions on Aβ clearance across

BBB (Deane et al., 2004a). Thus, while LRP1 mediates efflux of soluble Aβ40 into the

peripheral circulation preventing Aβ accumulation in the brain, RAGE can mediate influx

of Aβ from plasma to brain ISF and promote Aβ accumulation (Deane et al., 2003;

Deane et al., 2004a; Zlokovic, 2004). Elegant studies by the Zolkovic group suggest that

LRP1-mediated Aβ clearance across BBB may be more rapid (~ 6 fold faster) than ISF

flow (Bell et al., 2007) and may account for about 10-15% of the total Aβ clearance

(Shibata et al., 2000).

While lipidated apoE enhances enzymatic degradation of Aβ, the data on its effect on

Aβ transport across the BBB is conflicting. ApoE was initially postulated to enhance Aβ

clearance across the BBB through its interaction with Aβ as well as its ability to be

exported to the peripheral circulation via LRP1 in vascular endothelial cells (Deane et

al., 2004b). However, using in situ, real-time microdialysis, Bell et al reported that

80

association of Aβ40 with lipid-poor human apoE3 decreased while lipidated apoE

virtually blocked Aβ transport across the BBB in mice (Bell et al., 2007). The findings of

Bell et al were replicated in a later study where lipidation of all three apoE isoforms

significantly reduced the transport of apoE as well as apoE:Aβ complexes across the

BBB (Deane et al., 2008). In the same study Deane et al also showed that apoE, in fact

disrupts the transport of Aβ from brain ISF in an isoform-specific manner (apoE4 >

apoE3 > apoE2). Thus, binding of apoE to Aβ redirected the rapid export of free Aβ40

and Aβ42 from LRP1 to the VLDL receptor, which in comparison to LRP1, has a much

slower interaction with apoE-Aβ complex, resulting in 15- and 9-fold retention of Aβ40

and Aβ42, respectively (Deane et al., 2008). Along similar lines, apoE deficiency leads

to reduction of fibrillar Aβ deposits in AD mice (Irizarry et al., 2000) which may be due in

part to by reduced Aβ retention as suggested by the above microdialysis studies.

Although current evidence indicates that apoE interacts with Aβ in an isoform-specific

manner, its influence on Aβ clearance has been challenged by a recent study by

Holtzman and coworkers. In this study, Verghese et al mixed astrocyte-derived, human

CSF-derived or reconstituted lipidated apoE with cell-derived Aβ using a physiologically-

relevant apoE:Aβ ratio of 50-150:1 and found negligible levels of apoE-Aβ complexes

(Verghese et al., 2013). Moreover, the authors reported that apoE and Aβ competed for

LRP1-dependent cellular uptake by astrocytes. These data indicate that there is yet

much to learn about the relationship between apoE and Aβ.

81

Besides its influence on Aβ, apoE is also shown to serve other important role in the

CNS including synaptic function and synaptogenesis and inflammation.

1.6.4.9 ApoE, Synaptic Function, and Synaptogenesis

Neuronal membranes are rich in lipids, which are crucial in neuronal homeostasis and

function as well as synaptogenesis. Since adult neurons are almost entirely dependent

on external source for their lipid demands it is conceivable that apoE plays a crucial role

in the normal neuronal function including synaptic activity and synaptogenesis. In

support of this hypothesis, apoE modulates axonal growth in vitro. Incubation of cultured

fetal rabbit dorsal root ganglion cells with β-VLDL, which is rich in apoE promoted

neurite outgrowth and branching via LDLR-mediated mechanism (Handelmann et al.,

1992). Intriguingly, incubation with purified, lipid free apoE along with β-VLDL or

cholesterol resulted in increased neurite outgrowth but decreased branching

(Handelmann et al., 1992). Later studies showed that the neurotrophic effects of apoE

were isoform dependent; apoE3 promoted whereas apoE4 reduced neurite extension

and branching through an LDLR-mediated mechanism (Nathan et al., 1994; Holtzman

et al., 1995; Fagan et al., 1996; Nathan et al., 2002). The isoform-specific effect of apoE

was further demonstrated in murine neuroblastoma cells stably transfected with human

apoE3 and apoE4 wherein apoE3-secreting cells promoted neurite extension and

branching while apoE4 suppressed this effect (Bellosta et al., 1995). The inhibitory

effect of apoE4 on the neurite outgrowth was shown to be associated with apoE4-

mediated depolymerization of microtubules (Nathan et al., 1995).

82

In addition to synaptic remodeling, apoE is also thought to (in)directly affect learning

and memory by modulating synaptic plasticity. For example, apoE-deficient mice exhibit

spatial memory deficits as assessed by poor Morris water maze performance (Gordon

et al., 1995) and display impaired long-term potentiation (LTP) in hippocampal CA1

neurons (Valastro et al., 2001). Moreover, recent studies indicate that apoE can alter

synaptic plasticity in isoform-dependent manner. For example, LTP in the dentate gyrus

neurons is significantly impaired in an apoE isoform-dependent manner with apoE4 >

apoE2 and apoE-/- > apoE3 and wild-type mice (Trommer et al., 2004). Intriguingly,

opposite results were reported when LTP was assessed in the Schaffer collateral

synapses in the CA1 area with greatest LTP enhancement in apoE4 mice and least

amount of LTP induction in apoE2 mice (Korwek et al., 2009). ApoE isoform effects on

changes in LTP seems to be age-dependent with LTP enhancement seen in younger

apoE4 mice that disappears in old apoE4 mice, while LTP enhancement is unaltered in

the mice expressing apoE3 (Kitamura et al., 2004). The exact molecular mechanisms of

isoform-dependent effects of apoE on LTP are not fully understood. Reelin, an

endogenous ligand that regulates neuronal migration during embryonic development, as

well as neurotransmission in the adult brain enhances LTP through clustering of the two

apoE receptors, apoE receptor-2 (ApoER2) and VLDL receptors activating Src family

non-receptor tyrosine kinases, which in turn phosphorylate NMDA receptors, the key

receptors mediating LTP (Herz and Chen, 2006). Recently, apoE4 was shown to

selectively impair Reelin-mediated LTP by reducing surface expression of ApoER2 (by

locking ApoER2 in the intracellular compartment), thus interfering with Reelin-

dependent phosphorylation and activation of NMDA receptors (Chen et al., 2010). The

83

isoform-dependent effects on LTP are also reflected in hippocampal-dependent spatial

memory wherein female apoE4 mice show worse memory deficits compared to apoE3

mice (Grootendorst et al., 2005).

1.6.4.10 Role of ApoE in Inflammation

Studies in apoE-deficient mice suggest that apoE may have immunomodulatory roles

both in the periphery and in the CNS. ApoE-deficient mice show impaired immunity after

bacterial challenge with Listeria monocytogenes (Roselaar and Daugherty, 1998),

increased susceptibility to endotoxemia after intravenous lipopolysaccharide (LPS)

administration and inoculation with Klebsiella pneumonia (de Bont et al., 1999) and

increased systemic inflammatory response and higher mortality following LPS injection

(Van Oosten et al., 2001). Further, apoE modulates inflammation in an isoform-

dependent manner. Macrophages expressing apoE3 show lowest LPS-induced

expression of TNF-α and IL-6 compared to apoE4 and apoE2 (Tsoi et al., 2007). Similar

results were obtained in cultured primary microglia wherein apoE stimulated secretion of

proinflammatory cytokines such as TNF-α, IL-1β as well as prostaglandin E2 in an

isoform-specific manner with apoE4 > apoE3 > apoE2 (Chen et al., 2005; Maezawa et

al., 2006). The isoform-specific modulation of inflammation has been also studied in

apoE target replacement mice wherein apoE4 mice had significantly elevated TNF-α

and IL6 levels in brain as well as in the periphery compared to apoE3 mice (Lynch et al.,

2003). Moreover, brain autopsies revealed increased numbers of scattered microglia

and microglial activation in AD patients carrying the apoE4 allele (Egensperger et al.,

1998).

84

1.6.5 ApoE and TBI

Given the role of apoE in the lipid metabolism, synaptic function, inflammation, Aβ

metabolism, and possible in tau metabolism as well as the pathophysiological

similarities between TBI and AD, it is conceivable that apoE may modulate quite a few

aspects of TBI pathophysiology. While the role of apoE in AD pathology is well

established, its role in TBI is yet elusive. Two studies in apoE-deficient mice subjected

to closed head injury (CHI) give us a glimpse of the possible contribution of apoE in TBI

pathophysiology. In these studies, apoE-deficient mice showed greater cerebral edema

along with increased TNF-α mRNA levels (Lynch et al., 2002) as well as had greater

motor deficits, impaired spatial memory and overt neuronal death compared to wild-type

mice after CHI (Chen et al., 1997).

CNS injury leads to the release of large amounts of lipids from the degenerating axonal

membranes and myelin. Since CNS lipid metabolism is almost entirely based on apoE,

it is possible that apoE dynamics may change in response to TBI. In humans for

example, TBI results in a striking 14-fold selective reduction in CSF apoE-containing

lipoproteins from 2 to 5 days following injury and a 2- and 4-fold increase in total and

free cholesterol (Kay et al., 2003), reflecting a pronounced uptake of apoE particles by

the injured brain. Although data on the dynamics of apoE synthesis and utilization are

not yet available from human subjects, studies in animal models support a role for

increased synthesis of apoE in the CNS in the hours to weeks after many types of CNS

injury including TBI and spinal cord injury (SCI) (Poirier et al., 1991; Poirier et al.,

85

1993b; White et al., 2001a; Seitz et al., 2003; Iwata et al., 2005). For example, lesions

to the entorhinal cortex have been shown to induce coordinated increases in apoE

mRNA and its receptor, LDLR in parallel to clearance of lipid and membrane debris

supporting the idea that apoE may accept cholesterol and lipids released from the

damaged neuronal membranes and transport them back to the neurons undergoing

reinnervation and taken up through LDLR (Poirier et al., 1991; Poirier et al., 1993a;

White et al., 2001a). Iwata et al reported a biphasic upregulation of apoE and another

CNS lipoprotein, apoJ following fluid percussion injury in rats (Iwata et al., 2005). In this

model the initial peak appears at 2 days followed by return of apoE to baseline. ApoE

levels then again gradually increased over 6 months post-injury. Intriguingly, the authors

reported an uncoupling between the apoE protein and mRNA levels with mRNA levels

showing a delayed and sustained increase following an increase in brain apoE protein

levels, suggesting that apoE may enter into the injured brain by transfer through the

BBB or CSF (Iwata et al., 2005). ApoE isoform may also influence synaptic plasticity

after brain injury as apoE4 mice exhibit impaired synaptic plasticity compared to apoE3

mice following entorhinal cortex lesions (White et al., 2001b).

Epidemiological studies indicate that apoE may influence TBI outcomes in an isoform-

dependent manner although this relation is not as clear as in AD. Many studies have

reported poor post-TBI outcomes in apoE4 carriers compared to non-carriers (Sorbi et

al., 1995; Teasdale et al., 1997; Lichtman et al., 2000; Fleminger and Ponsford, 2005;

Smith et al., 2006) while others have found no such association (Jiang et al., 2008b;

Hiekkanen et al., 2009; Pruthi et al., 2010). These conflicting data may have resulted

86

from various confounding factors such as short follow-up times after TBI and a large

diversity of the subject population with respect to age, gender, and injury severity.

Nonetheless, a recent meta-analysis of 14 studies with a total of 2527 subjects

concluded that apoE4 does not affect initial injury severity but rather compromises

recovery at 6 months post-injury (Zhou et al., 2008). Moreover, accumulating evidence

suggests that age at injury may alter the association of apoE4 with TBI outcome. For

example, in a study of mTBI in pediatric subjects it was observed that the apoE4-

carriers were significantly more likely to have a worse GCS score compared to non-

carriers, yet performed equivalently, if not better, than non-apoE4 carriers on

neuropsychological tests over 12 months of follow-up (Moran et al., 2009). Another

recent study of CTE cases, which are also highly representative of brain injuries

experienced in late adolescence and early adulthood, also reported that the association

of apoE4 with worse outcomes is weaker than originally anticipated (Katsnelson, 2011).

1.7 TBI: An Overview of Animal Models

Since no single animal model can replicate the entire spectrum of human TBI

pathophysiology, several large and small animal models have been developed to mimic

particular pathophysiological aspects. Large animal models (e.g., non-human primates,

pigs, dogs, sheep, cats) are useful for investigating questions related to the response of

the gyrencephalic brain to injury, and may be particularly applicable for advanced

preclinical evaluation of therapeutic agents before introduction into humans (Kwon et

al., 2010). However, the high cost of large animal models and considerable ethical

concerns associated with their use limits their wide adoption into preclinical studies. In

87

contrast, rodent (e.g., rat and mouse) models have emerged as the most commonly

used animal models in TBI research, as they are widely available, cost-effective,

amenable to many behavioral, physiological and drug discovery evaluations, and,

particularly for mice, offer a wide variety of genetically modified strains.

Animal TBI models can be broadly classified into open head and closed head injury

models (Figure 1.9). The following section briefly summarizes important animal TBI

models along with their key advantages and disadvantages.

88

Figure 1.9: Animal models of TBI.

1.7.1 Open Head Injury Models

In open head injury (OHI) models, the mechanical force is applied directly to the dura,

which is exposed by a craniotomy and as such results in almost no head movement.

Animal Models of TBI

Open Head Injury

Fluid Percussion

Controlled-Cortical Impact

Penetrating Ballistic-Like Brain Injury

Needle Stab

Closed Head Injury

Impact-Acceleration (e.g., weight-drop)

Inertial Acceleration

Blast

89

The two most popular OHI models are fluid percussion (FP) and controlled cortical

impact (CCI).

1.7.1.1 Fluid Percussion (FP)

FP models generate brain injury by rapidly injecting fluid onto the intact dura through a

craniotomy, typically made laterally over the parietal cortex (lateral fluid percussion) or

medially over the sagittal suture midway between the bregma and lambda (midline fluid

percussion) (Figure 1.10) (Gurdjian et al., 1966; Thompson et al., 2005; Alder et al.,

2011). Fluid propulsion is driven by dropping a pendulum onto a fluid reservoir, and

pulse pressure, load duration and pulse velocity are reported. The pressure pulse of the

fluid can be varied to produce more or less severe injury, thus enabling FP models to

mimic a variety of human TBI injuries. FP induces mixed injury including petechial and

SAH, vascular damage at the gray/white matter interface, DAI, focal necrosis, and cell

loss (Povlishock and Kontos, 1985; McIntosh et al., 1987; Dixon et al., 1988; McIntosh

et al., 1989; Wang et al., 1997; Thompson et al., 2005). The biomechanical studies of

FP models have mainly concentrated on characterizing the input parameters such

volume loading, pressure peaks, and rate of fluid flow (Stalhammar et al., 1987). High-

speed X-ray imaging of the fluid pulse in rats (Dixon et al., 1988) and ferrets (Lighthall et

al., 1989) shows that the complex movement of fluid in the epidural space induces

gross movement of brain tissue. Data on the biomechanical response of the brain tissue

to the fluid movement, however is limited to a physical model (Thibault et al., 1992).

Although very widely used, particularly for rats, FP exhibits a high mortality rate due to

apnea and considerable variability of outcomes among different laboratories (Cernak,

90

2005; Marklund and Hillered, 2011). This variability is hypothesized to be due in part to

the surgical precision required to generate reproducible craniotomy position and angle

as well as minimal control over injury parameter (i.e., the height of the pendulum). As a

result, use of FP models often requires extensive operator training.

Figure 1.10: Fluid percussion injury model.

The injury is caused by a rapid pulse of saline caused by a pendulum striking the saline

reservoir. The inset shows movement of fluid in the epidural space. Reproduced with

permission from Macmillan Publishers Ltd: [Nature Reviews Neuroscience] (Xiong et al.,

2013).

1.7.1.2 Controlled Cortical Impact (CCI)

CCI models use a weight drop device (Feeney et al., 1981), a pneumatic (Dixon et al.,

1991; Smith et al., 1995), an electromechanic (Onyszchuk et al., 2007) or, most

recently, an electromagnetic (Brody et al., 2007) piston to deliver precisely controllable

tissue deformation to an exposed dura (Figure 1.11). This method induces primarily

focal damage with extensive cortical loss, hippocampal and thalamic damage,

intracranial hemorrhage, edema, and increased ICP (Saatman et al., 2006; Marklund

91

and Hillered, 2011), and is therefore used primarily to mimic severe TBI with frank

tissue destruction. The injury severity and pathology are affected by several variables

including craniotomy placement (midline vs lateral vs frontal), the impactor tip

configuration (size and shape), impact velocity and depth, and whether or not the

craniotomy is resealed after impact, which can be adjusted to produce various degrees

of injury with excellent precision and reproducibility. CCI models do not suffer from

“rebound injury” typically seen in weight-drop models (see below). On the other hand,

CCI models tend to induce severe injury destroying large cortical areas, which is

observed only in severe human TBI (Marklund and Hillered, 2011). Moreover, the

cortical damage during impact is difficult to visualize if the craniotomy is barely larger

than the impactor size, which may cause variation in the injury severity.

Figure 1.11: Controlled cortical impact model. The injury is caused by a piston compressing the exposed dura. The inset shows

direction of compression of the brain tissue. Reproduced with permission from

Macmillan Publishers Ltd: [Nature Reviews Neuroscience] (Xiong et al., 2013).

1.7.2 Closed Head Injury Models

92

In closed head injury (CHI) models, the injury is induced through the intact skull by

impact (e.g., dropping a weight on the intact skull or striking the intact skull with a

piston), non-impact (e.g., blast) or inertial loading (by rapid rotation of head in sagittal,

coronal or oblique planes). CHI models are characterized by varying degrees of head

acceleration.

1.7.2.1 CHI by Impact

In impact CHI, the mechanical force is delivered to the intact skull through an impact

that is delivered by various methods. The most common and simplest method is by

dropping a weight of known mass from a pre-determined height that falls under gravity

over the animal’s head (Marmarou et al., 1994; Chen et al., 1996; Tang et al., 1997;

Beni-Adani et al., 2001; DeFord et al., 2002; Pan et al., 2003; Zohar et al., 2003; Ucar et

al., 2006; Flierl et al., 2009; Zohar et al., 2011). Injury severity can be adjusted by

varying the mass of the drop weight or drop height. Impact is also delivered using a

pneumatically or electromagnetically-driven piston. Injury severity is usually controlled

by adjusting piston speed. In another method, Kilbourne et al used a steel ball that was

rolled down on a metal ramp to deliver impact (Kilbourne et al., 2009). The impact force

can be directed to the vertex (Marmarou et al., 1994; Tang et al., 1997; DeFord et al.,

2002; Zohar et al., 2003; Creeley et al., 2004; Wang et al., 2006), the dorso-lateral

surface (Laurer et al., 2001; Shitaka et al., 2011), lateral side (Viano et al., 2009, 2012;

Ren et al., 2013) or to the frontal side (Kilbourne et al., 2009) of the skull. The animal’s

skull is usually surgically exposed (but without any craniotomy) to direct impact on a

particular location on the skull, although in recent models (including the one developed

93

by us and described in the Chapter 3 of this thesis) can deliver impact to the head

without any surgery (see below).

Various support systems have been used to stabilize the animal’s head during impact

that also affect the degree of skull movement after impact and injury characteristics. For

example, the rodent’s head is supported on a thick block of foam or gel that allows

partial head acceleration causing moderate to severe brain injury with DAI as revealed

by the APP histochemistry (Foda and Marmarou, 1994; Marmarou et al., 1994; Viano et

al., 2012). An important variable is the spring constant of the foam used to support the

head that can influence the head movement after impact. Unfortunately, physical

characteristics of the foam, including the spring constant are rarely reported,

challenging the replication of the model across laboratories. In another configuration,

the animal head is held in a stereotaxic frame and undergoes the least amount of

movement and show mild APP staining (Laurer et al., 2001; Shitaka et al., 2011).

Recently, Whalen and colleagues described a new murine impact CHI model

characterized by a completely unrestrained head movement (Khuman et al., 2011;

Meehan et al., 2012). In this model, isoflurane-anesthetized mice were supported on a

piece of Kimwipes®, manually grasped by the tail and placed under a vertical guide

tube. The impact was delivered without surgically exposing the animal skull, between

coronal and lambdoid sutures with a 53 g metal bolt. Upon impact the mouse head tore

the Kimwipes support with free rotation in the anterior-posterior plane. This method

however, resulted in a moderately high post-CHI mortality rate (~ 20%). In a recent

variant of the ‘Kimwipe model’ the animal’s body was supported on a piece of aluminum

94

foil suspended over an empty case with a thick foam pad at the base of the case (Kane

et al., 2012). Upon impact, the aluminum foil completely tears off allowing unrestricted

head and body movement of the animal as it falls into the cage. These authors reported

repeatable, mild, concussive injury and suggested that the technique could be used for

repeated impacts. Ren and colleagues recently developed a ‘hit and run’ TBI model in

which the anesthetized mouse was hung head up vertically from its incisor teeth by a

mounted metal ring and impacted by a pneumatically-driven piston over the lateral side

of the head (Ren et al., 2013). Following impact the mouse fell onto a soft pad. The

authors reported that driving piston at 4.8 m/s resulted in mild injury characterized by

diffuse cortical astrogliosis, while impacting head at 5.2 m/s resulted in moderate injury

characterized by frank tissue disruption or cavitation.

Advantages of CHI models include the concordance of the mechanism with the vast

majority of human TBI that occur without skull fracture and, relative to FP and CCI,

employ less-invasive and simpler surgical methods that allow rapid behavioral

assessment of injury severity. Weight drop models, however, pose a number of very

significant limitations. For example, high-speed videography analysis shows that many

weight drop models deliver a primary and a rebound impact to the head. As well, weight

drop models are limited in that velocity of head impact and head displacement cannot

be varied independently as they can be when using an actuator. In addition, appropriate

release of head constraint to allow human-like head kinematics is challenging with

weight drop models and actuators that move downwards to create impact. A principal

caveat of most current CHI animal models is that both the input parameters (e.g.,

95

mechanical loading, method of mechanical input, response of animal’s head to

mechanical loading) as well as the cognitive, histological, and biochemical endpoints

used can vary considerably among different laboratories, which has challenged the

reproducibility of results and thus, the translational potential of these models. Much

more needs to be learned about how these CHI models alter animal head kinematics

(e.g., calculation of linear vs. angular acceleration, quantification of head acceleration),

how head kinematics relate to injury severity, how brain movement relates to head

movement in the species used, and to functional, molecular, and histological outcomes.

1.7.2.2 Non-Impact CHI by Blast Waves

Improvised explosive devices are a major source of TBI in military personnel. Blasts can

induce TBI through several mechanisms including the blast wave itself,

acceleration/deceleration forces, and particles that impact the head during the blast. To

simulate non-impact, blast TBI conditions in animals, the animal is secured at one end

of a long shock tube and is subjected to blast waves generated by compressed air or

gas or explosives (Figure 1.12) (Rubovitch et al., 2011; Goldstein et al., 2012; Rafaels

et al., 2012). The animal body is usually protected while the head is exposed to the

blast waves. These injuries typically present with diffuse brain edema accompanied by

hyperemia and delayed vasospasm. Again, the degree of head motion after blast

appears to have a significant effect on behavioral and neuropathological outcomes as

stabilizing the rodent head during blast was recently reported to be neuroprotective

(Goldstein et al., 2012).

96

Figure 1.12: Blast Wave TBI Model.

An animal is secured at one end of the shock tube. Blast waves are generated by

compressed air or explosives at the other end of the tube which travel toward the animal

causing head trauma. Reproduced with permission from Macmillan Publishers Ltd:

[Nature Reviews Neuroscience] (Xiong et al., 2013).

1.7.2.3 Non-Impact CHI by Inertial Loading

Many laboratories have developed inertial loading CHI models in which rotational force

causes rapid acceleration of the animal head followed by a longer deceleration phase.

These models have been developed for larger animals, including non-human primates

(Gennarelli et al., 1981; Gennarelli et al., 1982), pigs (Smith et al., 1997; Smith et al.,

2000; Browne et al., 2011), and rabbits (Gutierrez et al., 2001; Runnerstam et al., 2001;

Hamberger et al., 2003; Krave et al., 2005; Krave et al., 2011), as well as for small

animals such as rats (Davidsson and Risling, 2011). In these models, the animal head

is secured to a mechanical system that consists of a snout clamp or a skull fixation plate.

Linear motion induced by a piston is converted to rotational motion of the device, which

in turn results in a rotational acceleration of the head, along either the coronal

(Gennarelli et al., 1982; Smith et al., 1997; Smith et al., 2000), sagittal (Gennarelli et al.,

1982; Gutierrez et al., 2001; Runnerstam et al., 2001; Davidsson and Risling, 2011;

97

Krave et al., 2011), axial (Smith et al., 2000; Browne et al., 2011), lateral or oblique

(head turned 30° to right or left side and moving anterior or posterior with the center of

rotation in the lower cervical spine) planes (Gennarelli et al., 1982). The severity of

trauma can be adjusted by varying the angle of rotation or pulse duration. Studies using

these models have shown that post-traumatic neurological status, such as coma, is

related to the energy and form of rotation induced (Gennarelli et al., 1981; Gennarelli et

al., 1982; Browne et al., 2011; Krave et al., 2011). These models have also shown that

rotational forces can induce pathologies including SAH, increased ICP and elevated

serum levels of an astrocyte-specific protein, S100B (Runnerstam et al., 2001;

Davidsson and Risling, 2011). A distinctive neuropathological feature of these models is

the DAI, as evident by the formation of axonal bulbs and accumulation of APP (Smith et

al., 2000; Hamberger et al., 2003; Davidsson and Risling, 2011). Other microscopic

pathologies include gliosis (Smith et al., 1997; Gutierrez et al., 2001), neuronal death

(Smith et al., 1997; Runnerstam et al., 2001), and accumulation of phosphorylated

heavy neurofilament subunits at neuronal cell bodies (Smith et al., 1997; Smith et al.,

2000; Hamberger et al., 2003).

1.7.3 Challenges Associated With Current TBI Models

The choice of the most appropriate preclinical TBI model to use depends on what

factors of human injury need to be modeled. Compared to CHI models the input

parameters of FP and CCI models are well-controlled and thus induce highly

reproducible injuries but have some disadvantages. First, these models represent only a

small minority (0.8-3%) of human injuries that involve penetration of dura (Masson et al.,

98

2001; Myburgh et al., 2008; Wu et al., 2008). Second, these models, especially CCI,

tend to induce severe injury through destruction of large cortical areas, which is typically

observed only in severe human TBI (Marklund and Hillered, 2011). Third, these

methods often require a precisely positioned craniotomy that can reduce secondary

brain swelling by mimicking decompressive craniotomy often used to lower ICP

(Zweckberger et al., 2006; De Bonis et al., 2010). Fourth, these methods of injury cause

almost no head movement, which is rarely observed in human TBI.

CHI models, that apply impulsive (impact) loads to the head, by contrast, are

increasingly attractive for modeling the majority (>95%) of human TBIs that occur

without skull fracture or penetration of brain tissue. However, most CHI model systems

have significant caveats with respect to their high incidence of skull fracture

(necessitating immediate euthanasia of the animal) and highly variable outcomes

between laboratories. We can gain further insights into the relative strengths and

weaknesses of the various rodent injury models if we interpret these models in the

context of what is known about the biomechanics of human, large animal, and rodent

TBI. The vast majority of rodent CHI studies however, lack systematic biomechanical

rigor, particularly for weight drop models where only gross mechanical input parameters

such as the mass of the weight and drop height are typically reported. The

reproducibility of these mechanical inputs and measures to decrease variability in

mechanical input are rarely stated, and the mechanical response of animal’s head to the

forces applied is even more rarely recorded. As a result, the outcomes of studies even

99

with similar reported mechanical inputs can vary from no detectable injury to death

(Figure 1.13).

Figure 1.13: Variability in input parameters and outcomes in weight-drop TBI studies reported in the literature.

The figure depicts variation in two most commonly reported input parameters, drop

height and drop mass (A) and associated theoretical values of input energy (B) used in

mouse weight drop impact TBI studies. Panel C shows variation in mortality rate

associated with input energy. Reproduced from Namjoshi et al. (2013a).

One potential source of variability especially relevant for CHI models is the relationship

of head kinematics (moment of the head) following impact to both behavioral and

pathological outcomes after injury. The head kinematics can vary depending on how the

head is supported during impact. A few recent studies have begun to report details of

head kinematics following impact CHI in rats ((Viano et al., 2009; Li et al., 2011; Viano

100

et al., 2012). For example, Li et al. (2011) recently studied the biomechanical

parameters of the weight drop-based Marmarou impact acceleration model (Marmarou

et al., 1994). Head acceleration, change in head velocity, linear and angular head

acceleration were measured and compared to 1) kinematic responses predicted by a

finite element model and 2) axonal damage as assessed by APP histochemistry. The

authors reported that the severity of DAI was related to the linear and angular response

of the rat head but, against expectation, not with the drop height. Viano and colleagues

developed a rat model of CHI with unrestricted head movement to mimic concussions

experienced by professional football players (Viano et al., 2009, 2012). The ability to

study the detailed biomechanical responses of the rats and to allow unrestricted motion

in response to impact and the proposed methods for scaling these responses to human

concussions offer new insights into factors that may increase the translational potential

of future CHI studies in rodents.

In animal models, classification of injury severity as mild, moderate or severe is often

determined on functional outcomes such as cognitive and motor impairment that

develop over hours to weeks. Importantly, these outcomes can vary widely among

different research groups. For example, motor impairment may (Tang et al., 1997;

DeFord et al., 2002; Andoh and Kuraishi, 2003; Creeley et al., 2004; Rubovitch et al.,

2011) or may not (Tsenter et al., 2008) be observed with “mild” CHI. The severity,

temporal response, and extent of secondary injury response of brain can also vary

among rodent species and also within different strains of the same species. Because

the severity of the primary injury at the time of impact is often unknown and may be

101

below the threshold of detection even when using methods such as MRI, understanding

how biomechanical data affect the functional and histological outcomes of rodent CHI

may offer an alternative approach to improve the predictive and translational power of

preclinical TBI research in rodents.

Economic, ethical and scientific drivers have shifted the focus of preclinical TBI studies

from large animals to mainly rodents, even though large animals are better mimics of

the size and anatomy of the human brain (Denny-Brown and Russell, 1940; Gennarelli

et al., 1981; Gennarelli et al., 1982; Duhaime, 2006). One important consideration for

selecting an appropriate rodent TBI model is that that the rodent and human brain

differs considerably with respect to skull anatomy, brain mass and size, craniospinal

angle, gray-to-white matter ratio and cortical folding, all of which influence the

biomechanical parameters of brain injury. For example, rotational loading cannot be

linearly scaled between human and rodent TBI (Finnie, 2001). Furthermore, the long

axis of the rodent brain and spinal cord are nearly linear, as opposed to perpendicular

axes of the human brain and spinal cord, which causes less rotational shearing in the

rodent brain and renders rodents less vulnerable to diffuse axonal damage. Although

the anatomical differences between the rodent and human brain will always pose a

caveat for preclinical research, rodent-based experiments are expected to form the

majority of mechanism of action and proof-of-concept studies that can then be further

validated in large animal species. This progression will provide the greatest adherence

to ethical principles and scientific feasibility.

102

A common question regarding rodent models of TBI is whether the response of the

lissencephalic rodent brain to impact, acceleration or blast injury is representative of

that of the gyrencephalic human brain. For example, cases of CTE, which occurs from

repeated concussions, commonly show more extensive pathology in sulcal depths than

elsewhere along the cortex, suggesting that shear strain may be maximal at the sulcal

depths (Smith et al., 2013). Further work is warranted to investigate the role that these

anatomical differences may play in the context of mTBI research. Even if the

lissencephalic vs. gyrencephalic differences are found to result in an inability to

reproduce anatomical injury patterns observed in humans, the vast potential of rodent

models for studies investigating secondary injury pathways and for drug discovery

research is a major incentive to continued use of these species.

1.8 Pharmacological and Non-pharmacological Management of TBI: An Overview

of Randomized Clinical Trials

A recent review of 100 randomized clinical trials (RCT) examined efficacy of

interdisciplinary interventions of 55 acute phase and 45 post-acute phase TBI trials (Lu

et al., 2012). Acute phase trials were defined as interventions that occurred within 24h

of TBI. These studies largely focused on patient stabilization and minimizing secondary

injury, and employed standard outcomes comprised of GCS scores and mortality. By

comparison, post-acute trials can be very heterogeneous in design. The post-acute

trials reviewed were initiated from days-to-years post-TBI and used a wide variety of

outcomes including cognitive, neuropsychological, and quality of life measures (Lu et

al., 2012).

103

In the acute phase, only 15 of 55 trials showed a positive treatment effect and, of these,

11 evaluated non-pharmacological interventions. Furthermore, many interventions that

were tested across multiple trials led to mixed results. For example, decompressive

craniotomy, hyperosmotic therapy and hypothermia have shown inconsistent effects,

with studies reporting positive (Cruz et al., 2001, 2002; Zhi et al., 2003; Jiang et al.,

2005; Qiu et al., 2005; Jiang et al., 2006; Qiu et al., 2007), negative (Cooper et al.,

2011), or no effect (Smith et al., 1986; Marion et al., 1997; Shiozaki et al., 2001; Clifton

et al., 2002; Lu et al., 2003; Cooper et al., 2004; Harris et al., 2009; Clifton et al., 2011)

on post-TBI outcomes. A single trial of hyperventilation to reduce ICP also revealed an

adverse effect (Muizelaar et al., 1991). Other non-pharmacological interventions that

have shown benefit include early nutritional supplementation with zinc within 48 hours

or total parenteral nutrition within 72 hours post-TBI (Young et al., 1996). Lastly, a pre-

hospital rapid sequence intubation trial also yielded positive results (Bernard et al.,

2010).

The 32 acute phase pharmacological trials reviewed were grouped according to the

targeted treatment mechanism. Anti-excitotoxic agents, insulin, magnesium,

corticosteroids, and targeting lipid peroxidation or free radical damage either failed to

show a positive treatment effect or led to adverse effects. Importantly, only 4 of 32

pharmacological acute phase RCTs for TBI showed positive treatment effects, including

the psychostimulant, methylphenidate (Ritalin) (Moein et al., 2006), 2 RCTs for

104

progesterone (Wright et al., 2007; Xiao et al., 2008) and 1 RCT for the calcium channel

blocker, nimodipine (Harders et al., 1996).

Of the 45 post-acute phase RCTs reviewed, 24 evaluated cognitive rehabilitation

approaches and of these 22 showed positive treatment effects. The approaches used

included comprehensive interdisciplinary rehabilitation, cognitive/academic exercises

and communication skill training, compensatory techniques and computer assisted

training, educational intervention, psychotherapy, and behavior modification. Six of 8

trials using physical rehabilitation also showed a positive treatment effects, and nutrition

and acupuncture were found to be beneficial in the single trials conducted thus far.

Potential TBI pharmacotherapies were tested in 11 post-acute RCTs, with positive

treatment effects reported in 6 studies, including methylphenidate (Whyte et al., 2004;

Willmott and Ponsford, 2009), CDP-choline (Calatayud Maldonado et al., 1991), and a

vitamin B6 analog, pyritinol (Kitamura, 1981). On the other hand, a single trial of

phenytoin and carbamazepine was negative (Smith et al., 1994), while sertraline

(Novack et al., 2009; Banos et al., 2010), rivsatigmine (Tenovuo et al., 2009) and

modafinil (Jha et al., 2008) were found to have no significant treatment effects.

A number of RCTs in TBI have been completed in 2012-2013. Highlights include a

Phase III randomized trial of hypothermia in children, termed the “Cool Kids” study,

which recently reported that 48 h of hypothermia followed by rewarming does not

reduce mortality or positively affect recovery from pediatric brain injury (Adelson et al.,

2013). An intervention trial randomizing high-risk post-concussion syndrome patients to

105

an early visit to a physician similarly was negative, with similar outcomes to the

treatment as usual group (Matuseviciene et al., 2013). Chesnut et al recently reported

on their much anticipated trial of continuous invasive ICP monitoring following severe

TBI, showing no advantage over imaging and clinical examination alone (Chesnut et al.,

2012). Citocoline, which demonstrated promise in earlier trials as a neuroprotective

agent, failed in a subsequent Phase III trial (Zafonte et al., 2012). A trial of 30

hyperbaric oxygen therapy sessions, done in military personnel with mTBI, did not

demonstrate a significant effect (Wolf et al., 2012). Finally, in perhaps the most

influential trial of 2012, amantadine, given to TBI patients who were minimally conscious

or in a vegetative state 4 to 16 weeks following TBI, showed a highly significant effect

on accelerating the pace of functional recovery (Giacino et al., 2012).

The emerging consensus from these 100 RCTs conducted on TBI subjects is that early

intervention in the acute phase and comprehensive rehabilitation approaches in the

post-acute phase provide the most beneficial outcomes. What is far less clear is

whether a deeper understanding of the pathways involved in TBI pathogenesis may

eventually offer effective pharmacologic therapies. A critical requirement for any drug

discovery program is the availability of preclinical models that are as biofidelic to the

human condition as possible. With respect to TBI, however, a working knowledge about

neurophysiology and biomechanics is necessary to select the most appropriate model

system to address the experimental question being considered.

1.9 Study Rationale, Hypothesis, and Objectives

106

As discussed above, repair of damaged CNS critically depends on the lipid transport

mediated by brain lipoproteins. Lipid-laden brain lipoproteins aid in neuronal membrane

repair and synaptogenesis and stimulate the degradation of Aβ peptides that

accumulate after TBI. Genetic and pharmacological methods to increase the cholesterol

and phospholipid carried on brain lipoproteins improve cognitive function and reduce

neuropathology in AD mice. However, it is not known whether similar strategies may

also be beneficial for acute CNS injury.

We and others have previously demonstrated that synthetic LXR agonists effectively

induce ABCA1 expression, increase apoE lipidation, reduce Aβ levels, lower

inflammation and restore cognitive function in AD mouse models. Given that TBI

pathology shares common features with that of AD pathology including Aβ deposition

and NFT formation as well as neuroinflammation, it is of considerable interest to

determine whether LXR agonists may also have potential therapeutic utility for TBI. The

importance of addressing this question is two-fold. First, LXR treatment may minimize

neuronal damage and promote recovery in the acute phase after injury by reducing

inflammation and promoting apoE-mediated neuronal repair processes. Second, by

facilitating Aβ clearance in the first weeks after injury, LXR treatment may also lessen

the increased long-term risk of AD decades later.

Accordingly I hypothesized that the synthetic liver X receptor agonist, GW3965 will

promote neuronal recovery following closed-head mild, repetitive traumatic brain

injury in mice via apoE-mediated mechanisms.

107

Thus, the first objective of this thesis was to study therapeutic utility of the non-

steroidal LXR agonist, GW3965 in mild, repetitive closed head TBI in adult male mice

and whether the hypothesized beneficial effects of GW3965 were mediated through an

apoE-based mechanism. I tested this hypothesis using a widely-used weight drop CHI

model due to its several advantages over OHI models discussed previously. The

experimental details of the first objectives are presented in Chapter 2 of this thesis.

While weight drop TBI is one of the most popular rodent CHI model, it suffers from

several caveats like skull fractures and considerable experimental variation across

cognitive, histological, and biochemical outcome measures. This can be due to poor

control over the input parameters used to induce injury (e.g., mechanical loading,

method of mechanical input, response of the animal’s head to mechanical loading). In

addition, surgical manipulations required to localize the impact on the exposed skull

necessitate the use analgesic cocktails, sedatives and long-term exposure to

anesthetic. For example, in the weight drop CHI model used in the first objective, ethical

constraints necessitated us to use a non-steroidal anti-inflammatory drug (NSAID),

meloxicam for post-TBI pain suppression as well as ketamine-xylazine cocktail to keep

the animal sedated during the CHI procedure. These agents can prolong the post-

anesthetic recovery interfering with the immediate post-TBI assessment of neurological

outcomes. In addition, NSAIDs can suppress post-TBI inflammation, while, ketamine, a

NMDA receptor blocker, can suppress post-TBI glutamate-induced excitotoxicity,

interfering with the secondary injury pathways. Thus, it was essential to develop a

108

neurotrauma model that would mitigate these limitations and thus increase its

translational potential. While the second objective of this thesis was not driven by a

specific hypothesis, it was derived from the caveats encountered in the first study

objective.

Accordingly, the second objective of this thesis was to develop and a CHI model that:

1) offers tight control over injury parameters such as impact location and impact energy,

2) allows unrestricted head movement after impact mimicking most of the human TBI

conditions, 3) allows CHI without any surgical manipulation thus reducing anesthetic

analgesic and sedative requirement, 4) allows assessment of head kinematics during

and following impact, and 5) allows immediate post-TBI assessment of neurological

outcomes with minimal interference from anesthetics and sedatives. I further

characterized acute post-TBI outcomes in mice using this new CHI model.

The experimental details of the second objective are presented in Chapter 3 of this

thesis.

109

Chapter 2: The Liver X Receptor Agonist GW3965 Improves Recovery from Mild

Repetitive Traumatic Brain Injury in Mice Partly Through Apolipoprotein E

2.1 Summary

Traumatic brain injury (TBI) increases Alzheimer’s disease (AD) risk and leads to the

deposition of neurofibrillary tangles and amyloid deposits similar to those found in AD.

Agonists of Liver X receptors (LXRs), which regulate the expression of many genes

involved in lipid homeostasis and inflammation, improve cognition and reduce

neuropathology in AD mice. One pathway by which LXR agonists exert their beneficial

effects is through ABCA1-mediated lipid transport onto apolipoprotein E (apoE). To test

the therapeutic utility of this pathway for TBI, we subjected male wild-type (WT) and

apoE-/- mice to mild repetitive traumatic brain injury (mrTBI) followed by treatment with

vehicle or the LXR agonist GW3965 at 15 mg/kg/day. GW3965 treatment restored

impaired novel object recognition memory in WT but not apoE-/- mice. GW3965 did not

significantly enhance the spontaneous recovery of motor deficits observed in all groups.

Total soluble Aβ40 and Aβ42 levels were significantly elevated in WT and apoE-/- mice

after injury, a response that was suppressed by GW3965 in both genotypes. WT mice

showed mild but significant axonal damage at 2 d post-mrTBI, which was suppressed

by GW3965. In contrast, apoE-/- mice showed severe axonal damage from 2 to 14 d

after mrTBI that was unresponsive to GW3965. Because our mrTBI model does not

produce significant inflammation, the beneficial effects of GW3965 we observed are

unlikely to be related to reduced inflammation. Rather, our results suggest that both

110

apoE-dependent and apoE-independent pathways contribute to the ability of GW3965

to promote recovery from mrTBI.

2.2 Introduction

Several epidemiological studies suggest that TBI may increase the risk of dementia,

particularly AD, although this association is not always observed (May et al., 2011). TBI

has also been associated with an earlier onset of AD (May et al., 2011). Both

neurofibrillary tau tangles and amyloid plaques are found in post-mortem TBI brain

tissue. However, the widespread white matter involvement in TBI that is not present in

AD results in noteworthy differences in the pattern and distribution of their common

neuropathological features. For example, TBI neuropathology is largely that of tau

deposition, as only approximately 30% of TBI patients also contain Aβ deposits

(Roberts et al., 1994). Aβ plaques can appear within hours and remain detectable up to

47 years after a single moderate-severe TBI (Johnson et al., 2012). NFTs are also

detectable after a single TBI and are remarkably prominent in mild, repetitive,

concussive injuries (Johnson et al., 2012). Despite the low prevalence of amyloid

deposits in the post-mortem TBI brain, APP accumulation in damaged axons is a

striking histological hallmark of diffuse axonal injury (DAI) (Gentleman et al., 1993;

Sherriff et al., 1994b). Increased APP in the post-TBI brain is hypothesized to trigger a

burst of Aβ production that can deposit in amyloid plaques (Johnson et al., 2010).

Intriguingly, microdialysis experiments in brain-injured humans have demonstrated that

ISF Aβ levels correlate positively with the patient’s GCS score (Magnoni and Brody,

111

2010; Magnoni et al., 2012), suggesting that Aβ release is associated with recovery of

synaptic function, as has been demonstrated in animals (Schwetye et al., 2010).

Apolipoprotein E (apoE) is the major lipid carrier in the brain (LaDu et al., 2000a).

Brain injury increases apoE levels in order to scavenge lipids released by degenerating

neurons and myelin, which are later delivered to surviving neurons during reinnervation

and synaptogenesis (Poirier, 1994; Laskowitz et al., 1998). In humans, TBI results in a

14-fold selective reduction in CSF apoE levels by 2 to 5 days following injury and a 2- and

4-fold increase in total and free cholesterol, respectively (Kay et al., 2003), reflecting the

pronounced lipid mobilization, remodeling, and uptake of CSF apoE lipoproteins by the

injured brain. In animal models, apoE is induced within hours to weeks after TBI

followed by increased neuronal uptake of apoE-containing lipoproteins through the

LDLR (Poirier, 1994; Iwata et al., 2005). In mice, apoE deficiency compromises

recovery from acute neurological injuries, including TBI (Chen et al., 1997; Han and

Chung, 2000).

ApoE also regulates Aβ metabolism, a function that underlies its genetic association

with AD. The human apoE4 allele increases AD risk and hastens its onset (Corder et al.,

1993; Poirier et al., 1993b), whereas apoE2 delays onset and reduces AD risk (Corder et

al., 1994). Recent microdialysis studies demonstrated that apoE genotype modulates the

Aβ half-life in the brain ISF, with the rate of Aβ degradation accelerated by apoE2 (t1/2 =

0.56 h) and prolonged by apoE4 (t1/2 = 1.1 h) compared to apoE3 (t1/2 = 0.71 h)

(Castellano et al., 2011). Although AD risk and age of onset are indisputably modified

112

by apoE genotype, the relationship between apoE genotype and TBI outcome is

complex. Some studies report that apoE4 carriers have significantly poorer outcomes

compared to non-carriers (Teasdale et al., 1997; Lichtman et al., 2000; Smith et al.,

2006), while others find no association (Jiang et al., 2008b; Pruthi et al., 2010). Many

variables pose challenges to resolving these uncertainties, including relatively short

follow-up times after TBI and the diversity of subjects with respect to the age, gender,

and injury severity. Despite these challenges, a meta-analysis of 14 studies with 2,527

subjects concluded that apoE4 does not affect initial injury severity but rather may

compromise recovery at 6 months post-injury (Zhou et al., 2008).

Improving apoE function is thought to facilitate both neuronal repair and Aβ clearance

after TBI, thus potentially offering both acute and long-term benefits. One method to

enhance apoE function is to promote its lipidation by ABCA1, which is the rate-limiting

step in generating apoE-containing lipoprotein particles in the CNS. ABCA1 deficiency

leads to poorly-lipidated apoE in the CNS (Hirsch-Reinshagen et al., 2004; Wahrle et

al., 2004), and increases amyloid load in AD mice (Hirsch-Reinshagen et al., 2005;

Koldamova et al., 2005a; Wahrle et al., 2005). Conversely, overexpression of ABCA1 in

the CNS increases apoE lipidation and greatly decreases amyloid deposits in AD mice

(Wahrle et al., 2008). Transcription of ABCA1 and apoE is induced by agonists of LXRs,

which regulate many genes involved in lipid metabolism and inflammation (Jamroz-

Wisniewska et al., 2007; Wojcicka et al., 2007). Genetic deficiency of LXRs increases

amyloid burden in AD mice (Zelcer et al., 2007). Synthetic LXR agonists including

TO901317 and GW3965 cross the BBB, induce ABCA1 and apoE expression, improve

113

memory and reduce Aβ levels in AD mice (Jiang et al., 2008a; Donkin et al., 2010; Fitz

et al., 2010). Importantly, ABCA1 is required for several beneficial effects of GW3965 in

AD mice (Donkin et al., 2010), including increased CSF apoE levels, reduced amyloid

load, and improved memory. These observations provide a compelling rationale for

testing the therapeutic potential of LXR agonists for TBI. Indeed, TO901317 reduces Aβ

accumulation and promotes cognitive recovery in a CCI model of moderate-severe brain

injury in mice (Loane et al., 2011).

Approximately 80% of human TBI are mild injuries without skull fracture and loss of

consciousness (Decuypere and Klimo, 2012). It is increasingly appreciated that

repetitive, mild TBI, commonly experienced by athletes in high-contact sports may lead

to chronic traumatic encephalopathy (CTE), which is characterized by cognitive,

executive, and motor function disturbances, tau deposition and, in some cases, amyloid

deposits similar to those found in AD (Gavett et al., 2011).

To determine the therapeutic utility of LXR agonists for this type of brain injury, we

established a mouse model of mild repetitive TBI (mrTBI), wherein a gravity-driven

weight drop device was used to deliver two consecutive injuries 24h apart. Here we

report that GW3965 improves cognitive recovery and suppresses axonal damage in an

apoE-dependent manner. Surprisingly, apoE was not required for GW3965 to suppress

the transient increase in Aβ levels. Our results provide additional support for the

therapeutic potential of LXR agonists in TBI, and demonstrate that both apoE-

dependent and apoE-independent pathways contribute to their beneficial effects.

114

2.3 Materials and Methods

2.3.1 Animals

Male C57Bl/6 (WT) and apoE-/- (strain name: B6.129P2-Apoetm1Unc/J) mice were

obtained from Jackson Laboratories (Bar Harbor, ME). ApoE-/- mice are on C57Bl/6

background. Animals were housed with a reverse 12 h light - 12 h dark cycle for at least

10 days before TBI. The animals were fed with standard chow diet (PMI LabDiet® 5010,

containing 24% protein, 5.15 fat, and 0.03% cholesterol) and had free access to water.

All animal procedures were approved by the University of British Columbia Committee

on Animal Care (Protocol # A07-0706) and adhered to the Canadian Council for Animal

Care guidelines.

2.3.2 Mild Repetitive Traumatic Brain Injury (mrTBI)

WT (mean body weight: 29.6 ± 0.24 g) and apoE-/- (mean body weight: 30 ± 0.26 g)

mice at 4 months of age were subjected to two mild CHI spaced 24 h apart using a

gravity-driven weight-drop device obtained from Dr. Esther Shohami (The Hebrew

University of Jerusalem). The device consists of ~ 50 cm long graduated acrylic guide

tube held vertically over a metal plate (Figure 2.1A). A Teflon-tipped cone (tip diameter:

2 mm) is located at the base of the tube that can be freely raised inside the tube. A 95 g

brass weight is hung inside the tube with a thread. The height of weight is manually

adjusted by a thread holding lever. For inducing injury, the animal’s head is positioned

under the lower end of the guide tube and the Teflon-tipped cone is rested over the

exposed skull. In the original Shohami model, the animal’s head was supported by a

115

rigid Teflon disk. We modified this support by introducing ~ 3 mm thick computer mouse

pad between the animal’s head and the Teflon disk to reduce incidence of skull fracture

by partially allowing head movement following impact. This modification was made after

conducting a small pilot study with and without the mouse pad. The weight is raised to

the desired height and dropped over the Teflon-tipped cone, which directs the impact

force to a localized area of the animal skull.

All surgical procedures were conducted using aseptic technique. One day before the

TBI procedure, animals were weighed and the scalp was shaved under isoflurane

anesthesia (induction: 3-4%, maintenance: 1.5% - 2%). On the day of TBI surgery,

animals were anesthetized with isoflurane (induction: 3-4%, maintenance: 1.5% - 2%) in

oxygen (0.9 L/min). Lubricating eye ointment was applied to prevent corneal drying.

Animals were kept warm with an electric heating pad. The scalp was scrubbed three

times with 4% chlorhexidine solution and 70% ethanol. Ketamine (15 mg/kg) and

xylazine (1.75 mg/kg) mixture [subcutaneous (s.c.)], meloxicam (1 mg/kg, s.c.), and

bupivacaine (0.125 mg/kg, s.c., under the scalp) were administered for analgesia.

Sterile saline (1 ml/100 g body weight) was administered s.c. for hydration. The surgical

plane of anesthesia was confirmed with the absence of toe-pinch reflex and a midline

longitudinal incision was made in the scalp to expose the calvarium. After clearing the

parietal bone, a sterile piece of gauze (~ 3 mm x 5 mm), pre-moistened with

bupivacaine solution (2.5 mg/ml) was overlaid over the left parietal bone to minimize the

incidence of skull fracture (Figure 2.1 B). Anesthesia was temporarily discontinued (not

more than 30 s) and the animal was quickly moved under the weight-drop device. The

116

animal remained sedated with ketamine-xylazine during this temporary discontinuation

of anesthesia. The head was manually held in place such that the tip of the Teflon-

tipped cone rested over the piece of gauze in the left mid-parietal bone [approximately

between -1 and -3 mm from bregma and 1 mm left of the sagittal suture (Paxinos and

Franklin, 2001) ] (Figure 2.1 B and C). A 95 g brass weight was dropped from a pre-

determined height to result in injury across the left hemisphere. The precise drop height

(cm) of the weight was adjusted to be one-third the weight (g) of the animal (Mean drop

height: WT, 10 ± 0.08 cm; apoE-/-, 10.1± 0.09 cm). Mean body weight and drop height

did not defer significantly across the genotypes, time points, and treatment groups

(Figure 2.2). Anesthesia was reinstituted following injury. After removing the gauze and

confirming the absence of skull fracture, the incision was closed and the animals were

returned to their cages. Twenty-four hours after the first injury, the incision was

reopened under isoflurane anesthesia and a second identical TBI was induced as

described above. Mice suffering from skull fracture after injury were immediately

euthanized. Post-TBI recovery was monitored until animals moved freely and fed

normally. Sham controls received isoflurane anesthesia, pre-surgical analgesia, and

skin incision only. Mice within each group were randomly assigned to one of three post-

TBI assessment time points, viz., 2, 7, and 14 days after the second injury. The time of

the 2nd TBI was defined as time zero.

117

Figure 2.1: Weight drop CHI model and impact location. (A) Schematic of weight drop device and mouse head under the device. (B) Illustration

of surgical preparation for TBI showing an anesthetized mouse with the midline scalp

incision exposing the calvarium. A sterile piece of cotton gauze is overlaid across the

skull to help reduce incidence of skull fracture upon impact. The location of impact is

indicated by red dot. (C) Schematic diagram of murine skull showing approximate

location of the weight drop impact (red circle) on the left parietal bone. The dotted

rectangle represents cotton gauze.

118

Figure 2.2: Average animal body weight and drop height across all study groups.

To induce mrTBI, a 95 g brass weight was dropped on the animal’s skull using a height

(cm) adjusted to one third of the animal’s body weight (g) (see Methods and Materials).

The figure depicts body weight (A, B) and drop height (C, D) of WT and apoE-/- mice,

respectively, across post-mrTBI time points and treatment groups. Neither drop height

nor body weight differed significantly across genotypes, time points or treatment groups.

Data within each genotype were analyzed by two-way ANOVA followed by a Bonferroni

post-hoc test. Legend: V: untreated mice, open bars, G: GW3965-treated mice, black

bars.

2.3.3 GW3965 Treatment

Injured mice were randomized into treated (GW3965) and untreated (vehicle) groups

(WT mice: N=45/group, apoE-/- mice: N=30/group). Since, the animals do not become

fully ambulatory and start feeding for up to 2 h following TBI, the GW3965 treatment

was started with a single intraperitoneal (i.p.) bolus (20 mg/kg) 30 min after the 2nd

119

injury. GW3965 treatment was continued thereafter by feeding mice with standard

rodent chow (Research Diets Inc., New Brunswick, NJ) in which GW3965 was

compounded at 120 mg/kg until the end of the study, resulting in an average daily dose

of 15 mg/kg upon determination of average daily food intake. Because we had

previously demonstrated that GW3965 doses of 2.5 mg/kg/day and 33 mg/kg/day

provided cognitive and neuropathological benefits in APP/PS1 mice (Donkin et al.,

2010), we aimed to achieve a dose between these two extremes. A complete dose-

response study was beyond the scope of this thesis project. Because GW3965 has a

short plasma half-life of 2 h and high oral bioavailability (70%) (Collins et al., 2002)

delivering GW3965 in chow provides more constant plasma levels than by once-daily

oral gavage. Injured mice in the untreated group received a single i.p. bolus of

equivalent volume of the vehicle used for dissolving GW3965, dimethyl sulfoxide 30 min

after 2nd injury. These mice were thereafter fed standard rodent chow. Sham-operated

mice (WT mice: N=24; apoE-/- mice: N=18) received neither GW3965 nor vehicle and

were fed standard rodent chow. We have previously determined that GW3965 at a dose

of 33 mg/kg/day does not cause hepatotoxicity in mice (Donkin et al., 2010).

2.3.4 Cognitive Assessment by Novel Object Recognition

Medial temporal lobe function was assessed using the novel object recognition (NOR)

task performed at 2, 7, or 14 days following mrTBI (Ennaceur and Delacour, 1988). This

task was selected to facilitate comparisons to our previous NOR study of GW3965

efficacy in APP/PS1 mice (Donkin et al., 2010), and because fine motor impairments

observed after TBI may confound tasks such as Morris Water Maze. Each animal was

120

assessed with the NOR task at only one time point after sham or mrTBI surgery.

Twenty-four hours prior to the NOR task, mice were habituated to an empty open field

(14” x 24” x 14”) for 10 min in a brightly-lit room. The NOR task consisted of two phases:

training and testing. During training, mice were placed in the center of the open field

containing two identical objects spaced equidistant from the walls (Figure 2.3). The time

the animal explored each of the two objects during a 5 min epoch was quantified, with

exploring defined as sniffing, climbing on, or touching the object and at a body length or

less away from the object while oriented toward it. Trials were tracked using an

overhead digital camera and video tracking system (ANY-maze, Stoelting Co, Wood

Dale, IL). Between animals, the open field and objects were cleaned with 70% ethanol

to eliminate olfactory cues. Four hours after training, mice were tested by replacing one

of the identical objects with a novel object of similar size but different shape and color

(Figure 2.3). The side of the open field with the novel and familiar objects was

alternated for each mouse to avoid any side preference. Each animal was tracked for a

period of 5 min. Novel object recognition was quantified using the discrimination index

(DI) calculated as: DI = (TN – TF)/(TN + TF), where TN and TF represent exploration time

for the novel and familiar objects, respectively. DI values range from +1 to –1, where a

positive score indicates preference for the novel object, a negative score indicates

preference for the familiar object, and zero indicates no discrimination (Antunes and

Biala, 2012).

121

Figure 2.3: Novel object recognition phases.

During the training phase the mouse is placed in a rectangular box containing two

identical objects (indicated by yellow circles) and exploration time for each object is

recorded for 5 min. After a retention period of 4 h, the animals are re-introduced to the

same box with one familiar (yellow circle) and a novel object (red hexagon).

2.3.5 Motor Function Assessment by Accelerating Rotarod

Motor performance was evaluated using an accelerating rotarod (Model 7650, Ugo

Basile, Collegeville, PA). Mice were first trained for two days before the 1st TBI to walk

on accelerating rotarod (0 to 30 rpm in 210 s). The average of the last two trials of the

second training day was used to establish baseline failing latency. Unless the animal

was scheduled for sacrifice, rotarod latencies were repeatedly assessed at 1, 2, 7, and

14 d post-mrTBI. At each test day, every mouse underwent two accelerating speed

trials (0 to 30 rpm in 210 s) separated by an inter-trial interval of at least 15 min. The

animal was recorded as failing during the 210 s period if: (1) the animal fell completely

off the rotating rod before 210 s, or (2) the animal gripped the rod and made one

complete passive revolution without actively walking on the rod. The average failing

latency was calculated from the two trials for each animal.

2.3.6 Biochemical Analyses

122

2.3.6.1 Tissue Collection

Mice were anesthetized with an i.p. injection of 150 mg/kg ketamine (Bimeda-MTC) and

20 mg/kg xylazine (Bayer) and transcardially perfused with ice-cold phosphate-buffered

saline (PBS) containing heparin (0.5 units/mL) for 7 min. For biochemistry, brains were

longitudinally bisected and the ipsilateral half-brains were rapidly frozen over dry ice and

stored at -80°C until analysis (WT mice N: injured = 10/group/time point, sham = 6/time

point; apoE-/- mice N: injured = 5/group/time point, sham = 4/time point). For histology,

anesthetized mice were perfused with heparinized PBS and whole brains were

immersion-fixed in 10% neutral buffered formalin for 48 h followed by cryoprotection in

30% sucrose in PBS at 4°C for additional two days (N: injured = 5/treatment group/time

point/genotype, sham = 2/time point/genotype).

2.3.6.2 Protein Extraction

Frozen ipsilateral half-brains were thawed over ice and homogenized in 1.5 mL of ice-

cold RIPA lysis buffer (5mM EDTA, 50mM NaCl, 10mM sodium pyrophosphate, 50mM

NaF, and 1% NP40 alternative, pH: 7.4) containing complete protease inhibitor cocktail

(Roche Applied Science) with a Tissuemite homogenizer at full speed for 20 s. The

homogenate was sonicated at 20% output for 10 s followed by centrifugation at 4°C for

10 min at 9,000 rpm in a microcentrifuge (Eppendorf). The supernatant was separated

and used for analysis of ABCA1, apoE, APP, APP-CTF, LDLR, IL-6, TNF-α, MCP-1,

Aβ40, and Aβ42. Aliquots of all lysates were immediately frozen at –80°C until analysis.

Protein concentration was determined by DC Protein assay (Bio-Rad, Hercules, CA).

123

2.3.6.3 Western Blot Analyses

RIPA lysate (80 µg protein) was electrophoresed through 10% SDS-polyacrylamide

gels, transferred to PVDF membrane (Millipore), and immunodetected for the detection

of ABCA1, apoE, and LDLR. APP and APP-CTF were analyzed by resolving RIPA

lysate through 4-12% NuPAGE Bis-Tris gradient gels (catalogue # NP0335, Invitrogen).

The details of primary antibodies used for Western blotting are presented in the Table

2.1.

Antibody Clone Concentration Source

ABCA1 AC10 1:1000

Gift from Dr. M. R. Hayden, Centre for Molecular Medicine and Therapeutics, The University of British Columbia

ApoE M-20 1:1000 Santa Cruz Biotechnology, Santa Cruz, CA

APP Holoprotein 22C11 1:2000 Chemicon APP-CTF C1/6.1 1:1000 Covance

LDLR Polyclonal 1:1000 R&D Systems

Table 2.1: Details of primary antibodies used for Western blotting.

All membranes were probed with anti-GAPDH antibody (clone 6C5, 1:5000, Chemicon)

as a loading control. Blots were developed using enhanced chemiluminescence

(Amersham ECL, GE Healthcare) according to the manufacturer’s recommendations.

Films were scanned into TIFF format at 600 dpi resolution and pixel counts were

determined using ImageJ (version 1.46, NIH). Levels of ABCA1, apoE, APP, APP-CTF,

and LDLR were normalized to GAPDH and expressed as fold difference compared to

sham controls.

124

2.3.6.4 Quantitative Assessment of Endogenous Aβ by ELISA

Endogenous Aβ40 and Aβ42 levels in ipsilateral half-brain lysates were quantified by

ELISA (KMB3481 and KMB3431, Invitrogen) following the manufacturer’s instructions.

Levels of soluble Aβ40 and Aβ42 were normalized to total protein.

2.3.6.5 Quantitative Assessment of IL-6, TNFα, and MCP-1 by ELISA

IL-6, TNFα and monocyte chemotactic protein-1 (MCP-1) levels in ipsilateral half-brain

lysates were determined by ELISA. Briefly, 96-well plates were coated and incubated

overnight at 4°C with primary antibody against murine IL-6, TNFα, and MCP-1. Brain

lysates were incubated with primary antibody at room temperature for 1 h. Following

washing, samples were incubated with biotinylated secondary antibodies at room

temperature for 1h. All the primary and secondary antibodies were obtained from

eBioscience. The details of primary and secondary antibodies are presented in the

Table 2.2. Samples were developed with avidin-horseradish peroxidase (eBioscience)

and TMB substrate.

Cytokine Primary Antibody Secondary Antibody Clone Concentration Clone Concentration

IL-6 MP5-20F3 1:1000 MP5-32C11 1:1000 TNFα TN3-19.12 1:1000 Polyclonal 1:1000,

125

(Catalogue # 13-7341) MCP-1 4E2 1:1000 2H5 1:1000

Table 2.2: Details of primary and secondary antibodies used for cytokine ELISA.

2.3.7 Histology

Brains were cut into coronal sections (40 µm) on a cryotome and all sections starting

from the genu of corpus callosum (~1.1 mm anterior to bregma) to the posterior

hippocampus (~3.3 mm posterior to Bregma) were collected in 1X PBS with sodium

azide.

2.3.7.1 Assessment of Axonal Injury by Silver Staining

Three to four coronal brain sections, separated by 400 µm, from the bregma to the

posterior hippocampus, were stained using FD NeuroSilver Kit II (FD

NeuroTechnologies Inc, Columbia, MD) according to the manufacturer’s protocol, which

is based on the method originally described by Gallyas et al (1990). The manufacturer’s

protocol was further optimized with the following modifications: 1) Sections were stored

in 4% paraformaldehyde in saline-free phosphate buffer at 4°C for a minimum of 5 d

prior to staining. 2) Step 4 was decreased from 2x2 min incubations to 1x2 min

incubation to increase staining intensity. 3) Steps 3, 4, and 5 were performed with

vigorous shaking. From each section, 40X-magnified images of ipsilateral corpus

callosum, cingulum, external capsule, internal capsule, and optic tracts were captured

with a Zeiss Axio Imager A1 microscope attached to an Olympus DP72 camera and

digitized using cellSens® Standard digital imaging software (version 1.4.1, Build 8624,

OLYMPUS). A semiquantitative scale was developed to score staining intensity, ranging

126

from 0 (0 to <10% staining) to 3 (> 70% staining). The scale was based on the extent

and intensity of argyrophilic structures (cell bodies and axonal fibers) within an image

field. Every image was independently scored by two trained raters who were blinded to

the injury, genotype, and treatment status. Scores from the two raters were averaged. A

minimum of 3 sections per brain were scored and combined to define the average silver

stain score of each brain region. The silver staining was carried out on samples

containing 5 injured brains/genotype/time point/group and 2 sham brains/genotype/time

point.

2.3.7.2 Assessment of Microglial Activation by Iba1 Immunohistochemistry

Microglial activation induced by mrTBI was assessed with Iba1 immunohistochemistry.

Briefly, 3-4 coronal brain sections, separated by 400 µm, starting from the bregma to the

posterior hippocampus were washed in Tris-buffered saline (TBS), treated with 0.3%

hydrogen peroxide for 10 min and blocked with 3% normal goat serum (NGS) in TBS

containing 0.25% Triton-X (TBS-X) for 30 min. Sections were incubated overnight in

TBS-X containing anti-Iba1 polyclonal antibody (Catalogue # 019-19741, 1:1000, Wako

Chemicals, Richmond, VA) and 1% NGS. Following washing, sections were incubated

in biotinylated goat-anti-rabbit secondary antibody (1:1000 in TBS-X) for 1 h. Sections

were visualized with horseradish peroxidase (Vectastain Elite ABC Kit PK-6120, Vector

Laboratories (Canada) Inc, Burlington, ON) and DAB substrate. Sections were mounted

on Superfrost® Plus (Fisher Scientific) slides, dehydrated in ascending concentrations

(50, 70, 90, and 95%) of alcohol, cleared with xylene, and coverslipped in DPX

127

mounting medium (Electron Microscopy Sciences, Hatfield, PA). Images were captured

and digitized as above.

The mrTBI and GW3965 treatment protocol is summarized in Figure 2.4.

Figure 2.4: Summary of mrTBI, Gw3965 treatment and post-injury outcomes.

The post-TBI time points are relative to the second TBI. GW: GW3965, NOR: Novel

object recognition, V: Vehicle (DMSO) used for dissolving GW3965 for single i.p. bolus,

WT: Wild-type mice.

2.3.8 Statistical Analyses

All data are presented as mean ± SEM. For analysis of NOR data, because all mice

equally explored two similar objects during the training phase (DI ~ 0), DI scores during

training for animals of the same genotype were pooled. Similarly, DI scores during

testing were pooled for sham mice within each genotype. DI scores were compared

128

using two-way ANOVA followed by a Bonferroni post-hoc test. For rotarod analysis,

baseline latencies of mice from all time points within genotype and treatment groups

were pooled. Rotarod data were analyzed for time and treatment as well as time and

genotype effects using two-way repeated measures ANOVA followed by a Bonferroni

post-hoc test, as animals were tested repeatedly until sacrifice. For biochemical and

histological analyses, data from sham mice from all time points within each genotype

were pooled. Biochemistry and silver stain data were analyzed for time and treatment

as well as for time and genotype effects using two-way ANOVA followed by a Bonferroni

post-hoc test. A p value of < 0.05 was considered significant. All statistical analyses

were performed using GraphPad Prism (version 5.04, GraphPad Software Inc).

2.4 Results

2.4.1 ApoE is Required for GW3965 to Improve Cognitive Function After mrTBI

Non-spatial, working memory was assessed using the novel object recognition (NOR)

task at 2, 7, and 14 days following mrTBI. Discrimination between the novel and familiar

objects was determined using the discrimination index (DI) (Antunes and Biala, 2012).

During training, all animals displayed an equivalent time exploring each of the two

identical objects (DI ~ 0) (Figure 2.5) indicating that genotype, surgical procedures,

treatment, and time did not affect baseline exploratory behavior. Sham-operated WT

(Figure 2.5 A, gray bar, p < 0.001 compared to training, two-way ANOVA) and apoE-/-

(Figure 2.5 B, gray bar, p < 0.01 compared to training, two-way ANOVA) mice retained

preference to novelty as indicated by positive DI values. Untreated WT (Figure 2.5 A,

open bars) and apoE-/- (Figure 2.5 B, open bars) failed to discriminate between the

129

novel and familiar objects (p > 0.05 compared to training, two-way ANOVA) at all post-

mrTBI time points. Although GW3965-treated WT mice showed impaired memory at 2d

post-mrTBI (Figure 2.5 A, black bars, p > 0.05 compared to training, two-way ANOVA)

object recognition was restored by day 7 and 14 post-mrTBI (p < 0.001 compared to

training, two-way ANOVA). Finally, GW3965 treatment failed to improve NOR

performance in apoE-/- mice at any post-mrTBI time points (Figure 2.5 B, black bars, p

> 0.05 compared to training, two-way ANOVA). Gross motor impairment can affect NOR

performance. To rule out this possibility, we assessed the spontaneous activity of the

animal in the open field by comparing the total distance covered by the animal in the 5

min epochs of training and testing phases. We found that the total path length covered

by sham and injured WT as well as apoE-/- mice were not significantly different

indicating no gross motor impairment (Figure 2.6, p > 0.05, two-way ANOVA).

130

Figure 2.5: ApoE is required for GW3965 to improve NOR performance after mrTBI.

NOR memory was evaluated in untreated (V) and GW3965-treated (G) WT (A) and

apoE -/- (B) mice at 2, 7, and 14 days post-mrTBI. Bars represent discrimination index

(DI) scores. Cohorts were: sham (WT, n=23, pooled; apoE-/-, n=18, pooled) and 2 day

(WT: untreated, n=18, GW3965-treated, n= 15; apoE-/-: untreated, n=10, GW3965-

treated, n=10), 7 day (WT: untreated, n=13, GW3965-treated, n=10; apoE-/-: untreated,

n=11, GW3965-treated, n=11), and 14 day (WT: untreated, n=14, GW3965-treated,

131

n=14; apoE-/-: untreated, n=9, GW3965-treated, n=9) post-mrTBI. DI scores were

analyzed using two-way ANOVA and Bonferroni post hoc test. **: p<0.01, ***: p<0.001,

###: p<0.001.

Figure 2.6: NOR performance was not affected by motor impairment.

To assess whether NOR performance was affected by motor impairment; the total path

length (m) covered by WT and apoE-/- mice during testing and training was measured.

(A, B), total path length covered by WT mice during training and testing, respectively.

(C, D), total path length covered by apoE-/- mice during training and testing,

respectively. The path lengths covered by WT and apoE-/- mice were not significantly

different from sham animals during testing, indicating that NOR was not affected by

motor impairment. *: p<0.05 and **: p<0.01. Data were analyzed by two-way ANOVA

followed by a Bonferroni post hoc test. Legend: V- untreated mice, open bars, G-

GW3965-treated mice, black bars.

132

2.4.2 Spontaneous Recovery of Motor Dysfunction After mrTBI is Unaffected by

GW3965

We next assessed post-mrTBI fine motor functions with accelerating rotarod. mrTBI

significantly impaired the motor performance of WT (Figure 2.7 A, p < 0.05, two-way

repeated measures ANOVA) and apoE-/- (Figure 2.7 B, p < 0.05, two-way repeated

measures ANOVA) mice at 1, 2, and 7 d post-injury as indicated by reduced failing

latencies on an accelerating rotarod, compared to the baseline performance. By 14 d,

however, motor performance had fully recovered in WT and apoE-/- animals (Figure 2.7

A and B, p > 0.05, two-way repeated measures ANOVA). GW3965 treatment of WT or

apoE-/- mice did not significantly reduce the severity of motor impairment at any time

point, nor did it accelerate the rate of spontaneous recovery in either genotype (Figure

2.7 A and B, curly brackets). In contrast to GW3965, loss of apoE-/- led to significantly

more severe motor impairment compared to WT animals at 1 and 2 d post-mrTBI in

both untreated and treated groups (Figure 2.7 C and D, p < 0.05, two-way repeated

measures ANOVA).

133

Figure 2.7: mrTBI-induced motor impairment recovers spontaneously independent of GW3965 and apoE.

Motor performance of WT (n=38) and apoE-/- (n=30) mice was evaluated by the

accelerating rotarod task (0 to 30 rpm in 210 s). (A) Rotarod latencies of untreated (V,

open squares) and GW3965-treated (G, black filled squares) WT mice before and

following mrTBI. (B) Rotarod latencies of untreated (V, open circles) and GW3965-

treated (G, black filled circles) apoE-/- mice before and following mrTBI. Asterisks in A

and B denote significant differences between baseline and post-mrTBI latencies within

each group. For both genotypes, the treatment effect was not significant at any post-TBI

time point (curly brackets). (C) Rotarod latencies of untreated WT (open squares) and

apoE-/- (open circles) mice before and following mrTBI. (D) Rotarod latencies of

GW3965-treated WT (black filled squares) and apoE-/- (black filled circles) mice before

and following mrTBI. # and § in C and D denote significant differences between the

latencies of WT and apoE-/- mice at the respective time points. Cohorts were: Untreated

WT mice: (1 d=34, 2 d=34, 7 d=24; 14 d=14); GW3965-treated WT mice: (1 d=35, 2

d=35, 7 d=25, 14 d=14); untreated apoE-/- mice: (1 d=24, 2 d=24, 7 d=20, 14 d=9), and

134

GW3965-treated apoE-/- mice: (1 d=27, 2 d=22, 7 d=20, 14 d=9). Data were analyzed

using two-way repeated measures ANOVA followed by a Bonferroni post hoc test. *:

p<0.05, **: p<0.01, ***: p<0.001, ****: p<0.0001, #: p<0.05, ##: p<0.01, §§: p<0.01.

2.4.3 GW3965 Prevents mrTBI-Induced Elevation of Endogenous Aβ Levels

We determined levels of soluble endogenous Aβ40 and Aβ42 species from ipsilateral

half brains of WT and apoE-/- mice at 2, 7, and 14 d after mrTBI. Compared to sham

controls, injured WT mice showed a sustained ~1.5-fold increase in Aβ40 (Figure 2.8 A,

open bars, p < 0.05, two-way ANOVA), and a transient ~1.7 fold increase in Aβ42 that

returned to baseline by 14 d post-injury (Figure 2.8 B, open bars, p < 0.05, two-way

ANOVA). In contrast to untreated controls, GW3965 suppressed the increase of both

Aβ40 and Aβ42 in WT mice (Figure 2.8 A and B, black bars, p > 0.05, two-way ANOVA)

as compared to the sham. Similarly, apoE-/- mice exhibited a sustained ~1.5-fold

increase in Aβ40 (Figure 2.8 C, open bars, p < 0.05, two-way ANOVA), and a transient

~1.4-fold increase in Aβ42 after mrTBI that returned to baseline by 14 d post-injury

(Figure 2.8 D, open bars, p < 0.05, two-way ANOVA). Surprisingly, GW3965 effectively

suppressed the increase in both Aβ40 and Aβ42 in apoE-/- mice (Figure 2.8 C and D,

black bars, p > 0.05, two-way ANOVA) as compared to the sham. The GW3965

treatment effect was significant in both, WT and apoE-/- mice for Aβ40 (Figure 2.8 A

and C, p < 0.05, two-way ANOVA), but only in WT mice for Aβ42 (Figure 2.8 B, p <

0.05, two-way ANOVA). Furthermore, compared to WT mice, loss of apoE did not lead

to significantly greater accumulation of Aβ species, with the single exception of Aβ40

levels 7 d post-mrTBI in both vehicle (Figure 2.8 A and C, #, p < 0.05, two-way ANOVA)

and GW3965-treated groups (Figure 2.8 A and C, §, p < 0.05, two-way ANOVA). APP

135

holoprotein and APP-CTFα levels remained unchanged in both untreated and treated

WT and apoE-/- mice following mrTBI (Figure 2.9, p > 0.05, two-way ANOVA),

indicating that increased APP processing does not account for the changes in Aβ levels.

136

Figure 2.8: GW3965 prevents mrTBI-induced accumulation of endogenous Aβ in WT and apoE-/- mice.

Total, soluble murine Aβ40 and Aβ42 levels were measured from ipsilateral half brains of

WT (A, B) and apoE-/- (C, D) mice, respectively, at 2, 7, and 14 d post-mrTBI. Data

from sham animals within each genotype were pooled (grey bars). Numbers inside the

bars indicate sample size. Asterisks above individual bars indicate significant difference

compared to the respective sham levels. *: p<0.05, **: p<0.01, ***: p<0.001, ****:

p<0.0001. # indicates a significant difference (p < 0.05) between Aβ40 levels of

untreated WT and apoE-/- brains at 7 d post-mrTBI. § indicates a significant difference

(p < 0.05) between Aβ40 levels of GW3965-treated WT and apoE-/- brains at 7 d post-

mrTBI. Data were analyzed by two-way ANOVA followed by a Bonferroni post hoc test.

137

Legend: S: sham-operated mice, gray bars, V: untreated mice, open bars, G: GW3965-

treated mice, black bars.

Figure 2.9: APP and APP-CTFα levels remain unchanged following mrTBI.

APP holoprotein (A, C) and APP-CTFα (B, D) protein levels in ipsilateral half brains

were analyzed by Western blotting, with representative blots shown for WT (A, B) and

apoE-/- (C, D) mice. Data are expressed as fold difference relative to sham values. Data

from sham animals within each genotype were pooled (grey bars). Numbers inside bars

138

indicate sample size. Data were analyzed by two-way ANOVA followed by a Bonferroni

post hoc test. Legend: S: sham-operated mice, gray bars, V: untreated mice, open bars,

G: GW3965-treated mice, black bars.

2.4.4 GW3965 Enhances ABCA1 Induction Aafter mrTBI:

To determine whether lipid mobilization pathways are induced in our mrTBI model, we

determined the levels of ABCA1, apoE, and LDLR in treated and untreated WT and

apoE-/- mice. Compared to sham controls, untreated WT brains showed a transient 1.6-

fold increase in ABCA1 levels on post-mrTBI day 7 (Figure 2.10 A, open bars, p < 0.01,

two-way ANOVA) that returned to baseline by 14 d. This finding is consistent to the

findings reported in a FP injury model (Cartagena et al., 2008) and suggests that

endogenous LXR agonists may be transiently induced after TBI. As ABCA1 is a

sensitive LXR target, GW3965 augmented this response, leading to a 1.5-2.4-fold

increase in ABCA1 protein levels over the time points examined (Figure 2.10 A, black

bars, p < 0.05, two-way ANOVA). These results are consistent with our previous

observations of ABCA1 upregulation in GW3965-treated APP/PS1 mice (Donkin et al.,

2010). In contrast to WT animals, untreated apoE-/- mice did not show a transient

elevation of ABCA1 after mrTBI (Figure 2.10 B, open bars, p > 0.05, two-way ANOVA),

suggesting that the endogenous signals that lead to ABCA1 induction after TBI may be

compromised in the absence of apoE. Nevertheless, similar to the WT mice GW3965

treatment significantly increased ABCA1 levels in injured apoE-/- mice (Figure 2.10 B,

black bars, p < 0.05, two-way ANOVA), suggesting that GW3965 may bypass apoE and

directly increase ABCA1 expression after mrTBI. Neither mrTBI nor GW3965 treatment

led to significant changes in apoE levels in WT mice (Figure 2.11 A, p > 0.05, two-way

139

ANOVA), a result consistent with the relative insensitivity of apoE compared to ABCA1

as a LXR target (Koldamova et al., 2005b; Zelcer et al., 2007; Donkin et al., 2010).

Similar to apoE, LDLR levels remained unaffected after mrTBI in both treated and

untreated WT (Figure 2.11 B) and apoE-/- (Figure 2.11 C) mice (p > 0.05, two-way

ANOVA).

Figure 2. 10: GW3965 augments ABCA1 levels in WT and apoE-/- mice following mrTBI.

ABCA1 protein levels were determined in ipsilateral half brains of WT (A) and apoE-/-

(B) mice following mrTBI using Western blots, with representative blots shown below

the graphs. Data are expressed as fold difference normalized to sham values. Data

from sham animals within each genotype were pooled (grey bars). Numbers inside the

bars indicate sample size. Asterisks above individual bars indicate significant difference

compared to the respective sham levels. *: p<0.05, **: p<0.01, ***: p<0.001, ****:

p<0.0001. Data were analyzed by two-way ANOVA followed by a Bonferroni post hoc

test. Legend: S: sham V: untreated mice, open bars, G: GW3965-treated mice, black

bars.

140

Figure 2.11: ApoE and LDLR levels are unaffected by mrTBI or GW3965.

Levels of apoE and LDLR protein in WT mice (A, B) and LDLR protein in apoE-/- mice

(C) were determined in ipsilateral half brains following mrTBI using Western blotting,

with representative blots shown on the right. Data are expressed as fold difference

normalized to sham values. Data from sham animals within each genotype were pooled.

Numbers inside bars indicate sample size. Data were analyzed by two-way ANOVA

followed by Bonferroni post hoc test. Legend: S: sham-operated mice, gray bars, V:

untreated mice, G: GW3965-treated mice.

141

2.4.5 ApoE is Required for GW3965-Mediated Suppression of Axonal Damage

After mrTBI

Histological assessment of axonal damage using silver staining revealed argyrophilic

structures in cell bodies and axons in several ipsilateral white matter regions, including

the corpus callosum, cingulum, external and internal capsules, and optic tracts, with

strikingly greater accumulation observed in apoE-/- mice (Figures 2.12, 2.13, and 2.14

arrows). Notably, intense argyrophilia was observed in the bilateral optic tracts of WT

and apoE-/- mice, suggesting the involvement of strong coup-contrecoup forces in our

model. In comparison, very mild silver staining was observed in grey matter regions

including the sensory-motor cortex directly under the impact site as well as the piriform

cortex, farthest from the impact site (not shown).

142

Figure 2.12: Loss of apoE exacerbates axonal injury after mrTBI.

Axonal damage following mrTBI was assessed with silver staining (arrows). The left

panel depicts representative images of silver-stained coronal sections at approximately

-1.58 mm from bregma (Paxinos and Franklin, 2001) of untreated WT (A) and apoE-/-

(B) mouse brains harvested at 7 d post-mrTBI. White matter areas with prominent silver

staining are indicated by the black squares. Injury location is indicated by the

arrowhead. The right panel (C-L) depicts representative 40x-magnified images of silver

staining in five white matter areas in the brains of untreated WT (C-G) and apoE-/- (H-L) mice harvested at 7 d post-mrTBI.

143

Figure 2.13: mrTBI leads to mild axonal damage in WT mice.

Axonal damage following mrTBI was assessed with silver staining. The far left column

depicts representative images of silver-stained coronal sections at approximately -1.82

mm from bregma (Paxinos and Franklin, 2001) of sham (A), untreated (B), and

GW3965-treated (C) WT brains harvested at 2 d post-mrTBI. Locations of the white

matter areas where silver staining was most prominent are indicated by black squares.

Injury location is indicated by the arrowhead. Columns on the right depict representative

40X-magnified images of silver staining in five white matter areas in sham-operated (D-H), untreated TBI (I-M), and GW3965-treated (N-R) brains of WT mice harvested at 2 d

post-mrTBI.

144

Figure 2.14: Loss of apoE exacerbates axonal damage after mrTBI.

Axonal damage following mrTBI was assessed with silver staining (arrows). The far left

column depicts representative images of silver-stained coronal sections at

approximately -1.82 mm from bregma (Paxinos and Franklin, 2001) of sham (A), untreated (B), and GW3965-treated (C) apoE-/- brains harvested at 7 d post-mrTBI.

Locations of the white matter areas where silver staining was most prominent are

indicated by black squares. Injury location is indicated by the arrowhead. The columns

on the right depict representative 40X-magnified images of silver staining in five white

matter areas in sham-operated (D-H), untreated (I-M), and GW3965-treated (N-R) brains of apoE-/- mice harvested at 7 d post-mrTBI.

145

In untreated WT mice, axonal damage was significantly elevated in the corpus

callosum, cingulum, external capsule, and internal capsule at 2 d post mrTBI as

revealed by the semiquantitative analysis of silver-stained images (Figure 2.15 A-D,

open bars, p < 0.05, two-way ANOVA). In this group, silver staining intensity returned to

baseline by 7 d, suggesting that endogenous neuronal repair pathways were efficiently

activated in these regions. More robust and sustained axonal damage was observed in

the optic tracts of untreated WT mice (Figure 2.15 E, open bars, p < 0.05, two-way

ANOVA). In WT mice, GW3965 effectively suppressed axonal damage in all regions

examined (Figure 2.15 A-E, black bars, p < 0.05, two-way ANOVA). Untreated apoE-/-

mice showed extensive axonal damage that peaked at 7 d post-injury and remained

elevated at 14 d in the corpus callosum, external and internal capsules, and optic tract

(Figure 2.15 F-J, open bars, p < 0.05, two-way ANOVA). In contrast to WT mice,

GW3965 failed to suppress axonal damage in apoE-/- animals (Figure 2.15 F-J, black

bars, p < 0.05, two-way ANOVA). In all regions examined, axonal damage was

significantly more extensive in apoE-/- mice compared to WT controls (Figure 2.15 F-J,

p < 0.05, two-way ANOVA; # and § denote comparison of WT and apoE-/- mice within

untreated and treated groups, respectively).

146

Figure 2.15: ApoE is required for GW3965 to suppress axonal damage after mrTBI.

Silver-stained images of white matter areas in WT and apoE-/- brains were analyzed

semiquantitatively using an arbitrary silver staining scale extending from 0 (<10%

argyrophilic structures covering the image field) to 3 (>70% argyrophilic structures

covering the image field). The bar graphs represent mean ± SEM silver stain intensity

score (arbitrary value) of WT (A-E) and apoE-/- (F-J) brains in the corpus callosum (A,

147

F), cingulum (B, G), external capsule (C, H), internal capsule (D, I), and optic tracts (E, J). Sample sizes were: sham, n=6/genotype (pooled); untreated mrTBI, n=5/time

point/genotype; GW3965-treated mrTBI, n=5/time point/genotype. Asterisks above

individual bars indicate significant differences compared to the respective sham values

(grey bars). *: p<0.05, **: p<0.01, ***: p<0.001, ****: p<0.0001. ### (p<0.001) and ####

(p<0.0001) represent significant differences in silver stain scores between untreated WT

and apoE-/- mice. § (p<0.05), §§ (p<0.01), §§§ (p<0.001), and §§§§ (p<0.0001)

represent significant differences in silver stain scores between GW3965-treated WT and

apoE-/- mice. Data were analyzed by two-way ANOVA followed by a Bonferroni post

hoc test. Legend: V- untreated mice, open bars, G- GW3965-treated mice, black bars.

2.4.6 Weight Drop TBI Model Produces Negligible Neuroinflammation

To investigate the extent to which our mrTBI model induces inflammation, we examined

post-injury microglial activation using Iba1 immunohistochemistry and measured IL-6,

TNFα, and MCP-1 levels in ipsilateral half brains. Examination of coronal sections

revealed uniform distribution of Iba1-immunoreactive microglia with negligible changes

in the immunoreactivity in the injured WT and apoE-/- mice in either the ipsilateral cortex

(Figure 2.16) or hippocampus (Figure 2.17) compared to sham controls. Further,

GW3965 treatment did not reduce Iba1 staining below baseline (Figure 2.16 and 2.17).

148

Figure 2.16: mrTBI does not induce microglial activation in the cortex.

Microglial activation in sham and injured WT and apoE-/- mice was assessed with Iba1

immunohistochemistry. The top panel depicts representative images of Iba1-stained

coronal sections at approximately -1.82 mm from bregma (Paxinos and Franklin, 2001).

The bottom panel depicts 10X-magnified images of ipsilateral cortex underlying the injury

site (indicated by the black rectangle). Legend: V- untreated mice, G- GW3965-treated

mice.

149

Figure 2.17: mrTBI induces negligible hippocampal microglial activation.

Microglial activation in sham and injured WT and apoE-/- hippocampi was assessed

with Iba1 immunohistochemistry. The top panel depicts representative images of Iba1-

stained coronal sections at approximately -1.82 mm from bregma (Paxinos and

Franklin, 2001). The bottom panel depicts 10X-magnified images ipsilateral

150

hippocampus (indicated by black rectangle). Legend: V- untreated mice, G- GW3965-

treated mice.

Although no animal in this study had any obvious skull fracture, approximately 4-8% of

mice showed evidence of microcontusions in the cortex directly under the impact site

(Figure 2.18, arrows). Increased Iba1 immunoreactivity was clearly associated with

these localized areas of more severe injury (Figure 2.18).

Figure 2.18: Pronounced microglial activation is localized only around contused areas.

In this study, approximately 4-8% of brains subjected to mrTBI showed micro contusions

in the absence of gross skull fracture. The left panel shows a Iba1-stained coronal section

at approximately -1.82 mm from bregma (Paxinos and Franklin, 2001) with a micro

contusion in the cortex below impact site (black square). The right panel shows

representative 10X-magnified images of Iba1-stained untreated WT and apoE-/- contused

cortices. Microcontusions are denoted by black arrows. Note the pronounced localized

activation of microglia around the contusion.

Consistent with this model inducing negligible microglial activation, IL-6 levels in the

ipsilateral brains were unchanged in both WT and apoE-/- mice, with or without

151

GW3965 (Figure 2.19, p > 0.05, two-way ANOVA). TNFα and MCP-1 levels were below

the detection limit (not shown).

Figure 2.19: Pronounced microglial activation is localized only around contused areas.

Interleukin- 6 (IL-6) levels were measured from ipsilateral half brains of WT and apoE-/- mice harvested at 2, 7, and 14 d post-mrTBI. Data from sham animals within each

genotype were pooled. Cohorts were: sham (WT, n=9, pooled; apoE-/-, n=10, pooled), 2

day (WT: untreated, n=5, GW3965-treated, n=5 ; apoE-/-: untreated, n=5, GW3965-

treated, n=5), 7 day (WT: untreated, n=4, GW3965-treated, n=4; apoE-/-: untreated, n=6,

GW3965-treated, n=6), and 14 day (WT: untreated, n=5, GW3965-treated, n=5; apoE-/-:

untreated, n=5, GW3965-treated, n=5) post-mrTBI. Data were analyzed by two-way

ANOVA followed by a Bonferroni post hoc test. Legend: Sham WT: open bars, sham

apoE-/-: grey bars, vehicle-treated WT: open striped bars, vehicle-treated apoE-/-: grey

striped bars, GW3965-treated WT: open stippled bars, GW3965-treated apoE-/-: grey

stippled bars.

2.4.7 Retrospective Power Analysis

In this study we did not perform prospective power calculations to determine the sample

size. A retrospective power analysis was thus carried out based on the collected data

152

using an online post-hoc power calculator (http://clincalc.com/stats/power.aspx). For

power calculations it was assumed that the post-TBI outcomes in apoE-/- mice will be

worse compared to those in the WT mice, hence any significant change in the WT mice

will likely indicate similar or more change in apoE-/- mice. In this study we there were

four factors that could influence the study outcome: genotype, injury status, treatment

and time point. Of these we hypothesized that the smallest effect will be of TBI and TBI

x treatment interaction in WT mice and thus a significant difference for these

interactions will likely indicate significant results for the other group comparisons. Based

on these assumptions the retrospective power analysis was carried for TBI effect and

treatment x TBI effect at 7 d time point.

The retrospective power analysis revealed that the study was adequately powered to

detect TBI effect in WT mice and was underpowered to detect treatment x TBI effect

although all the observed changes were statistically significant.

The results of power calculations for various post-TBI outcomes are summarized in the

Table 2.3.

Post-TBI Outcome

Time Point

TBI Effect Treatment x TBI Effect Observed Change

(%) N Power

Observed Change

(%) N Power

NOR 7d - - - 47 (Observed) 13 5.4

153

Rotarod WT 7d 10 24 73.7% 10

(Estimated) 24 9.6% (4.8%

change) Rotarod ApoE-/- 7d 16 20 95.5% 16

(Estimated) 20 5.4%

ABCA1 (WT) 7d 55 8 96.6% 71 8 25.8%

ABCA1 (ApoE-/-) 7d 26 5 29.4 68 5 63

Abeta 7d 15 12 99.1 15 8 95 Silver Stain

(WT-CC) 2d 100 5 83.4 50 5 93.8

Silver Stain

(ApoE-/-CC)

2d 107 2 84.3 50 5 4.7

Table 2.3: Summary of retrospective power calculations

The table summarizes results of retrospective power calculations for various post-TBI

outcomes that were assessed in the present study. All the observed changes reported

in this table were statistically significant.

2.5 Discussion

The goal of this study was to evaluate the ability of GW3965 to promote recovery in a

model of mrTBI specifically designed to mimic repeated concussion. We found that

therapeutic administration of GW3965 improved NOR performance, suppressed Aβ

accumulation, and reduced axonal damage after mrTBI. Loss of apoE exacerbated the

severity of motor impairment and axonal damage and eliminated the ability of GW3965

to restore NOR performance and to promote axonal recovery. These results are

consistent with the role of apoE in neuronal repair and synaptic restoration. ApoE levels

did not change after TBI or after GW3965 treatment, which suggests that injury severity

154

was not sufficient to elevate apoE as well as reflects the poor sensitivity of apoE as an

LXR target compared to ABCA1 (Donkin et al., 2010). However, it is possible that apoE

may show localized upregulation in regions with more severe damage where microglial

activation is pronounced. Moreover, it is also possible that GW3965-mediated lipidation

of the available apoE mass may have been sufficient to promote post-TBI recovery in

our model. Although we did not assess post-TBI lipidation status of apoE with or without

GW3965 treatment in the present study, future studies will test whether ABCA1-

mediated lipidation of apoE contributes to the beneficial effects of GW3965 after mrTBI.

We used a commonly-used weight drop model to induce CHI in mice. The rationale for

using this model was less invasive surgical manipulations including craniotomy that is

required for FPI and CCI models, and relative simplicity compared to FPI and CCI

model, and high throughput. In addition, being a CHI model, it is clinically more relevant

as about 90% of human TBI do not involve penetration of dura. Using the weight drop

device we established a TBI protocol wherein every mouse received two CHI spaced at

24 h. In our initial pilot studies, we found that for mice weighing 25 to 32 g, a drop height

(in cm) equal to the one third the body weight (in g) resulted in mild TBI with low (<

10%) incidence of skull fracture.

Surprisingly, apoE was not required for GW3965 to suppress the transient increase in

Aβ levels induced in our model. Further studies will be required to characterize these

apoE-independent pathways that promote Aβ clearance after TBI. This will be an

important endeavor, as axonal APP accumulation is a hallmark of TBI (Gentleman et al.,

155

1993; Sherriff et al., 1994b). Theoretically, the Aβ produced after TBI could trigger Aβ-

dependent toxic pathways that exacerbate damage (Johnson et al., 2010). In support of

this γ-secretase inhibitors, which block Aβ production, reduce cognitive and pathological

changes following CCI in mice (Loane et al., 2009). However, the relationship between

Aβ levels and TBI recovery is complex. In WT and AD mice, PBS-soluble Aβ levels in

brain ISF microdialysates decrease by up to 50% within 90 min after CCI with impact

depths ranging from 1-2 mm (Schwetye et al., 2010). Intriguingly, Schwetye et al. (2010)

found no evidence of transient increases in Aβ levels after CCI, which differs from our

findings and those from other investigators using relatively crude tissue lysates (see

references in the Schwetey et al. (Schwetye et al., 2010) paper/discussion). Intra-axonal

Aβ accumulation is observed in 3XTg-AD and APP/PS1 mice from 1-24 h after

moderate-severe CCI (Tran et al., 2011b; Tran et al., 2011c), again suggesting that

distinct pools of Aβ may differ in their responses after injury. Future studies are needed

to clarify the factors that influence Aβ dynamics after TBI, including apoE.

Although our mrTBI model clearly leads to behavioral deficits and axonal damage, it is

not severe enough to trigger significant inflammation except in localized regions near

microcontusions. Importantly, inclusion or exclusion of mice with microcontusions did

not influence the group outcomes for NOR, rotarod, and silver stain data. These results

suggest that Iba1-positive staining of reactive microglia may be a potent secondary

effect of structural brain damage, although it is possible that the time points examined in

this study may have missed a peak of more global inflammation in our model.

156

Interest in understanding the pathophysiology of mild closed-head TBI has led to the

development of various impact and blast models. However, there is considerable

variation in outcomes. For example, many groups have reported mild cognitive deficits

without motor impairment in weight drop-based mild TBI mouse models (Tang et al.,

1997; DeFord et al., 2002; Pan et al., 2003; Creeley et al., 2004; Zohar et al., 2011). In

contrast, the Shohami group observed motor deficits in both severe and mild TBI, but

found that severe injury was required for cognitive impairment (Tsenter et al., 2008).

Using an electromagnetically-controlled piston to produce mrTBI, Shitaka et al observed

cognitive deficits, white matter damage, and robust microglial activation without

significant structural damage or APP immunoreactivity (Shitaka et al., 2011). Our mrTBI

model also shows sustained cognitive and transient motor deficits, axonal damage, and

transient Aβ accumulation. The most noteworthy difference between our model and that

of Shitaka et al is that we observed very little microglial activation, suggesting that

marked inflammation is not required to lead to behavioral and axonal changes. Our

model also does not lead to significant increases in apoE or LDLR levels, which have

been observed in models of moderate-severe TBI (Poirier, 1994; Iwata et al., 2005;

Petit-Turcotte et al., 2007; Loane et al., 2011).

We and others have demonstrated that synthetic LXR agonists effectively enhance the

ability of lipidated apoE to reduce Aβ levels and restore cognitive function in AD mice

(Riddell et al., 2007; Donkin et al., 2010). Given that TBI also involves altered apoE and

Aβ metabolism, it is of considerable interest to determine whether LXR agonists also

have potential therapeutic benefits for TBI. The importance of addressing this question

157

is two-fold. First, LXR treatment may minimize neuronal damage and promote acute

recovery by reducing inflammation and promoting neuronal repair. Second, by

facilitating Aβ clearance in the first weeks after injury, LXR treatment may also reduce

the increased long-term risk of AD years or decades later. Although current LXR

agonists have safety issues such as hypertriglyceridemia and hepatic steatosis that

preclude their present clinical use (Grefhorst et al., 2002; Groot et al., 2005), ongoing

drug discovery efforts may lead to a safe and effective compound. Furthermore, the

metabolic risks of LXR agonists may be clinically tolerable if short-term treatment could

improve functional recovery as well as decrease long-term AD risk. Building on the

findings of Loane and colleagues (Loane et al., 2011), our study provides additional

support for the potential of LXR agonists to treat mrTBI.

Our study also illustrates that the mechanisms by which LXR agonists promote recovery

after TBI may not directly correspond to their effects in AD models. For example,

outcomes such as cognitive recovery and axonal damage are apoE-dependent, but

surprisingly, Aβ clearance is independent of apoE in our model. Additionally, because

our model produces negligible inflammation, the beneficial effects of LXR agonists on

TBI recovery may be independent from their established anti-inflammatory effects.

Future investigations will be designed to identify the LXR targets operative in our mrTBI

model, assess their efficacy to reduce tau pathology, and evaluate their therapeutic

utility in models of moderate-severe TBI.

158

One caveat of this study was the lack of prospective power analysis to determine

adequate sample size. We therefore performed retrospective power calculations and

found that the study was adequately powered to capture TBI effect. It should however

be noted that the use retrospective power calculations is controversial (Thomas, 1997).

For example, retrospective power analysis is generally carried out only for the negative

study results and as such will produce a low post-hoc power result. This may be

misinterpreted as the study having inadequate power, which is reflected by the low

power observed in the present study to detect TBI x treatment effect, although the

interaction was significant. Based on the results of the present studies future studies will

be designed to include prospective power analysis.

A caveat of the weight drop TBI model used in this study is the absence of a systematic

biomechanical assessment of injury. Although input parameters such as weight mass

and drop height are reported in most CHI studies including ours, the reproducibility of

these mechanical inputs and the response of the animal’s head to the forces applied are

not well understood. Li and colleagues recently studied the biomechanical parameters

of the weight-drop based Marmarou rat TBI model (Marmarou et al., 1994) and found

that DAI severity was related to the linear and angular response of the rat head but not

with the drop height (Li et al., 2011). The importance of free head acceleration after

application of mechanical forces was further underlined by a recent study by Goldstein

and colleagues who demonstrated that a single blast injury in mice can generate

significantly impaired memory performance, long-term potentiation, and axonal

conductance accompanied by tau pathology, myelinated axonopathy,

159

microvasculopathy, and neuroinflammation (Goldstein et al., 2012). Intriguingly,

immobilizing the animal’s head to prevent blast-induced head oscillations prevented

memory deficits (Goldstein et al., 2012). It is possible that the very mild pathology in our

model is due to relatively little head movement after impact.

The caveats associated with the weight drop model inspired us to develop a better TBI

model which can allow unrestricted head movement and integrate biomechanics with

TBI outcomes. In the next chapter of this thesis I have described the novel rodent

neurotrauma model called CHIMERA (Closed-Head Impact Model of Engineered

Rotational Acceleration) that we developed in collaboration with the biomechanical

engineers at the International Collaboration on Repair Disorders (ICORD), The

University of British Columbia.

160

Chapter 3: Merging Pathology and Biomechanics Using CHIMERA: A Novel,

Surgery-Free, Closed-Head Impact Model of Engineered Rotational Acceleration

3.1 Summary

Traumatic brain injury (TBI) is a major health care concern that currently lacks any

effective treatment. Despite promising outcomes from many preclinical studies, clinical

evaluations have failed to identify effective pharmacological therapies, suggesting that

the translational potential of preclinical models may require improvement. Rodents

continue to be the most widely used species for preclinical TBI research. As most

human TBIs result from impact to an intact skull, closed head injury (CHI) models are

highly relevant, however, traditional CHI models suffer from extensive experimental

variability that may be due to poor control over biomechanical inputs. Here we describe

a novel CHI model called CHIMERA (Closed-Head Impact Model of Engineered

Rotational Acceleration) that fully integrates biomechanical, behavioral, and

neuropathological analyses. CHIMERA is distinct from existing neurotrauma model

systems in that it uses a completely non-surgical procedure to precisely deliver impacts

of prescribed dynamic characteristics to a closed skull while enabling kinematic analysis

of unconstrained head movement. In this study we characterized head kinematics as

well as functional, neuropathological, and biochemical outcomes up to 14d following

repeated TBI (rTBI) in adult C57BL/6 mice using CHIMERA. Head kinematic analysis

showed excellent repeatability over two closed head impacts separated at 24h. Injured

mice showed significantly prolonged loss of righting reflex and displayed neurological,

motor, and cognitive deficits along with anxiety-like behavior. Repeated TBI led to

161

diffuse axonal injury with extensive microgliosis in white matter from 2-14d post-rTBI.

Injured mouse brains also showed significantly increased levels of TNF-α and IL-1β and

increased endogenous tau phosphorylation. Repeated TBI using CHIMERA mimics

many of the functional and pathological characteristics of human TBI with a reliable

biomechanical response of the head. This makes CHIMERA well suited to investigate

the pathophysiology of TBI and for drug development programs.

3.2 Introduction

Traumatic brain injury (TBI) is a leading worldwide cause of death and disability for

persons under 45 years of age with a cost to society of over 17 billion USD per year. In

the United States, the overall incidence of TBI is estimated to be 538 per 100,000

persons, which represents at least 1.7 million new cases per year since 2003

(Gerberding and Binder, 2003; Langlois et al., 2006; Faul et al., 2010). TBI incidence is

reportedly lower in Europe (235 per 100,000) and Australia (322 per 100,000) (Cassidy

et al., 2004; Tagliaferri et al., 2006) although recent epidemiological data suggests far

greater incidence (749 per 100,000) (Feigin et al., 2013). Mild TBI (mTBI), which

includes concussion, comprises over 75% of all TBIs (Gerberding and Binder, 2003).

The reported incidence of mTBI is rising and mTBI is increasingly recognized as an

injury for which medical attention should be sought.

Falls are the most prevalent cause of TBI, although motor vehicle accidents and

impacts against objects are also common causes (Cassidy et al., 2004; Faul et al.,

2010; Roozenbeek et al., 2013). TBI resulting from high-contact sports such as boxing,

162

American football, ice hockey, soccer, and rugby account for almost 21% of all head

injuries among children and adolescents, particularly for mTBI (Centers for Disease

Control and Prevention, 2007). In these situations, the skull experiences an impact

resulting in brain deformation and resulting injury that most often occurs without skull

fracture. TBI is also considered a “signature injury” in modern warfare, as approximately

20% of veterans from the Iraq or Afghanistan wars are reported to have experienced a

TBI, 80% of which involve both blunt impact and overpressure mechanisms (Taber et

al., 2006; Hoge et al., 2008; Elder and Cristian, 2009; Sevagan et al., 2013).

Furthermore, the growing awareness that mTBI may have long-lasting and severe

consequences (Gavett et al., 2011; Jordan, 2013; Rusnak, 2013; Smith et al., 2013)

highlights the urgency to understand much more about the acute and long-term

consequences of brain injury.

The rapid accelerations and rotations during impact TBI lead to vigorous movement and

deformation of brain tissue within the skull that can result in contusions as the brain

contacts the interior of the bony skull. These inertial and contact forces directly affect

neurons, glia, and blood vessels producing a primary injury that initiates secondary

processes within hours to weeks after the initial injury (McIntosh et al., 1996;

Blumbergs, 1997; Davis, 2000; Giza and Hovda, 2001; Werner and Engelhard, 2007;

McAllister, 2011). These secondary changes lead to a plethora of events including

edema, raised ICP, impaired CBF, increased BBB permeability, inflammation, axonal

injury, calcium influx, elevated oxidative stress, free radical-mediated damage,

excitatory neurotransmitter release, and cell death (McIntosh et al., 1996; Blumbergs,

163

1997; Davis, 2000; Giza and Hovda, 2001; Werner and Engelhard, 2007; McAllister,

2011). Although very few treatment options are available for the primary injury,

secondary injury pathways are potentially modifiable (Graham et al., 2000). An

increasingly wide variety of experimental animal models are therefore being developed

to characterize secondary injury processes and for the evaluation of candidate

therapeutic approaches.

No single animal model can replicate the entire spectrum of human TBI

pathophysiology; therefore, several large and small animal models have been

developed to mimic particular aspects. Popular rodent models include OHI models

including FP and CCI systems, and CHI models that use either gravity or mechanical

methods to impact the intact skull (reviewed in Chapter 1). Although FP and CCI models

employ highly reproducible mechanical inputs and can mimic many pathological

features of human TBI, the prominent tissue destruction and lack of head movement

decreases the resemblance to the majority of human injuries that occur due to impact

and/or acceleration on an intact skull. In contrast, CHI models employ methods that do

not generally cause overt brain tissue loss and can also allow rapid behavioral

assessment of injury severity. As such, CHI models are considered by some to better

mimic the majority of human TBI. However, a major limitation of most current CHI

animal models is that the input parameters used to induce injury (e.g., mechanical

loading, method of mechanical input, response of the animal’s head to mechanical

loading) are often poorly controlled and this can lead to considerable experimental

164

variation across cognitive, histological, and biochemical outcome measures (Namjoshi

et al., 2013a).

Here we report a novel neurotrauma model called CHIMERA (Closed-Head Impact

Model of Engineered Rotational Acceleration). CHIMERA was developed to address the

absence of a simple and reliable model of rodent CHI that mimics the majority of human

TBI cases. CHIMERA is distinct from existing neurotrauma model systems in that it is

completely surgery-free and fully integrates biomechanical, behavioral, and

neuropathological analyses after delivering impacts of defined energy to a closed skull

with unconstrained head motion after impact. We show that that repeated TBI (rTBI) in

mice using CHIMERA reliably induces motor deficits and anxiety-like behavior and

leads to diffuse axonal injury (DAI) with extensive white matter inflammation and

increased phosphorylation of endogenous tau.

3.3 Materials and Methods

3.3.1 CHIMERA Impactor

The CHIMERA impactor consists of an aluminum frame that supports an animal holding

platform above a pneumatic system platform (Figure 3.1 A).

3.3.1.1 Animal Holding Platform

The animal holding platform is composed of a fixed head plate that supports the

animal’s head in a supine position and a body plate that positions and secures the

animal’s torso. The head plate has a hole through which the tip of the impactor piston

165

projects to impact the animal’s head. A cushion made from closed-cell foam surrounds

the hole to minimize rebound impact when the animal’s head falls back on the head

plate following impacted by the piston. Two perpendicular lines across the piston hole

act as crosshairs for aligning the animal’s head over the hole. The body plate holds a

restraint system consisting of an animal bed of closed-cell foam contoured to the shape

of the animal’s body and two Velcro straps. The animal holding platform is attached to

the frame by hinges and its angle of inclination can be adjusted. In this study, the angle

was set to approximately 32° such that the frontal and parietal bones laid flat over the

hole in the head plate, thus delivering impact to the dorsal cortical region.

3.3.1.2 Pneumatic Impactor System

The pneumatic impactor system includes a 1.9 L accumulator air tank (Model F9124,

Firestone Industrial Products, Indianapolis, IN), pressure regulator (Model 700-BF,

ControlAir Inc., Amherst, NH), a digital pressure gauge (Model DPG409-150G, Omega,

Laval, QC), a two-way solenoid valve (Model 21HN2KY110, Granzow Inc., Charlotte,

NC), and a trigger button. The pressure regulator and digital pressure gauge allow

precise adjustment of air pressure with minimum increments of 0.1 psi (0.69 kPa),

enabling accurate delivery of piston velocity and impact energy (Figure 3.2).

3.3.1.3 Impact Piston and Barrel System

Impact is induced with a free-floating metal impact piston with a body diameter of 15

mm and a tip diameter of 5 mm whose trajectory is constrained to linear motion by a

stainless steel barrel (internal diameter: 16 mm, external diameter: 20 mm). Four

166

different masses of impact pistons are available with the current system: 50 g, 75 g and

100 g pistons machined out of a chrome-coated steel rod, and a 25 g piston machined

out of an aluminum rod, to allow impact velocity or impact energy to be independently

adjusted (Figure 3.1 B). Only the piston tip (length: 2 cm, diameter: 0.5 cm) extrudes

from the barrel. The barrel design includes a series of holes drilled along the length to

vent air as the piston moves towards the open end. The piston is accelerated by a

controlled pulse of compressed air along the length of the barrel until it clears the

venting holes. The barrel can be mounted in either a vertical or horizontal configuration

relative to the animal holding platform. In the vertical configuration, the barrel is located

below the supine animal’s head such that the impact cylinder strikes the head to initiate

flexion. In this configuration the animal holding platform can be inclined relative to the

head plate such that the frontal and parietal bones lie flat over the hole in the head

plate, thus delivering impact to the dorsal cortical region. Alternatively, the animal

holding platform can be adjusted to lie parallel to the head plate, such that the impact is

directed towards the interparietal and occipital bones. In the lateral configuration the

barrel is mounted orthogonally relative to the head plate and the body is plate is inclined

to keep the head flat over the hole. In this configuration, the piston travels horizontally to

impact the left side of the head and initiate lateral and twisting head movements.

167

Figure 3.1: CHIMERA impactor and impact pistons.

(A) CHIMERA with various parts labeled as follows: 1. head plate, 2. body plate, 3.

animal bed, 4. Velcro straps, 5. air tank, 6. air pressure regulator, 7. digital pressure

gauge, 8. two-way solenoid valve, 9 vertical piston barrel. B) Impact pistons. CHIMERA

impactor can use impact pistons of four different masses, a 25 g piston constructed of

aluminum and 50 g, 75 g and 100 g pistons constructed of chrome-coated steel.

3.3.1.4 CHIMERA Calibration

The impactor was first calibrated by determining the air pressure-piston velocity

relationship by measuring the exit velocity of the piston at various air pressures (0.5, 1,

1.5, 2, 3, 5, 7, 10, 15, 20, 30, 50, 70, 100, and 130 psi). The velocity was measured in

triplicate at each pressure value for all four piston masses in both horizontal and vertical

orientations. Each impact event was recorded by a high-speed video camera at 10,000

fps. Video motion analysis software (TEMA Motion, Image Systems AB, Sweden) was

used to track the trajectory of a specific point on the impact piston. The calibration

curves for all four piston masses showed an inflection point between 10 and 15 psi such

168

that they were concave upwards before this point and concave downwards after the

point. The cause of this is unknown, but it is believed to be due to the interaction

between friction, drag, and piston velocity. Because of this effect, two 2nd-order

polynomial curves were used to fit the data. An R2 value of > 0.995 was found for all

curves. A representative calibration curve of impact energy for the 50 g piston at air

pressure values between 0 and 12 psi is shown in the Figure 3.2. Using these curves,

the desired impact velocity or energy can be independently interpolated. By choosing

the appropriate combination of piston mass and air pressure, impacts of input energy

ranging from 0.01 J to 18 J can be precisely generated.

Figure 3.2: Piston energy-air pressure calibration curve.

Air pressure-energy calibration curve was obtained by driving a 50 g piston at

increasing air pressure values and calculating the resultant impact energy. The graph

depicts energy measurements in triplicates for each air pressure value.

All experiments reported in the present Chapter were performed using a 50 g piston and

a vertically configured barrel. We initially performed a pilot experiment using three

169

impact energies (0.4, 0.5 and 0.6 J) to determine the maximum allowable impact energy

without leading to animal mortality. In the experiments described in this Chapter we

used impact energy of 0.5 J with the corresponding piston velocity of 4.5 m/s.

3.3.2 CHIMERA Repetitive TBI (rTBI) Procedure

All animal procedures were approved by the University of British Columbia Committee

on Animal Care (protocol # A11-0225) and were carried out in strict accordance with the

Canadian Council on Animal Care guidelines. Male C57Bl/6 mice (mean ± SD body

weight 34 ± 4 g) at 4-5 months of age were housed with a reversed 12 h light-12 h dark

cycle for at least 10 days and were subjected to repetitive TBI (rTBI) using CHIMERA

impactor as follows.

The animal was first anesthetized in an anesthesia induction chamber with 5%

isoflurane in oxygen (0.9 L/min). Once the animal lost righting reflex, it was quickly

transferred to the CHIMERA device and kept in a prone position on the animal bed. The

anesthesia was maintained with 2.5-3% isoflurane. Lubricating eye ointment was

applied to prevent corneal drying. Meloxicam (1 mg/kg) and saline (1 mL/100 g body

weight) were administered by subcutaneous injections for pain control and hydration,

respectively. The animal was then reoriented in a supine position in the animal bed. The

body plate was inclined such that the top of the animal’s head laid flat over the piston

hole in the head plate and the head was aligned using crosshairs such that the piston

struck the vertex of the head covering a 5 mm area surrounding the bregma (Figure 3.3

B and C). The animal’s body was secured with two Velcro straps (Figure 3.3 A).

170

Figure 3.3: Mouse head and impact position.

(A) Close-up view of animal strapped on the holding platform. (B) Location of impact

relative to the mouse head and brain in sagittal plane. P: impact piston. (C) Estimated

location of impact on the mouse skull (left) and brain (right) based on mouse brain atlas

(Paxinos and Franklin, 2001). The area covered by the impactor tip is shown by black

circle.

Impact was induced by pressing a trigger button that fires the piston and when

connected, simultaneously activates a high-speed camera to record the resulting head

trajectory. Isoflurane delivery was immediately stopped following the impact, the animal

was transferred to recovery cage and continuously monitored until fully ambulatory.

Twenty-four hours after the first impact, a second identical impact (including same drug

regimen) was delivered. Sham animals underwent all of these procedures except for the

171

impact. Isoflurane exposure time was measured for each impact from the start of

isoflurane delivery till isoflurane discontinuation. Approximately ~ 4 % of animals

displayed hind limb paralysis that lasted for several hours after TBI and were

euthanized. These animals were excluded from the data analysis. Sham-operated and

rTBI mice were randomly assigned to one of three post-TBI assessment time points,

viz., 2, 7, and 14 days after the second injury. The time of the 2nd TBI was defined as

time zero.

3.3.3 High-Speed Videography and Kinematic Analyses

For kinematic analysis, a cohort of 8 mice was subjected to rTBI and impact events

were recorded at 5,000 frames per second using a high-speed video camera (Q-PRI,

AOS Technologies, Switzerland). Head motion was tracked using two markers, one

being non-toxic paint applied on lateral side of head to mark the cheek area. Because

the skin is loose over the bony skull, we also marked the position of the maxilla by

wrapping dental floss positioned just caudal to the upper incisors around the animal’s

snout (Figure 3.4). Videos were analyzed using ProAnalyst Professional motion analysis

software (v 1.5.6.8, Xcitex Inc., Woburn, MA) as follows. The origin was set using the

paint marker in the Cartesian coordinates (Figure 3.4 A). The X and Y coordinates of

the position of each marker were tracked on a frame-by-frame basis and were

processed with a 400-Hz low-pass Butterworth filter to mitigate the noise in the recorded

measurements. Linear kinematic parameters were assessed from trajectories obtained

by tracking the paint mark, which was closest to the CG of the animal’s head. Velocities

and accelerations were determined by discrete differentiation of the marker position

172

data. Resultant linear velocity and acceleration were calculated as the magnitude of

their respective X and Y components using the following equations.

Linear Head Displacement = �𝑋𝑋𝑋𝑋𝑋𝑋𝑋𝑋2 + 𝑌𝑌𝑋𝑋𝑋𝑋𝑋𝑋2

Where, Xpos and Ypos represent X and Y positions of the paint mark from the origin,

respectively.

Linear Head Veloxity = �Xvel2 + Yvel2

Where, Xvel and Yvel represent velocities at X and Y positions, respectively.

Linear Head Acceleration = �Xacc2 + Yacc2

Where, Xacc and Yacc represent acceleration at X and Y positions, respectively.

Angular rotation of the head during TBI was determined by the angle of the line joining

the dental floss and the paint mark with the horizon (Figure 3.4 B). Immediately after

video capture, the dental floss was removed. The kinetic energy (KE) transferred from

the piston to the head was determined by KE = 0.5 x Me X ΔV2 (Viano et al., 2012),

where Me is the effective mass and approximated by head mass measured from the 8

mice (3.4 g) and ΔV is change in head velocity. A scaling factor λ [λ = (mass of human

brain / mass of mouse brain)1/3 = 13.8] was used to estimate the human head-

equivalent kinematic parameters from the animal data (Viano et al., 2009).

173

Figure 3.4: External markers used for mouse head tracking.

(A) Head movement was tracked using two markers: a dental floss (red arrow in the first

image) wrapped around the maxilla and a non-toxic paint (yellow arrow in the first

image) applied at the lateral size of the head. Head trajectories and linear kinematic

parameters were obtained by tracking the paint (yellow arrow) marker. Two dots

separated by 10 mm were marked on the CHIMERA frame and were used to calibrate

the distance. (B) Angle of head deflection and angular kinematic parameters were

obtained by determining the angle of the line joining the two tracking markers with the

horizon.

3.3.4 Behavioral Analyses

Duration of loss of righting reflex (LRR) was calculated as the time interval from

isoflurane discontinuation to the first sign of righting after each impact. Neurological

impairment was assessed using the neurological severity score (NSS, Table 3.1) (Flierl

et al., 2009) determined at 1 h and at 1, 2, and 7 d after the second TBI. Motor

performance was evaluated at 1, 2, 7, and 14 d after the second TBI using an

accelerating rotarod as described in Chapter 2. Open field activity was assessed at 1, 7

174

and 14 d after the second TBI using a Plexiglas box (14” x 24” x 14”). The floor of the

box was virtually divided into 60 equal squares using an overhead digital camera and

video tracking software (ANY-maze, v. 4.99, Stoelting Co, Wood Dale, IL). The field was

further subdivided into a peripheral zone along the walls of the open field consisting of

28 squares that surrounded a central zone consisting of 32 squares (Figure 3.5). The

animal was placed in the center of the box and spontaneous activity was recorded over

a 10 min epoch, including quantification of the total distance traveled, number of line

crossings, and immobile time. The thigmotaxis index (TI) was calculated as: TI = (TP-

TC)/(TP+TC) where TP and TC represent time spent in the peripheral and central zones,

respectively. Working memory was assessed by the passive avoidance task from 7 to

10 days after second TBI. Passive avoidance testing was conducted in a device that

consisted of two adjoining compartments, one illuminated (20.3 x 15.9 x 21.3 cm) and

one darkened (20.3 x 15.9 x 21.3 cm), divided by a guillotine-style door (Med

Associates Inc., St. Albans, VT). The floor of the compartments consisted of steel rods

capable of delivering an electric foot-shock. The electric shock was delivered by a

programmable animal shocker (Med Associates Inc.). Each session consisted of placing

mice into the illuminated compartment and using a timer to record the latency of the

mice to cross into the darkened compartment. On day 7 after the second TBI (training)

mice received an electric foot-shock (0.3 mA, 2 s) as soon as they crossed from the

illuminated into the darkened compartment. Following foot-shock, mice were removed

from the apparatus and returned to their home cage. On days 8-10 after the second TBI

mice were tested for memory retention. The latency for the mice to cross into the

darkened compartment was recorded. No shock was delivered during testing. Mice that

175

did not cross over into the darkened compartment were allowed to remain in the

illuminated compartment for the full 5 min and assigned a latency of 300 s. Non-working

memory was assessed with the Barnes maze from 8-13 d post-rTBI. Barnes maze

testing was conducted on a grey circular platform (91 cm diameter) with 20 circular

holes (5 cm diameter; Stoelting Co) at its periphery located in a 3.0 x 3.6 m room. An

escape box was positioned beneath one of the holes. Extra-maze visual cues consisted

of black and white images (cross, lightning bolt, smiley face) printed on glossy white

paper and placed on the walls surrounding the maze. Other visual cues included a cart

with a laptop and the experimenter. For motivational purposes, mice were subjected to

mild food deprivation (1 g of food per day) from days 8-13 after the second TBI and

maintained at 90% of free feeding body weight. During testing mice were exposed to

aversive stimuli in the form of two lights (100 W) positioned at either side of the maze

and a buzzer (2.9 kHz, 65 dB) which hung 10 cm above the maze. On day 8 mice

completed one 5-min habituation trial to become familiar with the maze environment

and to practice descending into the escape box. On days 9-13 mice were tested for

memory retention. Mice were given four trials per day (15-min inter-trial interval) for five

days. For each trial the animal was placed in a black plastic start tube (7 cm diameter,

12 cm height) at the center of the maze. After 10 s, the start tube was raised and the

buzzer was turned on to start the trial. Mice that were unable to locate the escape hole

in 90 s were gently guided to it. Mice remained in the escape box for 10 s before being

returned to their home cage. Mice were tracked during each trial using a digital camera

placed 4 m above the center of the maze and ANY-maze tracking system.

176

Figure 3.5: Open field thigmotaxis. The floor of the open field was virtually divided into central (gray area) and peripheral

(white area) zones. Thigmotaxis was assessed by placing the mouse in the open field

and measuring time spent in central and peripheral zones.

Task Description Points

177

(success/failure)

Exit circle Ability to exit a circle of 30 cm diameter within 3 min 0/1

Monoparesis/hemiparesis Paresis of front and/or back limbs 0/1

Straight walk Alertness, initiative and motor ability to walk straight 0/1

Startle reflex Innate reflex; the mouse will bounce in response to a loud hand clap

0/1

Seeking behavior

Physiological behavior as a sign of 'interest' in the environment 0/1

Beam balancing

Ability to balance on a beam of 7 mm width for at least 10 s 0/1

Round stick balancing

Ability to balance on a round stick of 5 mm diameter for at least 10 s 0/1

3 cm wide beam walk

Ability to cross a 30-cm long beam of 3 cm width 0/1

2 cm wide beam walk

Same task, increased difficulty on a 2-cm wide beam 0/1

1 cm wide beam walk

Same task, increased difficulty on a 1-cm wide beam 0/1

Maximal Score 10

Table 3.1: Neurological severity score (NSS) tasks.

Post-TBI neurological impairment is assessed by NSS, which is a composite of ten

different tasks that mice perform. Failure in each task earns one point, while passing the

task gets no point. Total NSS is obtained by adding scores of individual tasks. Based on

the total NSS, injury can be classified as mild (NSS: < 5), moderate (NSS: 5-7) and

severe/fatal (NSS: 8-10). Adapted from (Flierl et al., 2009).

3.3.5 Tissue Collection and Processing

For histological analyses, mice were anesthetized with an i.p. injection of 150 mg/kg

ketamine (Zoetis) and 20 mg/kg xylazine (Bayer) at 2, 7, and 14 d post-rTBI, and brains

were collected from perfused animals as described in the Chapter 2 except that 4%

178

paraformaldehyde rather than 10% neutral buffered formalin was used to post-fix

hemisected brain tissue for histology. For biochemical analyses, brains were harvested

as described in the Chapter 2 at 6 and 12 d and at 2, 7, and 14 d post-rTBI,

longitudinally hemisected and rapidly frozen over dry ice and stored at -80°C until

analysis.

3.3.6 Iba1 Immunohistochemistry and Silver Staining

Microglial activation and axonal injury were assessed using Iba-1 immunohistochemistry

and silver staining, respectively as described in Chapter 2. Microglial activation was

quantified using fractal analysis using ImageJ (NIH). Three to four microglia were

randomly chosen from 40X-magnified Iba-1 stained images, converted to 8-bit images,

and thresholded. Thresholded images were converted to outline images and analyzed

using the FracLac plugin (Karperien, A., FracLac for ImageJ.

http://rsb.info.nih.gov/ij/plugins/fraclac/FLHelp/Introduction.htm. 1999-2013.). The box

counting method was used and the mean fractal dimension was analyzed. The number

of microglia in each white matter regions was quantified by scanning the entire area

using a Olympus BX61 microscope at 10X magnification with a ProScan motorized XY

stage (Prior Scientific Inc, Rockland, MA). The component images were stitched

together into a montage using ImagePro Plus image analysis software (Media

Cybernetics Inc., Rockville, MD). The number of Iba1-positive cells was manually

counted in the entire white matter track ROI. The area of ROI was measured by ImageJ

as pixels and scaled to mm2. Cell density was finally expressed as number of Iba-1

positive cells per mm2. Silver staining intensity was quantified using ImageJ (version

179

1.48, NIH) on 40X-magnified images. The images were thresholded and the region of

interest (ROI) was selected. The ratio of area of positive signal in ROI to total ROI area

was reported as percent positive.

3.3.7 Biochemical Analyses

3.3.7.1 Tissue Processing

For protein determination, half-brains were homogenized in RIPA lysis buffer as

described in Chapter 2.

3.3.7.2 Assessment of TNF-α and IL-1β by ELISA

Endogenous TNF-α and IL-1β protein levels in the half-brain homogenates were

quantified by commercial ELISA kits (BD Biosciences OptEIA 559603 and 555268,

respectively) following the manufacturer’s instructions.

3.3.7.3 Quantitative Assessment of Phosphorylated and Total Tau by Simple

Western Analysis

Phosphorylated and total tau were assessed using a completely automated capillary

electrophoresis-sized-based Simple Western system (O'Neill et al., 2006) using the

WES machine (ProteinSimple, San Jose, CA). Simple Western is a gel-free, blot-free,

capillary-based, automated protein immunodetection system that automates all the

steps following sample preparation including sample loading, size-based protein

separation, immunoprobing, washing, detection, and data analysis. All procedures were

performed with manufacturer’s reagents according to the user manual. Briefly, 5 µl of

180

RIPA lysate (2 µg of protein) was mixed with 1.2 µl of 5x fluorescent master mix and

heated at 95°C for 5 min. The samples, blocking reagent, wash buffer, primary

antibodies, secondary antibodies, and chemiluminescent substrate were dispensed into

designated wells in the manufacturer provided microplate. Following plate loading the

separation and immunodetection were performed automatically using default settings.

Data were analyzed with Compass software (ProteinSimple). All tau antibodies were

kind gifts from Dr. Peter Davies (Albert Einstein College of Medicine, Manhasset, NY,

USA). The details of tau antibodies are presented in the Table 3.2. GAPDH (clone 6C5,

1:25000, Chemicon) was used as a loading control. Levels of p-tau and total tau were

normalized to GAPDH. Levels of phosphorylated tau were expressed as fold difference

compared to sham controls at the respective time points.

Antibody Tau Form Epitope(s) Concentration RZ3 p-Tau Thr231 1:25

PHF1 p-Tau Ser396 & Ser404 1:25

CP13 p-Tau Ser202 & Thr404 1:25

DA9 Total Tau 102 to 140 1:5000

Table 3.2: Details of monoclonal antibodies for probing p-tau and total tau.

The rTBI protocol and post-rTBI end points are summarized in Figure 3.6.

181

Figure 3.6: CHIMERA-rTBI procedure and post-rTBI endpoints

The figure indicates time line for rTBI/sham procedure and behavioral, biochemical and

histological end points at various post-rTBI time points.

BM: Barnes maze, Iba-1: Iba-1 immunohistochemistry, NSS: neurological severity score,

OF: open field behavior, PA: passive avoidance, RR: rotarod, SS: silver stain

3.3.8 Statistical Analyses

The head kinematics data and graphs are presented as mean ± 95% CI. Behavioral

data and graphs are presented as mean ± SEM. All other data and graphs are

presented as mean ± SD unless otherwise specified. NSS, LRR, thigmotaxis, and

rotarod data were analyzed using repeated measures two-way ANOVA followed by the

Holm-Sidak post-hoc test, as animals were tested repeatedly until sacrifice. Passive

avoidance and Barnes maze data were analyzed by repeated measures two-way

ANOVA. Iba-1 and silver staining data were analyzed by two-way ANOVA followed by

Tukey’s post-hoc test. For all the above statistical analyses, a p value of <0.05 was

considered significant. Tau phosphorylation and cytokine protein expression at each

post-rTBI time point was compared to the respective sham values by t test followed by

182

Bonferroni correction of multiple comparison, with p value set to <0.01 (5 comparisons)

for detecting statistical significance. Statistical analyses of behavioral data were

performed using SigmaPlot (version 12.5, Systat Software Inc.). Statistical analyses for

rest of the data were performed using GraphPad Prism (version 6.04, GraphPad

Software Inc).

3.4 Results

3.4.1 Head Kinematics Following CHIMERA rTBI

Analysis of high-speed videography (5000 fps) was used to assess the biomechanical

responses of the head in a group of 8 mice during CHIMERA rTBI at impact energy of

0.5 J. Peak kinematic parameters are depicted in Figure 3.7 and Table 3.3. Following a

vertical upward impact of a pneumatically-accelerated piston onto an intact skull of a

supine mouse, the head follows a looped trajectory in the sagittal plane (Figures 3.7 A

and 3.8). The average head trajectories following two TBIs were highly consistent

(Figure 3.7 A). The head traveled a peak linear displacement of 49.6 ± 3.5 mm in 15.7 ±

2.4 ms (Figure 3.7 B) and a peak angular rotation of 2.6 ± 0.28 rad in 24.8 ± 3.1 ms

(Figure 3.7 C). The peak linear velocity was 6.6 ± 0.8 m/s at 3.4 ± 1.0 ms (Figure 3.7 D),

and the peak angular velocity was 305.8 ± 73.7 rad/s at 2.8 ± 1.9 ms following initial

impactor contact (Figure 3.7 F). The head experienced large linear and angular

accelerations following impact, achieving a peak linear acceleration of 385.3 ± 52 g in

1.5 ± 0.3 ms (Figure 3.7 E) and peak angular acceleration of 253.6 ± 69.0 krad/s2 in 0.8

± 1.1 ms (Figure 3.7 G). As the head was stationary before impact, the change in head

velocity (ΔV) equals peak head velocity and was found to be 6.6 m/s. The energy

183

transferred from the piston to the head was 0.07 J, which equals to about 86% loss of

the input energy (0.5 J).

184

Figure 3.7: Head kinematics following CHIMERA rTBI.

Head kinematic parameters during impacts were assessed in 8 mice subjected to rTBI.

Data are represented as the mean ± SD for each impact. (A) Head trajectory in the

sagittal plane following impact. (B) Head displacement-time graph following impact. (C)

Head deflection is measured as the angle between the snout, side marker and the

horizontal plane. Linear head velocity and linear head acceleration are depicted in (D)

185

and (E), respectively. (F) and (G) show angular head velocity and angular acceleration,

respectively. All graphs are presented as mean ± SD.

Figure 3.8: CHIMERA allows unrestricted head motion during TBI.

Before impact, the mouse head was freely supported on a foam pad in the supine

position. Velcro straps were applied to the torso. Impact from the piston deflects the

head, which then subsequently returns to its original position on the foam pad. The

images were taken at 5,000 fps, at an angle perpendicular to the direction of impact and

along the mouse sagittal plane. Each image shown was 12 ms apart.

186

Kinematic Parameter 1st TBI (mean ± SD)

2nd TBI (mean ± SD)

Average of Two TBI % CV

Head displacement (mm) 48.4 ± 2.5 50.8 ± 4.1 49.6 8 Linear velocity (m/s) 6.3 ± 0.93 6.9 ± 0.94 6.6 5 Linear acceleration (g) 362.6 ± 42 407.6 ± 57.3 385.3 13 Angle of head deflection (rad) 2.6 ± 0.24 2.6 ± 0.31 2.6 9 Angular velocity (rad/s) 273.3 ± 82.9 338.4 ± 85.7 305.8 20 Angular acceleration (krad/s2) 216.2 ± 77.2 291 ± 69.7 253.6 21

Table 3.3: Summary of peak values of head kinematic parameters.

The table depicts average peak values of head kinematic parameters for rTBI from 8

rTBI mice. The coefficient of variation (CV) was calculated as the average of day 1 and

day 2 peak values from all available recordings.

Using the equal stress/equal velocity approach (Holbourn, 1943; Gutierrez et al., 2001;

Viano et al., 2009) to scale our murine kinematic data to human-equivalent values, ΔV

was found to be comparable to National Football League (NFL) values and higher than

Olympic boxing values, whereas scaled linear and angular velocity and acceleration

parameters were lower than NFL values but comparable to Olympic boxing values

Table 3.4.

Study TBI Type Species / Model

Head Kinematic Parameters (Scaled to Human) HIC15

187

ΔV (m/s)

Angular Velocity (rad/s)

Acceleration Linear

(g) Angular (krad/s2)

CHIMERA Impact-acceleration Mouse 6.6 22.6 27.9 1.4 40

Goldstein et al. (2012) Blast Mouse NA 36.2 NA 5.0 NA

Xiao-Sheng et al. (2000)

Non-impact rotation Rat NA 72.9 14.2 16.9 NA

Fijalkowski et al. (2007)

Non-impact rotation Rat NA NA NA 3.0 NA

Viano et al. (2009)

Impact-acceleration Rat NA NA 93.5 NA NA

Li et al. (2010)

Non-impact rotation Rat NA NA 14.7 1.1 NA

Wang et al. (2010)

Non-impact rotation Rat NA 6.4 NA 1.1 NA

Davidsson and Risling (2011)

Non-impact rotation Rat NA NA 32.8-

68.6 2.5-17.4 NA

Li et al. (2011) Weight drop Rat NA 4.3-5.7 60.5-

82.5 1.4-1.5 NA

Pellman et al. (2003a), Pellman et al. (2003b)

NFL concussion Human NA 34.8 97.8 6.4 250

Viano et al. (2005), Peng et al. (2013)

Olympic boxing – hook punch

Hybrid III dummy 3.1 29.3 71.2 9.3 79

Olympic boxing – uppercut

Hybrid III dummy 2.8 17.5 24.1 3.2 17

MVA involving pedestrian head impact (50% probability resulting in moderate injury)

Hybrid III dummy

and FEM NA NA 116 11.4 825

Table 3.4: Comparison of head kinematic parameters between rodent TBI models and human TBI.

The table compares most commonly-reported head kinematic parameters between

different TBI models in mice and rats. All the kinematic parameters are scaled to human

188

values according to the equal velocity/equal stress approach. According to this

approach, ΔV is unscaled, linear acceleration and angular velocity are scaled by 1 scale

factor (λ), and angular acceleration is scaled by λ2. The λ values for mice and rats were

13.8 and 11.0, respectively. National Football League (NFL) concussion values and

data from punches by Olympic boxers are included for comparison. Head Injury

Criterion for 15 ms (HIC15) is included to compare the likelihood of head injury. HIC15

for the CHIMERA was calculated based on the definition:

HIC = max

12

5.2

12)()(1 2

1

− ∫ ttdttatt

t

t

where linear acceleration (a) is in g and scaled by λ, time (t) is in seconds and scaled by

1/λ, and t1 and t2 are determined to give maximum value to the HIC function such that t2

- t1 = 15 ms.

FEM: Finite element model

NA: Data not available/not applicable

HIC15: Head Injury Criterion for 15 ms

ΔV: Change in velocity

MVA: Motor vehicle accidents

3.4.2 CHIMERA rTBI Induces Behavioral Deficits

Loss of righting reflex (LRR) in animals after TBI is considered analogous to loss of

consciousness in humans after TBI and can be considered as a behavioral indicator of

injury severity (Dewitt et al., 2013). Mice subjected to CHIMERA rTBI showed

significantly increased LRR duration compared to sham animals (Figure 3.9 A; TBI

effect: F(1, 65) = 59.61, p < 0.001). LRR duration was consistent between the first and

189

second impacts (Figure 3.9 A). We further assessed injury severity using the NSS,

which is a composite of ten tasks that assess motor reflexes, alertness, and

physiological behavior (Flierl et al., 2009). The NSS of injured animals was significantly

higher than sham mice from 1 h to 7d post-procedure (Figure 3.9 B, TBI effect: F(1,

191) = 44.12, p < 0.001). In injured mice, the NSS score showed maximum deficits at

1h post-procedure (p < 0.001) followed by steady spontaneous improvement over 1-7d,

albeit remaining significantly higher than sham animals at each post-rTBI time point

(Figure 3.9 B, p < 0.001). Similarly, rTBI significantly impaired motor performance from

1-7d post-injury as indicated by reduced fall latencies on an accelerating rotarod

compared to sham controls (Figure 3.9 C, TBI effect: F(1, 220) = 11.99, p < 0.001). Fall

latencies showed both time effects (F(4, 220) = 13.70, p < 0.001) and TBI x Time

interaction (F(4, 220) = 11.22, p < 0.001). Motor deficits in injured mice peaked at 1d (p

< 0.001) and returned to baseline conditions by 14d post-injury (p = 0.22), whereas

sham mice did not show any motor deficit (p > 0.82). Injured mice showed anxiety-like

behavior as indicated by significantly increased thigmotaxis in an open field test (Figure

3.9 D, TBI effect: F(1, 53) = 12.30, p < 0.001). The thigmotactic behavior of both groups

significantly declined over time in similar trend (Figure 3.9 D, Time effect: F(2, 53) =

5.45, p = 0.007; TBI x Time interaction insignificant). Open field thigmotaxis was not

affected by gross motor activity as no significant differences in total distance traveled or

time immobile were observed between injured and sham-operated mice (Figure 3.10).

Repeated TBI also induced working memory impairment as indicated by decreased

latencies to enter the darkened compartment on the passive avoidance task (Figure 3.9

E, TBI effect: F(1, 28) = 4.6, p = 0.041). For all mice, a main effect of Day indicated that

190

mouse behavior changed over the training and testing sessions evaluated in this study

(F(3, 84) = 58.55, p < 0.001). For the main effect of Day, pairwise comparisons

indicated that mice entered the darkened compartment significantly faster on Day 7 (p <

0.001) than on Days 8–10 (p > 0.05) signifying that all mice remembered the shock to

some degree (Figure 3.9 E). Injured mice also showed spatial reference memory

impairment as indicated by increased latencies to locate the escape hole on the Barnes

maze (Figure 3.9 F, TBI effect: F(1, 28) = 6.27, p = 0.018). Under our experimental

conditions, cognitive performance did not spontaneously resolve by the end of our

testing period.

191

Figure 3.9: CHIMERA rTBI induces behavioral deficits. (A) Duration of loss of righting reflex (LRR) was assessed immediately following the

sham or TBI procedure. Cohort size: Sham, N = 31; rTBI, N = 39. (B) Neurological

severity score (NSS) was assessed at 1 h, 1 d , and 2 d and 7 d post-rTBI. Cohort size:

Sham (1h: N = 34, 1d: N = 31, 2d: N = 35, 7d: N = 21); rTBI (1h, 1d and 2d: N = 42, 7d:

N = 25). (C) Motor performance was assessed on an accelerating rotarod at 1, 2, 7 and

14 d post-rTBI. Cohort size: Sham (1d and 2d: N = 35, 7d: N = 21, 14d: N = 15); rTBI

(1d and 2d: N = 41, 7d: N = 25, 14d: N = 15). (D) Thigmotaxis was quantified at 1, 7 and

14 d post-rTBI. Cohort size: Sham (1d: N = 23, 7d: N = 16, 14d: N = 15); rTBI (1d: N =

24, 7d: N = 10, 14d: N = 15). LRR, NSS, rotarod and tigmotaxis data were analyzed by

repeated measures two-way ANOVA followed by Holm-Sidak post-hoc test. Working

memory was assessed by passive avoidance test (E) from 7 to 10 d post-rTBI. Non-

working memory was assessed by Barnes maze (F) from 9 to 13 d post-rTBI. Data at

each time point represent mean of four trials. Barnes maze and passive avoidance data

192

(Cohort size: sham, N = 15; rTBI, N = 15) were analyzed by repeated measures two-

way ANOVA. For all graphs, * indicates significant rTBI effect within time point and #

indicates significant time effect within each group. *, **, *** and ***** indicate p<0.05,

p<0.01, p<0.001, and p<0.0001, respectively. #, ### and #### indicate p<0.05,

p<0.001, and p<0.0001, respectively.

Figure 3.10: CHIMERA rTBI does not significantly affect general mobility. General mobility was tested by the open field test at 1, 7 and 14 d post-injury. There

was no rTBI effect at any time point for the total distance travelled (A) and time spent

immobile (B). Data are presented as the mean ± SEM and analyzed by repeated

measures two-way ANOVA followed by Holm-Sidak post-hoc test. In all graphs, *** and

# indicate significant time effect for rTBI (p<0.001) and sham (p<0.05) mice,

respectively.

3.4.3 CHIMERA rTBI Induces Widespread Diffuse Axonal Injury

Silver staining was used to assess post-rTBI axonal damage at 2, 7, and 14 d (Figure

3.11). rTBI brains revealed widespread axonal injury, as indicated by intense punctate

and fiber-associated argyrophilic structures in white matter tracts including the olfactory

nerve layer of the olfactory bulb, corpus callosum, and optic tracts (Figure 3.11 A and

B). Axonal injury was observed at both coup (corpus callosum) and contra-coup (optic

193

tract) regions, indicating a diffuse injury pattern. High-magnification of the affected areas

at 100X revealed numerous axonal varicosities (Figure 3.11 C), which is a characteristic

histological feature of human axonal pathology after TBI (Johnson et al., 2013).

Quantitative analysis revealed significant silver uptake in the injured olfactory nerve

layer (TBI effect: F(1, 27) = 16.89, p < 0.0001), which was maximum at 2d (p < 0.001)

and returned to sham levels over 7-14d (Figure 3.12 A). On the other hand, rTBI

induced persistent silver stain uptake in the corpus callosum (Figure 3.12 B, TBI effect:

F(1, 37) = 41.54, p < 0.0001; Time effect insignificant) (Figure 3.12 C). In optic tract,

silver uptake was most intense (Figures 3.12 B and C TBI effect: F(1, 36) = 107.4) and

there was a significant time and injury interaction (TBI x Time interaction: F(2, 36) =

11.66), indicating persistent increase in axonal degeneration in contrecoup regions.

194

Figure 3.11: CHIMERA rTBI induces diffuse axonal injury.

Axonal degeneration was assessed by silver staining at 2, 7 and 14 d post-rTBI. (A) Coronal sections showing white matter areas including olfactory nerve layer, corpus

195

callosum and optic tract with prominent silver staining indicated by black rectangles. (B) Representative 40X-magnified images of the same brain regions in sham-operated

(upper panel) or rTBI-induced (lower panel) animals. (C) 100X-magnified images of the

same brain regions in rTBI-induced animals. Axonal varicosities are indicated by

arrows.

Figure 3.12: CHIMERA rTBI induces sustained axonal injury

Silver stained images were quantified by calculating % of region of interest (ROI) in the

white matter tract area stained with silver stain. The bars indicate mean ± SEM percent

of ROI showing positive signal in sham and rTBI-induced animals in (A) olfactory nerve

layer, (B) corpus callosum and (C) optic tract. Data were analyzed using two-way

ANOVA followed Tukey post-hoc test. For all graphs * indicates rTBI effect within each

time point and # indicates time effect within each group. For all graphs, *, **, *** and ****

indicate p<0.05, p<0.01, p<0.001, and p<0.0001, respectively and ### and ####

indicate p<0.001, p<0.0001, respectively.

3.4.4 CHIMERA rTBI Induces Widespread Microglial Activation

Using Iba-1 immunohistochemistry, we observed significantly increased activated

microglia throughout several white matter tracts including the olfactory nerve layer,

corpus callosum, optic tracts, and brachium of superior colliculus of injured brains

compared to the sham controls as assessed using both fractal analysis and microglial

196

density (Figures 3.13 and 3.14). Quantification of microglial morphology by fractal

analysis revealed that microglia in sham animals displayed highly ramified and

extensively branched processes that are characteristic of the resting state (Figure 3.13

C, upper row, Figure 3.14 A-D, open bars). By contrast, microglia in the corpus

callosum, brachium of superior colliculus, and olfactory nerve layer of injured animals

had predominantly hypertrophic to bushy morphology with primary branches only,

whereas those in the optic tract showed amoeboid morphology characteristic of highly

activated microglia (Figure 3.13 C, lower row, Figure 3.14 A-D, black bars). Quantitative

analysis showed significant and persistent microglial activation in the injured olfactory

nerve layer (TBI effect: F(1, 35) = 13.64, p = 0.0008), optic tract (TBI effect: F(1, 37) =

9.77, p = 0.0034), corpus callosum (TBI effect: F(1, 38) = 29.51, p < 0.0001), and

brachium of superior colliculus (TBI effect: F(1, 38) = 24.5, p < 0.0001) as soon as 2d

(Figure 3.14 A-D).

In addition to changes in microglial morphology, we observed significant increases in

the number of microglia in the same white matter regions including the olfactory nerve

layer (Figure 3.14 E, TBI effect: F(1, 16) = 21.53, p = 0.0003), optic tract (Figure 3.14 G,

TBI effect: F(1, 16) = 90.30, p = 0.0001), corpus callosum (Figure 3.14 F, TBI effect:

F(1, 19) = 22.25, p = 0.0002) and brachium of superior colliculus (Figure 3.14 H, TBI

effect: F(1, 18) = 34.85, p < 0.0001), indicating that injury induced proliferation or

recruitment of immune cells. In olfactory bulb (Time effect insignificant), corpus

callosum (Time effect insignificant) and optic tract, microglia cell number was

persistently increased from 2d up to 14d (p < 0.05). In the brachium of superior

197

colliculus, a delayed but persistent increase in the Iba-1 positive cell number was

observed from 7d to 14d (p < 0.01).

198

Figure 3.13: CHIMERA rTBI induces widespread microglial activation.

Microglial activation was assessed using Iba-1 immunohistochemistry at 2 d post-rTBI.

(A) Representative images of Iba-1 stained coronal sections of olfactory bulb and brain.

Areas with prominent microglial activation are indicated by black rectangles. (B) Representative 40X-magnified images of olfactory never layer, corpus callosum, optic

199

tract, and brachium of superior colliculus showing resting microglia in sham brains

(upper row) and activated microglia in injured brains (lower row). (C) Representative

100X-magnified images showing the morphology of Iba-1-stained resting microglia in

sham (upper row) and activated microglia in rTBI (lower row) brains.

200

Figure 3.14: Quantitative analysis of microglial response to rTBI.

Bar graphs in the left column (A-D) indicate mean ± SD fractal dimension for microglial

morphology in (A) olfactory nerve layer, (B) corpus callosum, (C) brachium of superior

201

colliculus, and (D) optic tract. Bar graphs in the right column (E-H) show mean ± SD

number of Iba-1 positive cells per mm2 in the same white matter regions. Data were

analyzed by two-way ANOVA followed by a Tukey post-hoc test. Numbers inside the

bars indicate sample size. For all graphs, * indicates a significant rTBI effect within a

particular time point while # indicates a significant time effect within rTBI group. *: p <

0.05, **: p < 0.01, ***: p < 0.001, ****: p < 0.0001. ##: p < 0.01, ####: p < 0.0001.

3.4.5 CHIMERA rTBI Increases Proinflammatory Cytokine Levels

In addition to the microglial response, we also measured protein levels of the

proinflammatory cytokines TNFα and IL-1β in the half brain homogenates. Protein levels

TNFα (Figure 3.15 A) and IL-1β (Figure 3.15 B) were significantly higher at 2 d post-TBI

compared to the respective sham levels (p<0.01, t tests with Bonferroni corrections).

Figure 3.15: CHIMERA rTBI increases proinflammatory cytokine levels.

Values are expressed as mean % fold change in TNFα (A) and IL-1β (B) levels in rTBI

brain lysates compared to cytokine levels in the sham brain lysates at respective time

points. For both graphs, ** indicates p<0.01 in comparison of rTBI vs sham, using t-test

with Bonferroni correction of multiple comparisons (p = 0.05/5 = 0.01).

202

3.4.6 CHIMERA rTBI Increases Endogenous Tau Phosphorylation

We next assessed the phosphorylation levels of endogenous murine tau using three

antibodies directed against different phosphorylation sites, namely CP13 (pSer202),

RZ3 (pThr231), and PHF1 (pSer396 and pSer404). In addition, total murine tau levels

were determined by the antibody DA9. Western blot analysis showed significantly

increased phosphorylation of all the probed epitopes in rTBI brain lysates at 6 h, 12 h

and 2d compared to the respective sham brain lysates (Figure 3.16 A-C and G-I,

p<0.01, t tests with Bonferroni correction). The change in tau phosphorylation was due

to a significant increase in ratio of phosphorylated tau:total tau, but not as a result of

change in total tau level (Figure 3.16 D-F and G-I, p<0.01, t tests with Bonferroni

corrections).

203

Figure 3.16: CHIMERA rTBI increases endogenous tau phosphorylation.

The graphs in the left column (A-C) depict fold change in endogenous phosphorylated

tau levels in rTBI half-brain homogenates compared to the sham brains using antibodies

CP13 (pSer202 and pThr205), RZ3 (pThr231) and PHF1 (pSer396 and pSer404),

respectively. Graphs in the middle column (D-F) depict quantitation of phosphorylated

tau as a proportion of total tau (DA9). Representative immunoblots of phosphorylated

and corresponding total tau are depicted in the right column (G-I). Arrows on the left of

the blots indicate position of 66 kDa marker. Data are presented as the mean ± SEM

204

fold change in rTBI compared to the respective shams at each time point. For all

graphs, *, ** and *** indicate p<0.01 for the comparison of rTBI vs respective sham

values, using multiple t-tests with Bonferroni correction (p = 0.05/5 = 0.01).

3.5 Discussion

The major goal of this study was to develop a simple, reliable model of murine CHI that

replicates fundamental aspects of human impact TBI through precise delivery of known

biomechanical inputs. CHIMERA fulfills these criteria and offers several key advantages

over comparable rodent TBI models.

The first and foremost advantage is CHIMERA allows completely surgery-free induction

of TBI eliminating the need of longer anesthesia duration as well as sedatives such as

ketamine and xylazine (that were used in the weight drop model), as well as learning

complex surgery such as craniotomy that is required for FP and CCI models. Almost all

of the currently available TBI models require some surgery to deliver the mechanical

force to either exposed skull (e.g. in weight drop TBI) or to the dura (as in FP and CCI

models) that increases the learning curve of the operator. For example, a novice

learning FP injury may struggle at the beginning to perform the craniotomy without

disrupting the dura as well as gluing the hub for attaching fluid injection tube without

introducing the glue into the craniotomy site (Alder et al., 2011). Moreover, head

securing method such as a stereotaxic setup used for CCI model and surgery

themselves increase the duration of injury procedure. For example, the duration

required for weight drop injury model discussed in Chapter 2 is ~ 20 min. In contrast,

using CHIMERA a fully-trained individual can finish the whole procedure in ~ 4 and ~ 8

205

minutes with and without high speed videography, respectively. Shorter injury procedure

also reduces the duration of anesthesia exposure. Duration of anesthesia exposure is

important as general anesthetic may influence injury outcomes. For example, isoflurane,

the most commonly used general anesthetic for rodent surgery is shown to exert

neuroprotective actions in TBI (Statler et al., 2000; Statler et al., 2006).

A second advantage of CHIMERA is that it allows unrestricted head movement after

impact replicating clinically encountered head impact situations as opposed to other TBI

models (e.g. weight drop) in which head support systems restrict head movement to

varying degree (Namjoshi et al., 2013a). As discussed in Chapter 1, head acceleration,

rather the impact itself is thought to induce more brain damage. In support of this,

CHIMERA-rTBI induces robust and sustained axonal damage and significant

neuroinflammation compared to the weight drop rTBI as discussed in Chapter 2.

Third advantage is the very high reproducibility of mechanical inputs (i.e. piston velocity

and impact energy) as exemplified by high R2 values (~ 0.99) of the piston velocity vs

air pressure calibration curve (Figure 3.2). This reflected in the high reproducibility in the

head kinematic parameters with a reasonable cohort size (N = 8). By fine adjustment in

the air pressure, CHIMERA allows precise control over mechanical input. Moreover, the

biomechanical inputs can be delivered by changing variety of parameters (impact

energy, location, and direction) that make CHIMERA a highly versatile model to study

TBI under a variety of mechanical input conditions. Thus, by choosing appropriate

combination of piston mass (25 g, 50 g, 75 g or 100 g) and air pressure, CHIMERA can

206

deliver impacts over a wide energy range from 0.01 J to 19 J. The impact can be

delivered either in vertical direction that results in head acceleration in the sagittal plane

or in the horizontal direction resulting head movement in an oblique plane. By adjusting

the angle between body and the head, the impact can be delivered at the vertex or on

the posterior aspect of the skull. By using the horizontal configuration, the impact can be

delivered to the lateral side of the head.

Lastly, the build-in mechanism of CHIMERA allows simultaneous high-speed video

recording that can be used to measure head kinematics following impact. This is very

important for the validation of impact procedures.

In this study, a 50 g piston delivered an input impact with a kinetic energy of 0.5 J.

Calibration curves show that the 50 g piston has an energy range of 0.01 J to 14.0 J

(useful range for mouse TBI is 0.1 J to 1 J) in minimum steps of 0.01 J, with highly

reproducible performance (Figure 3.2). Adjusting piston mass and dimensions will allow

other biomechanical parameters to be controlled, so that it may be possible to

experimentally model the biomechanical conditions observed under different types of

human impact TBI.

High-speed video integrated into CHIMERA enables measurement of mouse head

kinematic parameters that can be scaled to humans. The most-commonly used method

for scaling kinematic parameters between humans and animals is based on the equal

stress/equal velocity approach (Holbourn, 1943; Gutierrez et al., 2001; Viano et al.,

207

2009). For this certain assumptions are made such as: 1) neurons, blood and CSF all

have about the same density as water, 2) brain tissue is extremely incompressible, 3)

brain tissue has small modulus of rigidity and thus it is extremely easy to change shape

of the brain, and 4) human and animal brains have similar density and geometry, are

elastic and are isotropic (i.e. uniform in all directions) (Holbourn, 1943). According to

this principle, velocity does not scale and is the same for human and animal data. We

used a scaling factor λ [λ = (mass of human brain / mass of mouse brain)1/3 = 13.8] to

estimate the human head-equivalent kinematic parameters from our animal data.

Although this scaling approach has been widely used in the study of impact

biomechanics (Holbourn, 1943; Gutierrez et al., 2001; Viano et al., 2009), and has been

applied to a rat model (Viano et al., 2009), it is important to be cautious in its

extrapolation as the human and rodent brains differ in geometry, white : grey matter

ratio, ventricular volume and position, and cortical folding. All these factors render the

“scaling” of rodent data to human to be an approximation at best. Under our

experimental conditions, behavioral and neuropathological changes reliably occurred at

lower scaled values of all kinematic parameters (except for impact duration) than those

reported for NFL concussions (Pellman et al., 2003a; Pellman et al., 2003b). Because

clinical mTBI can occur under many different circumstances (e.g., falls, passengers and

pedestrians in motor vehicle accidents, non-NFL sports), an important area for future

research will be to determine how various impact characteristics lead to functional and

biochemical changes. The precision and flexibility CHIMERA offers with respect to

impact parameters will help to refine the relationship between impact characteristics and

physiological outcomes.

208

An additional advantage of CHIMERA is that it overcomes much of the variability

observed with most closed-head injury models. Typical weight-drop models (Marmarou

et al., 1994; Flierl et al., 2009) have poor control over biomechanical input parameters,

such as friction and air resistance inside the guide tube that may contribute to variable

outcomes, high incidence of skull fractures, and limited dynamic range, thereby posing

challenges in comparing results from different laboratories (Namjoshi et al., 2013a).

Head movement during many CHI impact models is often at least partially restricted by

anchoring the head within a stereotaxic frame or by resting the animal on a foam

support. In a recent modification, Kane et al supported mice on a piece of aluminum foil

that ruptures upon impact and leads to a 180° rotation of the animal (Kane et al., 2012).

While this modification allows unrestricted head movement, it still includes possible

sources of variability including stretching of the aluminum foil before yielding to the force

generated by the weight-drop and less reliable and adjustable positioning of impact

location compared to CHIMERA.

Loss of consciousness for <30 min is one of the clinical criteria for mTBI (Carroll et al.,

2004) and the analogous measure in mice is LRR. Interestingly, a second TBI occurring

24h after the first impact did not prolong LRR time. Post-concussion patients may also

show balance difficulties or postural instability up to several days (Guskiewicz, 2003;

Blume et al., 2011; Putukian, 2011), as well as mood changes such as irritability or

anxiety (Blume et al., 2011; Daneshvar et al., 2011). Though general locomotor

activities were not severely affected, CHIMERA TBI resulted in deficits in fine motor

209

coordination and neurological performance. Similar changes have also been reported

by other groups (Tsenter et al., 2008; Mouzon et al., 2012). CHIMERA also increased

thigmotaxis in TBI animals, suggesting an anxiety-like behavior (Simon et al., 1994).

Our model also revealed deficits in both working and non-working memory as assessed

by passive avoidance test and Barnes maze, respectively.

Human and experimental TBI induces rapid neuroinflammatory responses as

demonstrated by changes in cytokine levels (e.g., IL-1β and IL-6) and microglial

activation (Taupin et al., 1993; Shohami et al., 1997; Rooker et al., 2006; Shein et al.,

2007; Shitaka et al., 2011; Perez-Polo et al., 2013). Under our conditions, rTBI led to

elevated IL-1β and TNF-α levels at 2 d after injury, which was accompanied by

histological evidence of microglial activation. Because our animal ethics committee

required the use of meloxicam for pre-emptive pain control, it is possible that an

inflammatory response occurring during the first few hours after impact (Taupin et al.,

1993; Shohami et al., 1997; Shein et al., 2007) was suppressed (Engelhardt, 1996). Iba-

1-positive activated microglia were particularly evident along white matter tracts

throughout the brain, whereas grey matter was essentially spared. Microglial activation

as assessed by fractal analysis and cell density was significant at 2d and persistent until

14d. However, as Iba-1 does not distinguish the source of immune cells, the increase in

cell number in this study may be due to proliferation of resident immune cells in the

brain or recruitment of immune cells from the periphery, or both.

210

Diffuse axonal injury (DAI) is one of the characteristic pathologies of TBI (Adams et al.,

1989; Johnson et al., 2013). Using silver staining, we observed increased argyrophilic

fibers and punctate structures in several white matter tracts across the brain, suggesting

a diffuse pattern of damaged axons. Axonal varicosities, a classical feature of DAI in

humans, were also present (Johnson et al., 2013). Interestingly, both white matter areas

that were close to (e.g., corpus callosum) and distant from (e.g., optic tract and olfactory

nerve layer) the impact site were affected, suggesting coup and contra-coup injuries are

present in our model. Affected white matter areas, including the corpus callosum, optic

tract, and the olfactory system, have been reported in other CHI models that induce

impact at the superior side of skull (Foda and Marmarou, 1994; Creeley et al., 2004;

Shitaka et al., 2011; Namjoshi et al., 2013b). Several white matter areas, including

olfactory nerve layer, corpus callosum and optic tract showed both increased

argyrophilic staining and microglial activation, suggesting a possible relationship

between axonal damage and neuroinflammation, in agreement with previous reports

(Ciallella et al., 2002; Shitaka et al., 2011; Lin and Wen, 2013). Interestingly, axonal

injury in corpus callosum and optic tract continued to increase over 14 days while

microglial activation resolved in all white matter tract areas over 2-7 days. On the other

hand, axonal injury was resolved in the olfactory neuronal layer within 7 days

suggesting efficient neural repair/neuroregeneration in this white matter region. Future

studies will characterize the dynamics of post-rTBI axonal damage over long-term follow

up (up to 6 months).

211

Hyperphosphorylation of the cytoskeletal protein tau is a pathological event observed in

many neurodegenerative diseases including AD (Ballatore et al., 2007) and CTE

(McKee et al., 2013). In our model, we demonstrated that tau hyperphosphorylation is

an early event after rTBI in wild-type mice, again in agreement with other models of CHI

(Genis et al., 2000; Goldstein et al., 2012). It should, however be noted that post-TBI

changes in phosphorylation of murine tau does not predict whether human tau will show

similar dynamics. Further experiments using transgenic human tau mice will be required

to investigate the influence of rTBI on tau deposition.

Future studies will be conducted to characterize the relationships between kinematics

and the resulting behavioral and neuropathological responses across a variety of impact

parameters. The significant advantages CHIMERA offers over comparable rodent TBI

models are expected to facilitate the acquisition of preclinical data with improved

relevance to human TBI, thereby accelerating the pace of successful research to

understand the mechanisms of TBI and to develop effective therapeutic approaches for

this devastating condition.

212

Chapter 4: Conclusions and Future Directions

The experiments described in Chapter 2 of this thesis were designed to assess the role

of apoE in the recovery from TBI. As discussed in the introductory chapter, TBI, while a

pathological entity in itself, is also suggested to increase risk of dementia in TBI

survivors. Although the cause and effect mechanisms linking these two pathological

conditions are still unclear, several observations common to both diseases may help to

understand the relationship between the two. For example, post-mortem analyses of

traumatized human brains have revealed formation of diffuse Aβ deposits and NFTs,

which are the two pathological hallmarks of AD. Neurons are rich in lipids, which play

crucial role in neuronal function as well as synaptogenesis. Thus it is conceivable that

the brain lipid metabolism, which is based on apoE may serve important roles in

neuronal recovery following TBI. While the apoE4 isoform is the best established

genetic risk factor for AD, the association between apoE and TBI is still unclear.

Epidemiological data suggests that apoE may influence post-TBI recovery. The role of

apoE in AD pathogenesis stems from its role in Aβ metabolism, inflammation, and

synaptic remodeling. We and other groups have previously shown that ABCA1-

mediated lipidation of apoE is crucial for its role in Aβ clearance. Thus enhancing apoE

lipidation and function may be beneficial in AD pathology which is reflected by the

beneficial effects of LXR agonists on facilitating Aβ clearance and improving cognitive

function in AD mouse model.

213

Based on these observations we hypothesized that LXR agonists may have improved

recovery from TBI through apoE-based mechanisms. As described in Chapter 2 of this

thesis, we tested this hypothesis by treating wild-type and apoE-/- mice that were

subjected to mrTBI using popular weight drop method with a synthetic LXR agonist

GW3965.

4.1 Chapter 2: Conclusions

The first objective of this thesis was to test the therapeutic utility of synthetic LXR

agonist, GW3965 in mild, repetitive TBI (mrTBI), which is described in Chapter 2. I

found that GW3965 treatment improves NOR performance as well as promotes axonal

recovery following mrTBI and these outcomes require apoE. I further found that

GW3965 can also suppress the elevation of Aβ following TBI, although surprisingly this

outcome was apoE-independent. The overall results from this study suggest that

synthetic LXR agonists may be a promising pharmacological treatment for TBI.

Moreover, we observed treatment effects on the above-mentioned the post-TBI deficits

as early as two days following GW3965 administration, suggesting that LXR agonists

may be beneficial in the management of acute post-TBI phase.

We found that mrTBI induced significant NOR memory deficits in untreated WT as well

as apoE-/- mice that persisted from 2 to 14 days. While GW3965 treatment restored

NOR memory deficits in WT animals, it failed to show this beneficial effect in apoE-/-

mice. This indicates that apoE is required for the beneficial effect of GW3965 in post-

TBI cognitive deficits. Previously thought to assess pure working memory, NOR task is

214

now suggested to test episodic memory of the animal (Ennaceur, 2010). Although the

exact neural circuits for NOR task are not fully understood, studies indicate that

ventromedial prefrontal cortex (Akirav and Maroun, 2006) and hippocampus (Broadbent

et al., 2010) are essential for object recognition.

We tested TBI-induced motor deficits on an accelerating rotarod. The rationale for using

accelerating vs constant speed rotarod was that in the former method, the accelerating

rod constantly challenges the animal requiring motor coordination and planning while

the later may suffer from testing endurance rather than motor coordination (Deacon,

2013). Our TBI model induced impairment in finer motor tasks, such as motor

coordination and planning. While the exact mechanism for spontaneous recovery of

motor function observed in this study is presently unknown, it is possible that with the

repetitive rotarod testing after TBI we may have inadvertently introduced motor training

akin to post-TBI rehabilitation program in humans. It is well known that repeated motor

tasks promote development of spinal neural circuits, called central pattern generators

that when activated can produce rhythmic motor patterns such as walking in the

absence of sensory or descending inputs from brain (Marder and Bucher, 2001). The

above-mentioned possibility is further supported by a recent study, in which a 30 min

daily bilateral movement training in mice using the rotarod for 33 days improved motor

function and promoted axonal remodeling and rewiring of the corticospinal tract (the

major motor pathway that controls fine movements and posture) after unilateral

contusion of the motor cortex (Nakagawa et al., 2013). The authors further showed that

rotarod training increased neural activity in the spinal cord laminae VII, VIII and X. The

215

commissural interneurons located in these laminae of the rodent spinal cord are

involved in generating locomotor-like activity (Kjaerulff and Kiehn, 1996; Kiehn and

Kjaerulff, 1998). We further observed that loss of apoE further exacerbated the

magnitude of motor impairment following mrTBI. Intriguingly, GW3965 treatment neither

affected the severity of motor deficits nor the kinetics of spontaneous recovery. Future

studies will be necessary to assess specific motor pathways affected in our TBI model,

changes in the plasticity and remodeling of these neurons and role of apoE in these

processes.

Untreated WT brains showed a transient increase in ABCA1, the principal LXR target

protein that peaked at 7 days and dropped back to sham levels by day 14. It is

hypothesized that the damaged neurons release lipids that are taken up by the viable

cells as shown by increased levels of LDLR, a major apoE receptor with no change in

ABCA1 in first 3 days after CCI injury (Loane et al., 2011). In the same model, ABCA1

levels peaked at 7 days with corresponding decrease in LDLR (Loane et al., 2011). In

our study, while the LDLR levels were not affected at any post-TBI time point, the

ABCA1 dynamics was similar to that reported in the CCI model, i.e. with no change at 2

day post-TBI and peaking by day 7. This delayed increase ABCA1 may indicate

cholesterol efflux by the viable cells that can be used for synaptic repair and remodeling

during post-TBI recovery. Interestingly, unlike WT mice we did not see any changes in

ABCA1 levels in untreated apoE-/- mice following TBI. This suggests that the

endogenous signals leading to upregulation of ABCA1 might be compromised by the

absence of apoE. Tall and coworkers have shown that interaction of ABCA1 with apoA-

216

I, the peripheral counterpart of apoE, at the cell membrane inhibits calpain-mediated

degradation of ABCA1 that results in increase in the both cell surface expression and

activity of ABCA1 (Martinez et al., 2003). Extending this concept, it is also possible that

ABCA1 turnover is increased in the CNS leading to its rapid degradation in the absence

of apoE (Krimbou et al., 2004). As expected, GW3965 treatment caused sustained

increase in ABCA1 in both WT and apoE-/- mice.

We further found that neither TBI alone nor GW3965 treatment had any effect on apoE

levels in WT mice. This result is consistent with the relative insensitivity of apoE

compared to ABCA1 as a LXR target (Donkin et al., 2010). Since we determined apoE

levels in the whole half-brain extracts, it is possible that regional changes in the apoE

levels, if any may have been missed. While we did not assess the lipidation status of

apoE, it is also possible that even though the total apoE pool does not change per say

the ABCA1-mediated lipidation of the available apoE may have been sufficient to

promote recovery mechanisms after mild injury. More studies can be done to assess

whether the same holds true in case of moderate-to-severe TBI. Moreover, the

contribution of ABCA1 in LXR agonist-mediated recovery may be more important as

recently shown by our group that ABCA1 is required for the beneficial effects of

GW3965 in AD mice (Donkin et al., 2010). This hypothesis can be tested in future

studies using ABCA1-/- mice.

Surprisingly, contrary to our hypothesis GW3965 treatment blocked Aβ elevation even

in the absence of apoE. While, it is generally agreed that apoE modulates Aβ clearance

217

through enzymatic degradation as well as clearance through cerebrovasculature recent

data suggests that the interaction of apoE with Aβ and its influence on Aβ metabolism is

complex than thought (see section 1.6.4.7). Our observation indicates that LXR

activation may engage compensatory mechanisms in the absence of apoE to enhance

Aβ clearance. A potential candidate for such compensatory mechanism is apoA-I, the

peripheral counterpart of apoE. Although apoA-I is not synthesized in the brain, it is

present in the CSF at relatively high levels of 0.1%-0.5% relative to plasma levels,

suggesting that it crosses BBB (Stukas et al., 2014). A recent study from our lab has

shown that apoA-I levels remain significantly higher in the CNS than in plasma in

ABCA1-/- mice and that 8-week treatment of AD mice with GW3965 causes a dramatic

increase in the cerebral apoA-I levels (Stukas et al., 2011). In the present study we did

not observe any changes in brain as well as CSF apoA-I levels (data not shown) either

following TBI or with GW3965 treatment, which may be due to short duration of

treatment compared to long-term treatment (Stukas et al., 2011). Nonetheless, recent

studies suggest apoA-I as another apolipoprotein that may play important role in injured

brain for several reasons. First, AD mice that lack apoA-I show increased

cerebrovascular amyloid deposits and impaired spatial learning and memory function

(Lefterov et al., 2010) whereas overexpression of human apoA-I protects age-

associated cognitive deficits and lowers neuroinflammation and CAA (Lewis et al.,

2010). Interestingly, deletion or overexpression of apoA-I seem to affect Aβ and amyloid

deposits only in the cerebrovasculature and not in the brain parenchyma (Lefterov et al.,

2010; Lewis et al., 2010) suggesting that role of apoA-I may be more prominent in the

maintenance of the cerebrovasculature (Stukas et al., 2014). Second, in vitro studies

218

have shown that apoA-I binds to Aβ (Koldamova et al., 2001), prevents Aβ aggregation

and reduces Aβ-induced cytotoxicity, oxidative stress, and neurodegeneration (Lefterov

et al., 2010; Lewis et al., 2010). Lastly, serum apoA-I levels were recently shown to be

elevated in patients with mTBI and assessment serum apoA-I levels in combination with

S100B, an astrocyte-specific protein has been proposed to be a useful biomarker for

classification of mTBI (Bazarian et al., 2013). ApoA-I therefore may have more

important role in the CNS function and lipid metabolism than it is appreciated (Stukas et

al., 2014). Thus, future studies will be designed to get a deeper insight into the role of

apoA-I in the brain, which may offer novel approaches to promote neuronal recovery

after TBI.

4.2 Chapter 2: Study Limitations and Future Directions

While the results from this study indicate that synthetic LXR agonists may offer

beneficial therapeutic effects in TBI partly through apoE-mediated mechanism there are

several limitations to this study that may help guide future studies.

Although the LXR stimulation offers beneficial effects in mrTBI, current LXR agonists

have safety issues such as hypertriglyceridemia and hepatic steatosis (Grefhorst et al.,

2002; Groot et al., 2005) that has virtually precluded the development of LXR agonists

for chronic diseases. To date, only one Phase I study of a synthetic LXR agonist, LXR-

623, in humans has been reported (Katz et al., 2009). In this study, the safety,

pharmacokinetic and pharmacodynamic profile of LXR-623, was studied in healthy

human volunteers after a single ascending dose. Although the authors reported a

219

favorable safety profile with dose-dependent increase in ABCA1 and ABCG1

expression, further development of this compound was discontinued due to the

development of adverse CNS events at the highest doses used (150 and 300 mg),

including confusion, forgetfulness, decreased comprehension, lightheadedness,

drowsiness, paresthesias, palpitations, decreased concentration, altered time

perception and paranoid ideation (Katz et al., 2009). It is therefore of interest to develop

strategies that can retain the beneficial effects of LXR activation and circumvent their

unwanted side effects. For example, development of LXRβ-selective agonists is an

active area of research, as LXRα underlies much of the undesired lipogenic effects

(Quinet et al., 2006). Another promising approach is to develop tissue-specific LXR

agonists or antagonists, as the magnitude of induction of LXR response genes varies

considerably among tissues (Brunham et al., 2006). As shown in this study, the

beneficial effects of LXR agonist were seen even in the acute post-TBI phase, which

suggests that these agents may be useful for short-term rather than chronic treatment,

which could improve functional recovery as well as decrease long-term AD risk. While

studying long-term effects of LXR treatment was beyond the scope of this thesis project,

future studies could be designed for this. Finally, less potent compounds or compounds

with LXR-agonist like activity may offer another approach, as proton pump inhibitors

have some LXR activity in murine primary astrocytes and human glial-derived cell lines

(Cronican et al., 2010).

Our study suggests that apoE is not necessary for LXR-mediated suppression of

endogenous Aβ. While the amyloid plaques are one of the pathological hallmarks of AD,

220

the overall influence of Aβ in TBI pathology is questionable as amyloid deposits are

found in only 30% of the clinical TBI cases. On the other hand, tau

hyperphosphorylation is common neuropathology of TBI, which suggests that the

cascade of events that occur following TBI may lead directly to tau pathogenesis

independent of Aβ. This hypothesis is supported by a recent study where tau pathology

still occurs when Aβ production is blocked (Tran et al., 2011a). Although in the present

study we did not undertake systematic examination tauopathy using specific anti-tau

antibodies, the possibility of tauopathy may come from indirect evidence. For example,

silver uptake is associated with both 3-repeat and 4-repeat tau deposits (Uchihara et al.,

2005). Silver uptake into damaged axons in white matter tracts is by far the most

sensitive histological marker of injury in our weight drop mrTBI model as it is detectable

even in the absence of marked microglial activation or cell death. This suggests that

these white matter tracts may be sites where tau pathology begins after mrTBI. This

possibility is further supported by robust silver uptake along with increased tau

phosphorylation observed in CHIMERA-TBI (discussed in the next section). Moreover,

apoE-/- mice clearly show profoundly greater and prolonged silver staining of damaged

axons, suggesting that apoE’s role in repair of damaged neuronal membranes may

(in)directly influence pathways that lead to tau phosphorylation. Supporting this

hypothesis is our observation that promoting apoE function with GW3965 suppresses

silver uptake in WT but not in apoE-/- mice. The influence of apoE on tau

phosphorylation is unclear although the current evidence suggests that apoE may inhibit

tau phosphorylation by inhibition of tau kinases, including GSK3β (Ohkubo et al., 2003;

221

Hoe et al., 2006). Thus, it will be now important to determine the interaction between

murine as well as human apoE on tau phosphorylation in mrTBI.

Last but not least, although we used the widely-used weight drop model in this study,

this model has several caveats, including poor control over mechanical parameters and

variable injury outcomes (discussed in Chapter 1). Although currently underappreciated,

rigorous biomechanical validation of TBI models is essential to minimize variations in

the outcomes and increase their translational potential (Namjoshi et al., 2013a). As

described in the Discussion section of Chapter 2, we did not conduct a systematic

biomechanical evaluation of the weight drop model. This strongly motivated us to

develop a novel TBI model that can mitigate the caveats associated with the weight

drop model, as well as allow us to study injury biomechanics. Accordingly in

collaboration with the biomechanical engineers at the University of British Columbia, we

developed a novel neurotrauma model called CHIMERA that uses completely non-

surgical procedure to induce closed head injury in mice. In the Chapter 3 of this thesis, I

have discussed the biomechanical, functional as well as neuropathological

characterization of CHIMERA.

As will be discussed in the next section of this Chapter, we now have a novel, simple

yet highly versatile TBI model that can reliably mimic some neuropathological features

of human TBI with high experimental reproducibility. Based on the results discussed in

this thesis, we believe that CHIMERA will prove a highly useful and biofidelic model for

conducting future studies including those proposed above.

222

4.3 Chapter 3: Conclusions

The study described in Chapter 3 was not driven by a specific hypothesis but rather

describes characterization of a new, simple, and reliable model of murine CHI,

CHIMERA that replicates fundamental aspects of human impact TBI through precise

delivery of known biomechanical inputs.

CHIMERA is completely nonsurgical and requires only isoflurane anesthesia, therefore

enabling immediate neurological severity assessments using LRR and NSS measures.

Being nonsurgical, CHIMERA is ideal for studies investigating multiple impacts as well

the long-term consequences of impact TBI. These advantages overcome many

limitations of surgically-induced TBI models, including longer anesthetic exposure as

well as inclusion of opioid analgesics (e.g., buprenorphine) and sedatives like xylazine

that can interfere with rapid neurological assessment, low throughput, extensive

operator training, and limited suitability for studies involving repetitive TBI or long-term

TBI outcomes. CHIMERA produces injury using a simple, reliable and semi-automated

procedure that requires < 10 min per animal to produce defined injury. As the

biomechanical input parameters are highly adjustable, i.e., across impact energy,

velocity, and direction, CHIMERA offers a wide dynamic range of precisely controllable

inputs to reproduce the conditions that occur in human impact TBI. Importantly, the

kinematic analyses facilitated by CHIMERA enable head motion parameters to be

integrated with behavioral and neuropathological outcome measures, potentially

enabling improved relevance to human TBI. CHIMERA produces diffuse injury

223

characterized by activation of inflammatory reactions, axonal damage, and tau

phosphorylation, replicating many aspects of human impact TBI without overt focal

damage. Taken together, these attributes make CHIMERA a valuable new model to

investigate the mechanisms of TBI and for use in preclinical drug discovery and

development programs.

We found several differences in the neuropathological outcomes of CHIMERA-rTBI

compared to the weight drop rTBI. Thus, silver stain data revealed that the axonal injury

in CHIMERA-rTBI was different than that in weight drop rTBI in the terms of extent,

intensity and dynamics. In the weight drop rTBI, for example the axonal injury was

restricted in the coup and contrecoup white matter tracts below the impact site. In the

CHIMERA-rTBI in addition to the coup-contrecoup spread, we found significant axonal

pathology at the site distant from the impact site, such as olfactory nerve layer.

Qualitative comparison revealed robust axonal pathology in CHIMERA-rTBI compared

to mild axonal injury in weight drop rTBI. Lastly, while the axonal pathology in weight

drop rTBI resolved over 7-14 days, it continued to increase over the same duration in

CHIMERA-rTBI suggesting ongoing secondary injury pathways that may have

overwhelmed endogenous recovery mechanisms. CHIMERA-rTBI induced robust

microglial activation in the white matter tracts along with moderate increase in the

cytokine levels. On the other hand both microglial activation and elevation of cytokine

levels were absent in weight drop rTBI. There are several possibilities for these

differences. For example the mechanical inputs as well as head support system used in

these two models were different that may have affected head kinematics. For example,

224

while we did not systematically assessed head kinematics in the weight drop rTBI

model, a preliminary high-speed video analysis revealed almost no head moment but a

crush type of injury following weight drop (data not presented). On the other hand

CHIMERA-rTBI allows free head movement that translates into robust, widespread and

sustained axonal injury coupled with significant microglial activation and elevated

cytokine levels. Moreover, we cannot rule out influence of longer duration of anesthesia

as well as effects of ketamine, xylazine, and bupivacaine on the secondary injury

pathways in weight drop rTBI. Lastly, while CHIMERA-rTBI induced a transient increase

in phosphorylation of endogenous tau, tau phosphorylation in weight drop rTBI was

variable (data not shown).

Comparisons of weight drop rTBI and CHIMERA-rTBI models are summarized in the

Table 4.1

225

Weight Drop rTBI CHIMERA rTBI TBI Procedure: TBI Procedure: Surgery required Surgery-free Total duration of isoflurane anesthesia: ~ 20 min

Total duration of isoflurane anesthesia: ~ 4 min (without videography) and ~ 8 min (with videography)

Local anesthetic used: Bupivacaine No local anesthetic required Analgesics used: Meloxicam Analgesics used: Meloxicam Sedatives used: Ketamine and xylazine No sedatives required Total duration of injury procedure: ~ 20 min/animal

Total duration of injury procedure: 4-8 min/animal

Duration of recovery from anesthesia: 60-90 min

Duration of recovery from anesthesia: 6-10 min

Mortality rate: 10-15 % Mortality rate: 3-4 % Input Parameters and Head Kinematics: Input Parameters and Head Kinematics: Drop height is controlled, however friction and air resistance can have unpredictable influence on the velocity of gravity-driven mass

Input parameters are precisely controlled. Piston velocity is maintained throughout the piston travel by constant air pressure.

The head support system allows virtually no head movement following impact

The head support system allows free head movement following impact

Head requires to be held manually under the guide tube that can influence head kinematics following impact

Head does not require manual holding

High-speed videography requires manual trigger and cannot be easily integrated into the injury procedure

High-speed videography is triggered by CHIMERA impactor and is easily integrated into the injury procedure

Proposed Injury Mechanism Proposed Injury Mechanism Impact-head crush/compression Impact-head acceleration Behavioral Outcomes Behavioral Outcomes Injury induced cognitive deficits from 2-14 d as assessed by NOR

Injury induced cognitive deficits from 2-14 days as assessed by passive avoidance test and Barnes maze

Injury induced motor deficits from 1-7 d with spontaneous recovery by 14 d.

Injury induced motor deficits from 1-7 d with spontaneous recovery by 14 d.

Neuropathological Outcomes Neuropathological Outcomes Axonal injury restricted to coup-contrecoup white matter under the impact site

Widespread axonal injury including coup-contrecoup white matter as well as site distant from the impact (olfactory bulb)

Mild axonal injury Robust axonal injury Axonal injury in most of white matter resolves over 7-14 days

Axonal injury continues to increase over 2-14 days

No microglial activation Robust microglial activation Biochemical Outcomes Biochemical Outcomes No change in the proinflammatory cytokines Moderate and transient increase in the

proinflammatory cytokine levels Variable changes in phosphorylation of endogenous tau

Transient but significant increase in endogenous tau phosphorylation

Table 4.1: Comparison of weight drop-rTBI and CHIMERA-rTBI models.

226

4.4 Chapter 3: Study Limitations and Future Directions

While CHIMERA offers several advantages over traditional TBI models, we also

acknowledge several limitations of the present study, which stem principally from the

very nascent nature of the model. An interesting aspect of this study was although

several behavioral, inflammatory, and tau responses mostly returned to baseline by 14

days post-injury using the injury parameters used in this study, cognitive performance

and axonal pathology remained robust. Understanding the temporal response of the

various behavioral, biochemical and neuropathological outcomes used in this study

under a variety of injury conditions is now a very important future topic. Thus, the injury

characteristics using CHIMERA can be varied by independently adjusting various

parameters (e.g., impact energy, number of impacts, duration between successive

impacts, impact location and impact direction). Such assessment was beyond the scope

of this thesis.

While CHIMERA-TBI allows free head movement following TBI, high speed videos

revealed that the impact induced significant hyperflexion of the animal’s neck. It is

possible the frontal Velcro straps used to restrain the animal body may lead to

significant strain in the neck during such hyperflexion that can result in the spinal cord

injury. Indeed during the pilot study, post-mortem examination revealed that

approximately 30% of mice injured by CHIMERA develop visible contusions of the

cervical and thoracic spinal cord and these animals have significantly worse motor

behavior. Moreover the most common reason for the post-TBI mortality (~ 3.75%)

227

associated with CHIMERA was the development of hind limb paralysis. While seemingly

a caveat of the model, this can be exploited to develop a neurotrauma model that

induces either TBI alone or TBI along with spinal cord injury (SCI). Accordingly, future

studies are designed to control neck hyperflexion. One way to achieve this is by

gradually slowing head acceleration after peak acceleration phase (that is thought to

cause TBI) using soft foam. The utility of such model is further highlighted by recent

data indicating that TBI is a common co-occurring injury with SCI (Macciocchi et al.,

2008). Moreover, up to 30% of neurotrauma cases may also have some degree of

combined TBI/SCI, which can go undiagnosed and untreated relative to the most

obvious injury (Povolny and Kaplan, 1993). To our knowledge, there is only one animal

model of combined TBI/SCI that uses a double surgical method to induce separate TBI

and SCI (Inoue et al., 2013). By controlling neck flexion, we will be in a unique position

to study either TBI or combined TBI and SCI without any surgical manipulations in a

single model.

A major limitation of the present study is short post-TBI follow-up time. While inclusion

of longer post-injury time points was beyond the scope of this thesis, ongoing and future

studies are designed to address this. While thigmotaxis indicates only anxiety-like

behavior, more appropriate tests such as elevated plus maze will be used in future to

assess post-TBI anxiety.

Ongoing and future studies are also designed to undertake a detailed injury

characterization of CHIMERA to test influence of various input parameters such as full

228

dynamic range of CHIMERA, comparison of vertical vs lateral head impact and varying

the number or intervals of repetitive impacts. As mentioned previously, the wide impact

energy range and relatively simple design makes CHIMERA easily adaptable for

studying TBI in large rodents such as rats, which have been the traditional rodent model

of choice for most TBI studies due to their superior behavioral repertoire and larger

brains for imaging. Future studies are planned to compare injury responses in mice and

rats using CHIMERA.

Notwithstanding these limitations, the present study indicates that CHIMERA will be a

valuable neurotrauma model for conducting both the mechanistic as well as

pharmacological studies.

Taken together, the first study described in this thesis shows that the synthetic LXR

agonists may be further developed as a potential treatment for mild TBI. While this

study indicates that the beneficial effects of LXR in TBI are mediated partially through

apoE-dependent mechanisms there are several unanswered questions related to the

mechanism. We believe that these questions can be addressed in future studies using a

highly innovative neurotrauma model, CHIMERA described in the second study of this

thesis.

229

Bibliography

Abbott NJ, Ronnback L, Hansson E (2006) Astrocyte-endothelial interactions at the

blood-brain barrier. Nat Rev Neurosci 7:41-53.

Abildayeva K, Jansen PJ, Hirsch-Reinshagen V, Bloks VW, Bakker AHF, Ramaekers

FCS, de Vante J, Groen AK, Wellington CL, Kuipers F, Mulder M (2006) 24(S)-

hydroxycholesterol participates in a liver X recpetor-controlled pathway in astrocytes

that regulates apolipoprotein E-mediated cholesterol efflux. J Biol Chem 281:12799-

12808.

Adams JH (1980) Brain damage in fatal non-missile head injury in man. Amsterdam:

Elsevier Publishing company.

Adams JH, Doyle D, Graham DI, Lawrence AE, McLellan DR (1986) Gliding contusions

in nonmissile head injury in humans. Arch Pathol Lab Med 110:485-488.

Adams JH, Doyle D, Ford I, Gennarelli TA, Graham DI, McLellan DR (1989) Diffuse

axonal injury in head injury: definition, diagnosis and grading. Histopathology 15:49-59.

Adelson PD, Wisniewski SR, Beca J, Brown SD, Bell M, Muizelaar JP, Okada P, Beers

SR, Balasubramani GK, Hirtz D (2013) Comparison of hypothermia and normothermia

after severe traumatic brain injury in children (Cool Kids): a phase 3, randomised

controlled trial. Lancet Neurol 12:546-553.

230

Adeosun SO, Hou X, Zheng B, Stockmeier C, Ou X, Paul I, Mosley T, Weisgraber K,

Wang JM (2014) Cognitive deficits and disruption of neurogenesis in a mouse model of

apolipoprotein E4 domain interaction. J Biol Chem 289:2946-2959.

Adle-Biassette H, Duyckaerts C, Wasowicz M, He Y, Fornes P, Foncin JF, Lecomte D,

Hauw JJ (1996) Beta AP deposition and head trauma. Neurobiol Aging 17:415-419.

Aggerbeck LP, Wetterau JR, Weisgraber KH, Wu CS, Lindgren FT (1988) Human

apolipoprotein E3 in aqueous solution. II. Properties of the amino- and carboxyl-terminal

domains. J Biol Chem 263:6249-6258.

Akirav I, Maroun M (2006) Ventromedial prefrontal cortex is obligatory for consolidation

and reconsolidation of object recognition memory. Cereb Cortex 16:1759-1765.

Alder J, Fujioka W, Lifshitz J, Crockett DP, Thakker-Varia S (2011) Lateral fluid

percussion: model of traumatic brain injury in mice. J Vis Exp:e3063.

Alonso AdC, Li B, Grundke-Iqbal I, Iqbal K (2006) Polymerization of

hyperphosphorylated tau into filaments eliminates its inhibitory activity. Proc Natl Acad

Sci U S A 103:8864-8869.

Andersson S, Gustafsson N, Warner M, Gustafsson JA (2005) Inactivation of liver X

receptor beta leads to adult-onset motor neuron degeneration in male mice. Proc Natl

Acad Sci U S A 102:3857-3862.

231

Andoh T, Kuraishi Y (2003) Nitric oxide enhances substance P-induced itch-associated

responses in mice. Br J Pharmacol 138:202-208.

Andreadis A, Brown WM, Kosik KS (1992) Structure and novel exons of the human tau

gene. Biochemistry 31:10626-10633.

Antunes M, Biala G (2012) The novel object recognition memory: neurobiology, test

procedure, and its modifications. Cogn Process 13:93-110.

Apelt J, Ach K, Schliebs R (2003) Aging-related down-regulation of neprilysin, a putative

beta-amyloid-degrading enzyme, in transgenic Tg2576 Alzheimer-like mouse brain is

accompanied by an astroglial upregulation in the vicinity of beta-amyloid plaques.

Neurosci Lett 339:183-186.

Armstrong RA (1994) beta-Amyloid (A beta) deposition in elderly non-demented

patients and patients with Alzheimer's disease. Neurosci Lett 178:59-62.

Auboeuf D, Rieusset J, Fajas L, Vallier P, Frering V, Riou JP, Staels B, Auwerx J,

Laville M, Vidal H (1997) Tissue distribution and quantification of the expression of

mRNAs of peroxisome proliferator-activated receptors and liver X receptor-alpha in

humans: no alteration in adipose tissue of obese and NIDDM patients. Diabetes

46:1319-1327.

232

Avila J, Lucas JJ, Perez M, Hernandez F (2004) Role of tau protein in both physiological

and pathological conditions. Physiol Rev 84:361-384.

Ballatore C, Lee VM, Trojanowski JQ (2007) Tau-mediated neurodegeneration in

Alzheimer's disease and related disorders. Nat Rev Neurosci 8:663-672.

Banos JH, Novack TA, Brunner R, Renfroe S, Lin HY, Meythaler J (2010) Impact of

early administration of sertraline on cognitive and behavioral recovery in the first year

after moderate to severe traumatic brain injury. J Head Trauma Rehabil 25:357-361.

Barnes DE, Kaup A, Kirby KA, Byers AL, Diaz-Arrastia R, Yaffe K (2014) Traumatic

brain injury and risk of dementia in older veterans. Neurology 83:312-319.

Basak JM, Verghese PB, Yoon H, Kim J, Holtzman DM (2012) Low-density lipoprotein

receptor represents an apolipoprotein E-independent pathway of Abeta uptake and

degradation by astrocytes. J Biol Chem 287:13959-13971.

Bateman RJ, Munsell LY, Morris JC, Swarm R, Yarasheski KE, Holtzman DM (2006)

Human amyloid-beta synthesis and clearance rates as measured in cerebrospinal fluid

in vivo. Nat Med 12:856-861.

Baugh CM, Robbins CA, Stern RA, McKee AC (2014) Current understanding of chronic

traumatic encephalopathy. Curr Treat Options Neurol 16:306.

233

Bayly PV, Cohen TS, Leister EP, Ajo D, Leuthardt EC, Genin GM (2005) Deformation of

the human brain induced by mild acceleration. J Neurotrauma 22:845-856.

Bazarian JJ, Blyth BJ, He H, Mookerjee S, Jones C, Kiechle K, Moynihan R, Wojcik SM,

Grant WD, Secreti LM, Triner W, Moscati R, Leinhart A, Ellis GL, Khan J (2013)

Classification accuracy of serum Apo A-I and S100B for the diagnosis of mild traumatic

brain injury and prediction of abnormal initial head computed tomography scan. J

Neurotrauma 30:1747-1754.

Bell RD, Sagare AP, Friedman AE, Bedi GS, Holtzman DM, Deane R, Zlokovic BV

(2007) Transport pathways for clearance of human Alzheimer's amyloid beta-peptide

and apolipoproteins E and J in the mouse central nervous system. J Cereb Blood Flow

Metab 27:909-918.

Bellosta S, Nathan BP, Orth M, Dong LM, Mahley RW, Pitas RE (1995) Stable

expression and secretion of apolipoproteins E3 and E4 in mouse neuroblastoma cells

produces differential effects on neurite outgrowth. J Biol Chem 270:27063-27071.

Beni-Adani L, Gozes I, Cohen Y, Assaf Y, Steingart RA, Brenneman DE, Eizenberg O,

Trembolver V, Shohami E (2001) A peptide derived from activity-dependent

neuroprotective protein (ADNP) ameliorates injury response in closed head injury in

mice. J Pharmacol Exp Ther 296:57-63.

234

Bernard SA, Nguyen V, Cameron P, Masci K, Fitzgerald M, Cooper DJ, Walker T, Std

BP, Myles P, Murray L, David, Taylor, Smith K, Patrick I, Edington J, Bacon A,

Rosenfeld JV, Judson R (2010) Prehospital rapid sequence intubation improves

functional outcome for patients with severe traumatic brain injury: a randomized

controlled trial. Ann Surg 252:959-965.

Beschorner R, Nguyen TD, Gozalan F, Pedal I, Mattern R, Schluesener HJ, Meyermann

R, Schwab JM (2002) CD14 expression by activated parenchymal

microglia/macrophages and infiltrating monocytes following human traumatic brain

injury. Acta Neuropathol 103:541-549.

Biernat J, Gustke N, Drewes G, Mandelkow E, Mandelkow E (1993) Phosphorylation of

Ser262 strongly reduces binding of tau to microtubules: Distinction between PHF-like

immunoreactivity and microtubule binding. Neuron 11:153-163.

Biffi A, Greenberg SM (2011) Cerebral amyloid angiopathy: a systematic review. J Clin

Neurol 7:1-9.

Binder LI, Frankfurter A, Rebhun LI (1985) The distribution of tau in the mammalian

central nervous system. J Cell Biol 101:1371-1378.

Bjorkhem I, Meaney S (2004) Brain cholesterol: long secret life behind a barrier.

Arterioscler Thromb Vasc Biol 24:806-815.

235

Blumbergs PC (1997) Pathology. In: Head Injury - Pathophysiology and Management of

Severe Closed Head Injury (Reilly P, Bullock R, eds), pp 39-70. London: Chapman and

Hall.

Blume HK, Lucas S, Bell KR (2011) Subacute concussion-related symptoms in youth.

Phys Med Rehabil Clin N Am 22:665-681, viii-ix.

Boche D, Perry VH, Nicoll JA (2013) Review: activation patterns of microglia and their

identification in the human brain. Neuropathol Appl Neurobiol 39:3-18.

Bodzioch M, Ors¢ E, Klucken J, Langmann T, B”ttcher A, Diederich W, Drobnik W,

Barlage S, B �chler C, Porsch-™zc �r�mez M, Kam

Rothe G, Aslanidis C, Lackner KJ, Schmitz G (1999) The gene encoding ATP-binding

cassette transporter 1 is mutated in Tangier Disease. Nat Genet 22:347-351.

Bouma GJ, Muizelaar JP, Bandoh K, Marmarou A (1992) Blood pressure and

intracranial pressure-volume dynamics in severe head injury: relationship with cerebral

blood flow. J Neurosurg 77:15-19.

Bouzat P, Sala N, Payen JF, Oddo M (2013) Beyond intracranial pressure: optimization

of cerebral blood flow, oxygen, and substrate delivery after traumatic brain injury. Ann

Intensive Care 3:23.

236

Brain Injury Society of Toronto (2014) Brain injury facts and figure.

http://www.bist.ca/brain-injury-fact-figures. Accessed on May 22, 2014.

Brandenburg W, Hallervorden J (1954) Dementia pugilistica with anatomical findings.

Virchows Arch 325:680-709.

Bré MH, Karsenti E (1990) Effects of brain microtubule-associated proteins on

microtubule dynamics and the nucleating activity of centrosomes. Cell Motil

Cytoskeleton 15:88-98.

Brecht WJ, Harris FM, Chang S, Tesseur I, Yu GQ, Xu Q, Dee Fish J, Wyss-Coray T,

Buttini M, Mucke L, Mahley RW, Huang Y (2004) Neuron-specific apolipoprotein e4

proteolysis is associated with increased tau phosphorylation in brains of transgenic

mice. J Neurosci 24:2527-2534.

Broadbent NJ, Gaskin S, Squire LR, Clark RE (2010) Object recognition memory and

the rodent hippocampus. Learn Mem 17:5-11.

Brody DL, Magnoni S, Schwetey KE, Spinner ML, Esparza TJ, Stocchetti N, Zipfel GJ,

Holtzman DM (2008) Amyloid-β dynamics correlate with neurological status in the

injured human brain. Science 321:1221-1224.

Brody DL, Mac Donald C, Kessens CC, Yuede C, Parsadanian M, Spinner M, Kim E,

Schwetye KE, Holtzman DM, Bayly PV (2007) Electromagnetic controlled cortical

237

impact device for precise, graded experimental traumatic brain injury. J Neurotrauma

24:657-673.

Brooks-Wilson A et al. (1999) Mutations in ABC1 in Tangier disease and familial high-

density lipoprotein deficiency. Nat Genet 22:336-345.

Browne KD, Chen XH, Meaney DF, Smith DH (2011) Mild traumatic brain injury and

diffuse axonal injury in swine. J Neurotrauma 28:1747-1755.

Brunham LR, Kruit JK, Pape TD, Parks JS, Kuipers F, Hayden MR (2006) Tissue-

specific induction of intestinal ABCA1 expression with a liver X receptor agonist raises

plasma HDL cholesterol levels. Circ Res 99:672-674.

Bu G (2009) Apolipoprotein E and its receptors in Alzheimer's disease: pathways,

pathogenesis and therapy. Nat Rev Neurosci 10:333-344.

Buki A, Povlishock JT (2006) All roads lead to disconnection?--Traumatic axonal injury

revisited. Acta Neurochir (Wien) 148:181-193; discussion 193-184.

Buki A, Koizumi H, Povlishock JT (1999a) Moderate posttraumatic hypothermia

decreases early calpain-mediated proteolysis and concomitant cytoskeletal compromise

in traumatic axonal injury. Exp Neurol 159:319-328.

238

Buki A, Siman R, Trojanowski JQ, Povlishock JT (1999b) The role of calpain-mediated

spectrin proteolysis in traumatically induced axonal injury. J Neuropathol Exp Neurol

58:365-375.

Bullock MR, Chesnut R, Ghajar J, Gordon D, Hartl R, Newell DW, Servadei F, Walters

BC, Wilberger JE (2006) Surgical management of acute epidural hematomas.

Neurosurgery 58:S7-15; discussion Si-iv.

Bullock R (1994) Excitatory amino acids following brain injury. J Neurosurg 80:595-596.

Bullock R, Zauner A, Woodward JJ, Myseros J, Choi SC, Ward JD, Marmarou A, Young

HF (1998) Factors affecting excitatory amino acid release following severe human head

injury. J Neurosurg 89:507-518.

Burns MP, Noble WJ, Olm V, Gaynor K, Casey E, LaFrancois J, Wang L, Duff K (2003)

Co-localization of cholesterol, apolipoprotein E and fibrillar Abeta in amyloid plaques.

Brain Res Mol Brain Res 110:119-125.

Buxbaum JD, Thinakaran G, Koliatsos V, O'Callahan J, Slunt HH, Price DL, Sisodia SS

(1998) Alzheimer amyloid protein precursor in the rat hippocampus: transport and

processing through the perforant path. J Neurosci 18:9629-9637.

Calatayud Maldonado V, Calatayud Perez JB, Aso Escario J (1991) Effects of CDP-

choline on the recovery of patients with head injury. J Neurol Sci 103 Suppl:S15-18.

239

Canadian Institute for Health Information (2006) Head Injuries in Canada: A Decade of

Change (1994-1995 to 2003-2004). In: Analysis in Brief: Canadian Institute for Health

Information.

Cantu RC (2007) Chronic traumatic encephalopathy in the National Football League.

Neurosurgery 61:223-225.

Carpentier M, Robitaille Y, DesGroseillers L, Boileau G, Marcinkiewicz M (2002)

Declining expression of neprilysin in Alzheimer disease vasculature: possible

involvement in cerebral amyloid angiopathy. J Neuropathol Exp Neurol 61:849-856.

Carroll LJ, Cassidy JD, Holm L, Kraus J, Coronado VG (2004) Methodological issues

and research recommendations for mild traumatic brain injury: the WHO Collaborating

Centre Task Force on Mild Traumatic Brain Injury. J Rehabil Med:113-125.

Cartagena CM, Ahmed F, Burns MP, Pajoohesh-Ganji A, Pak DT, Faden AI, Rebeck

GW (2008) Cortical injury increases cholesterol 24S hydroxylase (Cyp46) levels in the

rat brain. J Neurotrauma 25:1087-1098.

Cassidy JD, Carroll LJ, Peloso PM, Borg J, von Holst H, Holm L, Kraus J, Coronado VG

(2004) Incidence, risk factors and prevention of mild traumatic brain injury: results of the

WHO Collaborating Centre Task Force on Mild Traumatic Brain Injury. J Rehabil

Med:28-60.

240

Castellano JM, Kim J, Stewart FR, Jiang H, DeMattos RB, Patterson BW, Fagan AM,

Morris JC, Mawuenyega KG, Cruchaga C, Goate AM, Bales KR, Paul SM, Bateman RJ,

Holtzman DM (2011) Human apoE isoforms differentially regulate brain amyloid-beta

peptide clearance. Sci Transl Med 3:89ra57.

Cavelier C, Lorenzi I, Rohrer L, von Eckardstein A (2006) Lipid efflux by the ATP-

binding cassette transporters ABCA1 and ABCG1. Biochim Biophys Acta 1761:655-666.

Centers for Disease Control and Prevention (2007) Nonfatal traumatic brain injuries

from sports and recreation activities--United States, 2001-2009. MMWR Morb Mortal

Wkly Rep 56:733-737.

Cernak I (2005) Animal models of head trauma. NeuroRx 2:410-422.

Chartier-Harlin M-C, Crawford F, Houlden H, Warren A, Hughes D, Fidani L, Goate A,

Rossor M, Roques P, Hardy J, Mullan M (1991) Early-onset Alzheimer's disease caused

by mutations at codon 717 of the [beta]-amyloid precursor protein gene. Nature

353:844-846.

Chawla A, Boisvert WA, Lee CH, Laffitte BA, Barak Y, Joseph SB, Liao D, Nagy P,

Edwards PA, Curtiss LK, Evans RM, Tontonoz P (2001) A PPAR g -LXR-ABCA1

pathway in macrophages is involved in cholesterol efflux and atherogenesis. Molec Cell

7:161-171.

241

Chen JD, Evans RM (1995) A transcriptional co-repressor that interacts with nuclear

hormone receptors. Nature 377:454-457.

Chen S, Averett NT, Manelli A, Ladu MJ, May W, Ard MD (2005) Isoform-specific

effects of apolipoprotein E on secretion of inflammatory mediators in adult rat microglia.

J Alzheimers Dis 7:25-35.

Chen W, Sun Y, Welch C, Gorelik A, Leventhal AR, Tabas I, Tall AR (2001) Preferential

ATP-binding cassette transporter A1-mediated cholesterol efflux from late

endosomes/lysosomes. J Biol Chem 276:43564-43569.

Chen XH, Johnson VE, Uryu K, Trojanowski JQ, Smith DH (2009) A lack of amyloid á

plaques despite persistent accumulation of amyloid á in axons of long-term survivors of

traumatic brain injury. Brain Pathol 19:214-223.

Chen XH, Siman R, Iwata A, Meaney DF, Trojanowski JQ, Smith DH (2004) Long-term

accumulation of amyloid-beta, beta-secretase, presenilin-1, and caspase-3 in damaged

axons following brain trauma. Am J Pathol 165:357-371.

Chen Y, Lomnitski L, Michaelson DM, Shohami E (1997) Motor and cognitive deficits in

apolipoprotein E-deficient mice after closed head injury. Neuroscience 80:1255-1262.

242

Chen Y, Durakoglugil MS, Xian X, Herz J (2010) ApoE4 reduces glutamate receptor

function and synaptic plasticity by selectively impairing ApoE receptor recycling. Proc

Natl Acad Sci U S A 107:12011-12016.

Chen Y, Constantini S, Trembovler V, Weinstock M, Shohami E (1996) An experimental

model of closed head injury in mice: Pathophysiology, histopathology, and cognitive

deficits. J Neurotrauma 13:557-568.

Chesnut RM, Temkin N, Carney N, Dikmen S, Rondina C, Videtta W, Petroni G, Lujan

S, Pridgeon J, Barber J, Machamer J, Chaddock K, Celix JM, Cherner M, Hendrix T

(2012) A Trial of Intracranial-Pressure Monitoring in Traumatic Brain Injury. New Eng J

Med 367:2471-2481.

Chiang JY, Kimmel R, Stroup D (2001) Regulation of cholesterol 7alpha-hydroxylase

gene (CYP7A1) transcription by the liver orphan receptor (LXRalpha). Gene 262:257-

265.

Chin LS, Toshkezi G, Cantu RC (2011) Traumatic encephalopathy related to sports

injury. US Neurology 7:33-36.

Chobanian AV, Hollander W (1962) Body cholesterol metabolism in man. I. The

equilibration of serum and tissue cholesterol. J Clin Invest 41:1732-1737.

243

Chodobski A, Zink BJ, Szmydynger-Chodobska J (2011) Blood-brain barrier

pathophysiology in traumatic brain injury. Transl Stroke Res 2:492-516.

Ciallella JR, Ikonomovic MD, Paljug WR, Wilbur YI, Dixon CE, Kochanek PM, Marion

DW, DeKosky ST (2002) Changes in expression of amyloid precursor protein and

interleukin-1beta after experimental traumatic brain injury in rats. J Neurotrauma

19:1555-1567.

Cirrito JR, Yamada KA, Finn MB, Sloviter RS, Bales KR, May PC, Schoepp DD, Paul

SM, Mennerick S, Holtzman DM (2005) Synaptic activity regulates interstitial fluid

amyloid-beta levels in vivo. Neuron 48:913-922.

Citron M, Oltersdorf T, Haass C, McConlogue L, Hung AY, Seubert P, Vigo-Pelfrey C,

Lieberburg I, Selkoe DJ (1992) Mutation of the beta-amyloid precursor protein in familial

Alzheimer's disease increases beta-protein production. Nature 360:672-674.

Clausen F, Lorant T, Lewen A, Hillered L (2007) T lymphocyte trafficking: a novel target

for neuroprotection in traumatic brain injury. J Neurotrauma 24:1295-1307.

Clifton GL, Miller ER, Choi SC, Levin HS, McCauley S, Smith KR, Jr., Muizelaar JP,

Marion DW, Luerssen TG (2002) Hypothermia on admission in patients with severe

brain injury. J Neurotrauma 19:293-301.

244

Clifton GL, Valadka A, Zygun D, Coffey CS, Drever P, Fourwinds S, Janis LS, Wilde E,

Taylor P, Harshman K, Conley A, Puccio A, Levin HS, McCauley SR, Bucholz RD,

Smith KR, Schmidt JH, Scott JN, Yonas H, Okonkwo DO (2011) Very early hypothermia

induction in patients with severe brain injury (the National Acute Brain Injury Study:

Hypothermia II): a randomised trial. Lancet Neurol 10:131-139.

Collins JL, Fivush AM, Watson MA, Galardi CM, Lewis MC, Moore LB, Parks DJ, Wilson

JG, Tippin TK, Binz JG, Plunket KD, Morgan DG, Beaudet EJ, Whitney KD, Kliewer SA,

Willson TM (2002) Identification of a nonsteroidal liver X receptor agonist through

parallel array synthesis of tertiary amines. J Med Chem 45:1963-1966.

Colton CA (2009) Heterogeneity of microglial activation in the innate immune response

in the brain. J Neuroimmune Pharmacol 4:399-418.

Conde C, Caceres A (2009) Microtubule assembly, organization and dynamics in axons

and dendrites. Nat Rev Neurosci 10:319-332.

Cook DG, Leverenz JB, McMillan PJ, Kulstad JJ, Ericksen S, Roth RA, Schellenberg

GD, Jin LW, Kovacina KS, Craft S (2003) Reduced hippocampal insulin-degrading

enzyme in late-onset Alzheimer's disease is associated with the apolipoprotein E-

epsilon4 allele. Am J Pathol 162:313-319.

Cooper DJ, Myles PS, McDermott FT, Murray LJ, Laidlaw J, Cooper G, Tremayne AB,

Bernard SS, Ponsford J (2004) Prehospital hypertonic saline resuscitation of patients

245

with hypotension and severe traumatic brain injury: a randomized controlled trial. J Am

Med Assoc 291:1350-1357.

Cooper DJ, Rosenfeld JV, Murray L, Arabi YM, Davies AR, D'Urso P, Kossmann T,

Ponsford J, Seppelt I, Reilly P, Wolfe R (2011) Decompressive craniectomy in diffuse

traumatic brain injury. N Engl J Med 364:1493-1502.

Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW,

Roses AD, Haines JL, Pericak-Vance MA (1993) Gene dose of apolipoprotein E type 4

allele and the risk of Alzheimer's disease in late onset families. Science 261:921-923.

Corder EH, Saunders AM, Risch NJ, Strittmatter WJ, Schmechel DE, Gaskell PC,

Rimmler JB, Locke PA, Conneally PM, Schmader KE, Small GW, Roses AD, Haines JL,

Pericak-Vance MA (1994) Protective effect of apolipoprotein E type 2 for late onset

Alzheimer disease. Nat Genet 7:180-184.

Corsellis JA, Bruton CJ, Freeman-Browne D (1973) The aftermath of boxing. Psychol

Med 3:270-303.

Costet P, Luo Y, Wang N, Tall AR (2000) Sterol-dependent transactivation of the ABC1

promoter by the liver X receptor/retinoid X receptor. J Biol Chem 275:28240-28245.

246

Creeley CE, Wozniak DF, Bayly PV, Olney JW, Lewis LM (2004) Multiple episodes of

mild traumatic brain injury result in impaired cognitive performance in mice. Acad Emerg

Med 11:809-819.

Cronican AA, Fitz NF, Pham T, Fogg A, Kifer B, Koldamova R, Lefterov I (2010) Proton

pump inhibitor lansoprazole is a nuclear liver X receptor agonist. Biochem Pharmacol

79:1310-1316.

Cruz J, Minoja G, Okuchi K (2001) Improving clinical outcomes from acute subdural

hematomas with the emergency preoperative administration of high doses of mannitol:

a randomized trial. Neurosurgery 49:864-871.

Cruz J, Minoja G, Okuchi K (2002) Major clinical and physiological benefits of early high

doses of mannitol for intraparenchymal temporal lobe hemorrhages with abnormal

pupillary widening: a randomized trial. Neurosurgery 51:628-637; discussion 637-628.

Csuka E, Morganti-Kossmann MC, Lenzlinger PM, Joller H, Trentz O, Kossmann T

(1999) IL-10 levels in cerebrospinal fluid and serum of patients with severe traumatic

brain injury: relationship to IL-6, TNF-α, TGF-β1 and blood–brain barrier function. J

Neuroimmunol 101:211-221.

Csuka E, Hans VH, Ammann E, Trentz O, Kossmann T, Morganti-Kossmann MC (2000)

Cell activation and inflammatory response following traumatic axonal injury in the rat.

Neuroreport 11:2587-2590.

247

Daneshvar DH, Riley DO, Nowinski CJ, McKee AC, Stern RA, Cantu RC (2011) Long-

term consequences: effects on normal development profile after concussion. Phys Med

Rehabil Clin N Am 22:683-700, ix.

Darvish KK, Crandall JR (2001) Nonlinear viscoelastic effects in oscillatory shear

deformation of brain tissue. Med Eng Phys 23:633-645.

Davidsson J, Risling M (2011) A new model to produce sagittal plane rotational induced

diffuse axonal injuries. Front Neurol 2:41.

Davis AE (2000) Mechanisms of traumatic brain injury: biomechanical, structural and

cellular considerations. Crit Care Nurs Q 23:1-13.

Dawson SL, Hirsch CS, Lucas FV, Sebek BA (1980) The contrecoup phenomenon:

Reappraisal of a classic problem. Hum Pathol 11:155-166.

De Bonis P, Pompucci A, Mangiola A, D'Alessandris QG, Rigante L, Anile C (2010)

Decompressive craniectomy for the treatment of traumatic brain injury: does an age limit

exist? J Neurosurg 112:1150-1153.

de Bont N, Netea MG, Demacker PNM, Verschueren I, Kullberg BJ, van Dijk KW, van

der Meer JWM, Stalenhoef AFH (1999) Apolipoprotein E knock-out mice are highly

susceptible to endotoxemia and Klebsiella pneumoniae infection. J Lipid Res 40:680-

685.

248

Deacon RMJ (2013) Measuring Motor Coordination in Mice. J Vis Exp:e2609.

Deane R, Wu Z, Zlokovic BV (2004a) RAGE (yin) versus LRP (yang) balance regulates

alzheimer amyloid beta-peptide clearance through transport across the blood-brain

barrier. Stroke 35:2628-2631.

Deane R, Sagare A, Hamm K, Parisi M, Lane S, Finn MB, Holtzman DM, Zlokovic BV

(2008) apoE isoform-specific disruption of amyloid beta peptide clearance from mouse

brain. J Clin Invest 118:4002-4013.

Deane R, Wu Z, Sagare A, Davis J, Du Yan S, Hamm K, Xu F, Parisi M, LaRue B, Hu

HW, Spijkers P, Guo H, Song X, Lenting PJ, Van Nostrand WE, Zlokovic BV (2004b)

LRP/Amyloid β-Peptide Interaction Mediates Differential Brain Efflux of Aβ Isoforms.

Neuron 43:333-344.

Deane R et al. (2003) RAGE mediates amyloid-beta peptide transport across the blood-

brain barrier and accumulation in brain. Nat Med 9:907-913.

Decuypere M, Klimo P, Jr. (2012) Spectrum of traumatic brain injury from mild to

severe. Surg Clin North Am 92:939-957, ix.

DeFord SM, Wilson MS, Rice AC, Clausen T, Rice LK, Barabnova A, Hamm RJ (2002)

Repeated mild brain injuries result in cognitive impairment in B6C3F1 mice. J

Neurotrauma 19:427-438.

249

DeKosky ST, Abrahamson EE, Ciallella JR, Paljug WR, Wisniewski SR, Clark RS,

Ikonomovic MD (2007) Association of increased cortical soluble abeta42 levels with

diffuse plaques after severe brain injury in humans. Arch Neurol 64:541-544.

Delacourte A, Robitaille Y, Sergeant N, Buee L, Hof PR, Wattez A, Laroche-Cholette A,

Mathieu J, Chagnon P, Gauvreau D (1996) Specific pathological Tau protein variants

characterize Pick's disease. J Neuropathol Exp Neurol 55:159-168.

DeMattos RB, Brendza RP, Heuser JE, Kierson M, Cirrito JR, Fryer J, Sullivan PM,

Fagan AM, Han X, Holtzman DA (2001) Purification and characterization of astrocyte-

secreted apolipoprotein E and J-containing lipoproteins from wild-type and human apoE

transgenic mice. Neurochem Int 39:415-425.

Denny-Brown D, Russell WR (1940) Experimental cerebral concussion. J Physiol

99:153.

Dewitt DS, Perez-Polo R, Hulsebosch CE, Dash PK, Robertson CS (2013) Challenges

in the development of rodent models of mild traumatic brain injury. J Neurotrauma

30:688-701.

Dietschy JM, Turley SD (2001) Cholesterol metabolism in the brain. Curr Opin Lipidol

12:105-112.

250

Dietschy JM, Turley SD (2004) Cholesterol metabolism in the central nervous system

during early development and in the mature animal. J Lipid Res 45:1375-1397.

Dixon CE, Lighthall JW, Anderson TE (1988) Physiologic, histopathologic, and

cineradiographic characterization of a new fluid-percussion model of experimental brain

injury in the rat. J Neurotrauma 5:91-104.

Dixon CE, Clifton GL, Lighthall JW, Yaghmai AA, Hayes RL (1991) A controlled cortical

impact model of traumatic brain injury in the rat. J Neurosci Methods 39:253-262.

Dong LM, Weisgraber KH (1996) Human apolipoprotein E4 domain interaction. Arginine

61 and glutamic acid 255 interact to direct the preference for very low density

lipoproteins. J Biol Chem 271:19053-19057.

Dong LM, Wilson C, Wardell MR, Simmons T, Mahley RW, Weisgraber KH, Agard DA

(1994) Human apolipoprotein E. Role of arginine 61 in mediating the lipoprotein

preferences of the E3 and E4 isoforms. J Biol Chem 269:22358-22365.

Donkin JJ, Stukas S, Hirsch-Reinshagen V, Namjoshi D, Wilkinson A, May S, Chan J,

Fan J, Collins J, Wellington CL (2010) ATP-binding cassette transporter A1 mediates

the beneficial effects of the liver-X-receptor agonist GW3965 on object recognition

memory and amyloid burden in APP/PS1 mice. J Biol Chem 285:34144-34154.

251

Donnelly BR, Medige J (1997) Shear properties of human brain tissue. J Biomech Eng

119:423-432.

Drew LB, Drew WE (2004) The contrecoup-coup phenomenon: a new understanding of

the mechanism of closed head injury. Neurocrit Care 1:385-390.

Duhaime AC (2006) Large animal models of traumatic injury to the immature brain. Dev

Neurosci 28:380-387.

Egensperger R, Kosel S, von Eitzen U, Graeber MB (1998) Microglial activation in

Alzheimer disease: Association with APOE genotype. Brain Pathol 8:439-447.

Eiband AM (1959) Human tolerance to rapidly applied accelerations: a summary of the

literature. In: NASA Memorandum. Cleveland, Ohio: NASA Lewis Research Center.

Eisenberg HM, Gary HE, Jr., Aldrich EF, Saydjari C, Turner B, Foulkes MA, Jane JA,

Marmarou A, Marshall LF, Young HF (1990) Initial CT findings in 753 patients with

severe head injury. A report from the NIH Traumatic Coma Data Bank. J Neurosurg

73:688-698.

Elder GA, Cristian A (2009) Blast-related mild traumatic brain injury: mechanisms of

injury and impact on clinical care. Mount Sinai J Med 76:111-118.

252

Engel S, Schluesener H, Mittelbronn M, Seid K, Adjodah D, Wehner HD, Meyermann R

(2000) Dynamics of microglial activation after human traumatic brain injury are revealed

by delayed expression of macrophage-related proteins MRP8 and MRP14. Acta

Neuropathol (Berl) 100:313-322.

Engelhardt G (1996) Pharmacology of meloxicam, a new non-steroidal anti-

inflammatory drug with an improved safety profile through preferential inhibition of COX-

2. Br J Rheumatol 35 Suppl 1:4-12.

Ennaceur A (2010) One-trial object recognition in rats and mice: methodological and

theoretical issues. Behav Brain Res 215:244-254.

Ennaceur A, Delacour J (1988) A new one-trial test for neurobiological studies of

memory in rats. 1: Behavioral data. Behav Brain Res 31:47-59.

Eppinger R, Sun E, Kuppa S, Saul R (2000) Suppliment: Development of Improved

Injury Criteria for the Assessment of Advanced Automotive Restraint Systems - II. In, pp

1-40: National Highway Traffic Safety Administration,

http://www.nhtsa.gov/DOT/NHTSA/NRD/Multimedia/PDFs/Crashworthiness/Air%20Bag

s/finalrule_all.pdf, Accessed on March 2012.

Evans FG, Lissner HR, Lebow M (1958) The relation of energy, velocity, and

acceleration to skull deformation and fracture. Surg Gynecol Obstet 107:593-601.

253

Fagan AM, Bu G, Sun Y, Daugherty A, Holtzman DM (1996) Apolipoprotein E-

containing high density lipoprotein promotes neurite outgrowth and is a ligand for the

low density lipoprotein receptor-related protein. J Biol Chem 271:30121-30125.

Fagan AM, Holtzman DM, Munson G, Mathur T, Schnieder D, Chang LK, Getz GS,

Reardon CA, Lukens J, Shah JA, Ladu MJ (1999) Unique lipoproteins secreted by

primary astrocytes from wild type, apoE (-/-), and human apoE transgenic mice. J Biol

Chem 274:30001-30004.

Fan L, Young PR, Barone FC, Feuerstein GZ, Smith DH, McIntosh TK (1995)

Experimental brain injury induces expression of interleukin-1 beta mRNA in the rat

brain. Brain Res Mol Brain Res 30:125-130.

Fan QW, Iosbe I, Asou H, Yanagisawa K, Michikawa M (2001) Expression and

regulation of apolipoprotein E receptors in the cells of the central nervous system in

culture: A review. J Am Aging Assoc 24:1-10.

Farris W, Mansourian S, Chang Y, Lindsley L, Eckman EA, Frosch MP, Eckman CB,

Tanzi RE, Selkoe DJ, Guenette S (2003) Insulin-degrading enzyme regulates the levels

of insulin, amyloid beta-protein, and the beta-amyloid precursor protein intracellular

domain in vivo. Proc Natl Acad Sci U S A 100:4162-4167.

Faul MM, Xu L, Wald MM, Coronado VG (2010) Traumatic brain injury in the united

states: emergency department visits, hospitalizations and death 2002-2006. In. Atlanta,

254

GA: Centers for Disease Control and Prevention, National Center for Injury Prevention

and Control.

Feeney DM, Boyeson MG, Linn RT, Murray HM, Dail WG (1981) Responses to cortical

injury: I. Methodology and local effects of contusions in the rat. Brain Res 211:67-77.

Feigin VL, Theadom A, Barker-Collo S, Starkey NJ, McPherson K, Kahan M, Dowell A,

Brown P, Parag V, Kydd R, Jones K, Jones A, Ameratunga S (2013) Incidence of

traumatic brain injury in New Zealand: a population-based study. Lancet Neurol 12:53-

64.

Feng B, Tabas I (2002) ABCA1-mediated cholesterol efflux is defective in free

cholesterol-loaded macrophages. J Biol Chem 277:43271-43280.

Fijalkowski RJ, Stemper BD, Pintar FA, Yoganandan N, Crowe MJ, Gennarelli TA

(2007) New rat model for diffuse brain injury using coronal plane angular acceleration. J

Neurotrauma 24:1387-1398.

Finnie J (2001) Animal models of traumatic brain injury: a review. Aust Vet J 79:628-

633.

Fitz NF, Cronican AA, Saleem M, Fauq AH, Chapman R, Lefterov I, Koldamova R

(2012) Abca1 deficiency affects Alzheimer's disease-like phenotype in human ApoE4

but not in ApoE3-targeted replacement mice. J Neurosci 32:13125-13136.

255

Fitz NF, Cronican A, Pham T, Fogg A, Fauq AH, Chapman R, Lefterov I, Koldamova R

(2010) Liver X receptor agonist treatment ameliorates amyloid pathology and memory

deficits caused by high-fat diet in APP23 mice. J Neurosci 30:6862-6872.

Flament S, Delacourte A, Verny M, Hauw JJ, Javoy-Agid F (1991) Abnormal Tau

proteins in progressive supranuclear palsy. Similarities and differences with the

neurofibrillary degeneration of the Alzheimer type. Acta Neuropathol 81:591-596.

Fleminger S, Ponsford J (2005) Long term outcome after traumatic brain injury. Br Med

J 331:1419-1420.

Fleminger S, Oliver DL, Lovestone S, Rabe-Hesketh S, Giora A (2003) Head injury as a

risk factor for Alzheimer's disease: the evidence 10 years on; a partial replication. J

Neurol Neurosurg Psychiatry 74:857-862.

Flierl MA, Stahel PF, Beauchamp KM, Morgan SJ, Smith WR, Shohami E (2009) Mouse

closed head injury model induced by a weight-drop device. Nat Protoc 4:1328-1337.

Foda MA, Marmarou A (1994) A new model of diffuse brain injury in rats. Part II:

Morphological characterization. J Neurosurg 80:301-313.

Fork M, Bartels C, Ebert AD, Grubich C, Synowitz H, Wallesch CW (2005)

Neuropsychological sequelae of diffuse traumatic brain injury. Brain Inj 19:101-108.

256

Forman BM, Ruan B, Chen J, Schroepfer GJ, Jr., Evans RM (1997) The orphan nuclear

receptor LXRalpha is positively and negatively regulated by distinct products of

mevalonate metabolism. Proc Natl Acad Sci U S A 94:10588-10593.

Franz G, Beer R, Kampfl A, Engelhardt K, Schmutzhard E, Ulmer H, Deisenhammer F

(2003) Amyloid beta 1-42 and tau in cerebrospinal fluid after severe traumatic brain

injury. Neurology 60:1457-1461.

Franze K, Reichenbach A, Käs J (2009) Biomechanics of the CNS. In:

Mechanosensitivity of the Nervous System (Kamkim A, Kiseleva I, eds), pp 173-213:

Springer Netherlands.

Fratiglioni L, Ahlbom A, Viitanen M, Winblad B (1993) Risk factors for late-onset

Alzheimer's disease: a population-based, case-control study. Ann Neurol 33:258-266.

French LR, Schuman LM, Mortimer JA, Hutton JT, Boatman RA, Christians B (1985) A

case-control study of dementia of the Alzheimer type. Am J Epidemiol 121:414-421.

Fujitsu K, Kuwabara T, Muramoto M, Hirata K, Mochimatsu Y (1988) Traumatic

Intraventricular Hemorrhage: Report of Twenty-six Cases and Consideration of the

Pathogenic Mechanism. Neurosurgery 23:423-430.

257

Fukumoto H, Deng A, Irizarry MC, Fitzgerald ML, Rebeck GW (2002) Induction of the

cholesterol transporter ABCA1 in central nervous system cells by liver X receptor

agonists increases secreted Abeta levels. J Biol Chem 277:48508-48513.

Funk JR, Duma SM, Manoogian SJ, Rowson S (2007) Biomechanical risk estimates for

mild traumatic brain injury. Annu Proc Assoc Adv Automot Med 51:343-361.

Gabbi C, Warner M, Gustafsson JA (2014) Action mechanisms of Liver X Receptors.

Biochem Biophys Res Commun 446:647-650.

Gadd GW (1966) Use of a Weighted-Impulse Criterion for Estimation Injury Hazard. In:

10th Stapp Car Crash Conf, pp 164-174: Society of Automotive Engineers.

Gallyas F, Guldner FH, Zoltay G, Wolff JR (1990) Golgi-like demonstration of "dark"

neurons with an argyrophil III method for experimental neuropathology. Acta

Neuropathol 79:620-628.

Gavett BE, Stern RA, McKee AC (2011) Chronic traumatic encephalopathy: a potential

late effect of sport-related concussive and subconcussive head trauma. Clin Sports Med

30:179-188.

Geddes JF, Whitwell HL, Graham DI (2000) Traumatic axonal injury: practical issues for

diagnosis in medicolegal cases. Neuropathol Appl Neurobiol 26:105-116.

258

Geddes JF, Vowles GH, Beer TW, Ellison DW (1997) The diagnosis of diffuse axonal

injury: implications for forensic practice. Neuropathol Appl Neurobiol 23:339-347.

Genis L, Chen Y, Shohami E, Michaelson DM (2000) Tau hyperphosphorylation in

apolipoprotein E-deficient and control mice after closed head injury. J Neurosci Res

15:559-564.

Gennarelli TA (1993) Mechanisms of brain injury. J Emerg Med 11 Suppl 1:5-11.

Gennarelli TA, Thibault LE (1982) Biomechanics of acute subdural hematoma. J

Trauma 22:680-686.

Gennarelli TA, Graham DI (2005) Neuropathology. In: Textbook of traumatic brain

injury, 1st Edition (Silver JM, McAllister TW, Yudofsky SC, eds), pp 27-50. Washington,

DC: American Psychiatric Publishing, Inc.

Gennarelli TA, Adams JH, Graham DI (1981) Acceleration induced head injury in the

monkey.I. The model, its mechanical and physiological correlates. Acta Neuropathol

Suppl 7:23-25.

Gennarelli TA, Thibault LE, Adams JH, Graham DI, Thompson CJ, Marcincin RP (1982)

Diffuse axonal injury and traumatic coma in the primate. Ann Neurol 12:564-574.

259

Gentleman SM, Nash MJ, Sweeting CJ, Graham DI, Roberts GW (1993) Beta-amyloid

precursor protein (beta APP) as a marker for axonal injury after head injury. Neurosci

Lett 160:139-144.

Gentleman SM, Leclercq PD, Moyes L, Graham DI, Smith C, Griffin WS, Nicoll JA

(2004) Long-term intracerebral inflammatory response after traumatic brain injury.

Forensic Sci Int 146:97-104.

Gentleman SM, Greenberg BD, Savage MJ, Noori M, Newman SJ, Roberts GW, Griffin

WS, Graham DI (1997) A beta 42 is the predominant form of amyloid beta-protein in the

brains of short-term survivors of head injury. Neuroreport 8:1519-1522.

Gerberding JL, Binder S (2003) The report to congress on mild traumatic brain injury in

the United States: Steps to prevent a serious public health problem. In, pp 1-45. Atlanta,

GA: Centers for Disease Control and Prevention.

Ghisletti S, Huang W, Ogawa S, Pascual G, Lin M-E, Willson TM, Rosenfeld MG, Glass

CK (2007) Parallel SUMOylation-Dependent Pathways Mediate Gene- and Signal-

Specific Transrepression by LXRs and PPARγ. Molecular Cell 25:57-70.

Giacino JT, Whyte J, Bagiella E, Kalmar K, Childs N, Khademi A, Eifert B, Long D, Katz

DI, Cho S, Yablon SA, Luther M, Hammond FM, Nordenbo A, Novak P, Mercer W,

Maurer-Karattup P, Sherer M (2012) Placebo-Controlled Trial of Amantadine for Severe

Traumatic Brain Injury. New England Journal of Medicine 366:819-826.

260

Giannakopoulos P, Herrmann FR, Bussiere T, Bouras C, Kovari E, Perl DP, Morrison

JH, Gold G, Hof PR (2003) Tangle and neuron numbers, but not amyloid load, predict

cognitive status in Alzheimer's disease. Neurology 60:1495-1500.

Giza CC, Hovda DA (2001) The Neurometabolic Cascade of Concussion. J Athl Train

36:228-235.

Giza CC, Kutcher JS, Ashwal S, Barth J, Getchius TS, Gioia GA, Gronseth GS,

Guskiewicz K, Mandel S, Manley G, McKeag DB, Thurman DJ, Zafonte R (2013)

Summary of evidence-based guideline update: Evaluation and management of

concussion in sports: Report of the Guideline Development Subcommittee of the

American Academy of Neurology. Neurology 80:2250-2257.

Goate A et al. (1991) Segregation of a missense mutation in the amyloid precursor

protein gene with familial Alzheimer's disease. Nature 349:704-706.

Goedert M, Spillantini MG, Cairns NJ, Crowther RA (1992) Tau proteins of Alzheimer

paired helical filaments: abnormal phosphorylation of all six brain isoforms. Neuron

8:159-168.

Goedert M, Spillantini MG, Jakes R, Rutherford D, Crowther RA (1989) Multiple

isoforms of human microtubule-associated protein tau: sequences and localization in

neurofibrillary tangles of Alzheimer's disease. Neuron 3:519-526.

261

Goldstein LE et al. (2012) Chronic traumatic encephalopathy in blast-exposed military

veterans and a blast neurotrauma mouse model. Sci Transl Med 4:143ra160.

Gong JS, Kobayashi H, Hayashi H, Zou K, Sawamura N, Fujita SC, Yanagisawas K,

Michikawa M (2002) Apolipoprotein E (ApoE) isoform-dependent lipid release from

astrocytes prepared from hyman ApoE3 and ApoE4 knock-in mice. J Biol Chem

277:29919-29926.

Goodwin B, Watson MA, Kim H, Miao J, Kemper JK, Kliewer SA (2003) Differential

regulation of rat and human CYP7A1 by the nuclear oxysterol receptor liver X receptor-

alpha. Mol Endocrinol 17:386-394.

Gordon I, Grauer E, Genis I, Sehayek E, Michaelson DM (1995) Memory deficits and

cholinergic impairments in apolipoprotein E-deficient mice. Neurosci Lett 199:1-4.

Gorrie C, Oakes S, Duflou J, Blumbergs P, Waite PM (2002) Axonal injury in children

after motor vehicle crashes: extent, distribution, and size of axonal swellings using beta-

APP immunohistochemistry. J Neurotrauma 19:1171-1182.

Gosch HH, Gooding E, Schneider RC (1970) The lexan calvarium for the study of

cerebral responses to acute trauma. J Trauma Acute Care Surg 10:370-385.

262

Gotz J, Gladbach A, Pennanen L, van Eersel J, Schild A, David D, Ittner LM (2010)

Animal models reveal role for tau phosphorylation in human disease. Biochim Biophys

Acta 1802:860-871.

Grady MS, McLaughlin MR, Christman CW, Valadka AB, Fligner CL, Povlishock JT

(1993) The use of antibodies targeted against the neurofilament subunits for the

detection of diffuse axonal injury in humans. J Neuropathol Exp Neurol 52:143-152.

Graham DI, Adams JH, Doyle D (1978) Ischaemic brain damage in fatal non-missile

head injuries. J Neurol Sci 39:213-234.

Graham DI, McIntosh TK, Maxwell WL, Nicoll JA (2000) Recent Advances in

Neurotrauma. J Neuropathol Exp Neurol 59:641-651.

Greaves LL, Van Toen C, Melnyk A, Koenig L, Zhu Q, Tredwell S, Mulpuri K, Cripton

PA (2009) Pediatric and adult three-dimensional cervical spine kinematics: effect of age

and sex through overall motion. Spine (Phila Pa 1976) 34:1650-1657.

Greenwald RM, Gwin JT, Chu JJ, Crisco JJ (2008) Head impact severity measures for

evaluating mild traumatic brain injury risk exposure. Neurosurgery 62:789-798;

discussion 798.

Grefhorst A, Elzinga BM, Voshol PJ, Bl”sch T, Kok T, Bloks VW, van der Sluijs FH,

Havekes LM, Romijn JA, Verkade HJ, Kuipers F (2002) Stimulation of lipogenesis by

263

pharmacological activation of the liver X receptor leads to production of large,

triglyceride-rich very low density lipoprotein particles. J Biol Chem 277:34182-34190.

Groot PH et al. (2005) Synthetic LXR agonists increase LDL in CETP species. J Lipid

Res 46:2182-2191.

Grootendorst J, Bour A, Vogel E, Kelche C, Sullivan PM, Dodart JC, Bales K, Mathis C

(2005) Human apoE targeted replacement mouse lines: h-apoE4 and h-apoE3 mice

differ on spatial memory performance and avoidance behavior. Behav Brain Res 159:1-

14.

Guo Z, Cupples LA, Kurz A, Auerbach SH, Volicer L, Chui H, Green RC, Sadovnick AD,

Duara R, DeCarli C, Johnson K, Go RC, Growdon JH, Haines JL, Kukull WA, Farrer LA

(2000) Head injury and the risk of AD in the MIRAGE study. Neurology 54:1316-1323.

Gurdjian ES (1975) Re-evaluation of the biomechanics of blunt impact injury of the

head. Surg Gynecol Obstet 140:845-850.

Gurdjian ES, Lissner HR (1945) Deformation of the skull in head injury; a study with the

stresscoat technique. Surg Gynecol Obstet 81:679-687.

Gurdjian ES, Lissner HR (1946) Study of deformations of the skull by the stresscoat

technique. Anat Rec 94:465.

264

Gurdjian ES, Webster JE (1947) The mechanism and management of injuries of the

head. J Am Med Assoc 134:1072-1077.

Gurdjian ES, Lissner HR (1947) Deformations of the skull in head injury as studied by

the stresscoat technic. Am J Surg 73:269-281.

Gurdjian ES, Lissner HR, Webster JE (1947) The mechanism of production of linear

skull fracture; further studies on deformation of the skull by the stresscoat technique.

Surg Gynecol Obstet 85:195-210.

Gurdjian ES, Webster JE, Lissner HR (1955) Observations on the mechanism of brain

concussion, contusion, and laceration. Surg Gynecol Obstet 101:680-690.

Gurdjian ES, Roberts VL, Thomas LM (1966) Tolerance curves of acceleration and

intracranial pressure and protective index in experimental head injury. J Trauma 6:600-

604.

Gurdjian ES, Lissner HR, Evans FG, Patrick LM, Hardy WG (1961) Intracranial pressure

and acceleration accompanying head impacts in human cadavers. Surg Gynecol Obstet

113:185-190.

Gururaj G (2002) Epidemiology of traumatic brain injuries: Indian scenario. Neurol Res

24:24-28.

265

Guskiewicz KM (2003) Assessment of postural stability following sport-related

concussion. Curr Sports Med Rep 2:24-30.

Gusmao SN, Pittella JE (2003) Acute subdural hematoma and diffuse axonal injury in

fatal road traffic accident victims: a clinico-pathological study of 15 patients. Arq

Neuropsiquiatr 61:746-750.

Gutierrez E, Huang Y, Haglid K, Bao F, Hansson HA, Hamberger A, Viano D (2001) A

new model for diffuse brain injury by rotational acceleration: I model, gross appearance,

and astrocytosis. J Neurotrauma 18:247-257.

Haddad BF, Lissner HR, Webster JE, Gurdjian ES (1955) Experimental concussion;

relation to acceleration to physiologic effect. Neurology 5:798-800.

Hamberger A, Huang YL, Zhu H, Bao F, Ding M, Blennow K, Olsson A, Hansson HA,

Viano D, Haglid KG (2003) Redistribution of neurofilaments and accumulation of beta-

amyloid protein after brain injury by rotational acceleration of the head. J Neurotrauma

20:169-178.

Han SH, Chung SY (2000) Marked hippocampal neuronal damage without motor

deficits after mild concussive-like brain injury in apolipoprotein E-deficient mice. Ann N

Y Acad Sci 903:357-365.

266

Handelmann GE, Boyles JK, Weisgraber KH, Mahley RW, Pitas RE (1992) Effects of

apolipoprotein E, beta-very low density lipoproteins, and cholesterol on the extension of

neurites by rabbit dorsal root ganglion neurons in vitro. J Lipid Res 33:1677-1688.

Harders A, Kakarieka A, Braakman R (1996) Traumatic subarachnoid hemorrhage and

its treatment with nimodipine. German tSAH Study Group. J Neurosurg 85:82-89.

Hardy J, Higgins G (1992) Alzheimer's disease: the amyloid cascade hypothesis.

Science 256:184-185.

Hardy J, Selkoe DJ (2002) The amyloid hypothesis of Alzheimer's disease: progress

and problems on the road to therapeutics. Science 297:353-356.

Hardy WN, Khalil TB, King AI (1994) Literature review of head injury biomechanics. Int

Journal Impact Eng 15:561-586.

Hardy WN, Foster CD, Mason MJ, Yang KH, King AI, Tashman S (2001) Investigation

of Head Injury Mechanisms Using Neutral Density Technology and High-Speed Biplanar

X-ray. Stapp Car Crash J 45:337-368.

Hardy WN, Mason MJ, Foster CD, Shah CS, Kopacz JM, Yang KH, King AI, Bishop J,

Bey M, Anderst W, Tashman S (2007) A study of the response of the human cadaver

head to impact. Stapp Car Crash J 51:17-80.

267

Harris FM, Tesseur I, Brecht WJ, Xu Q, Mullendorff K, Chang S, Wyss-Coray T, Mahley

RW, Huang Y (2004) Astroglial regulation of apolipoprotein E expression in neuronal

cells. Implications for Alzheimer's disease. J Biol Chem 279:3862-3868.

Harris FM, Brecht WJ, Xu Q, Tesseur I, Kekonius L, Wyss-Coray T, Fish JD, Masliah E,

Hopkins PC, Scearce-Levie K, Weisgraber KH, Mucke L, Mahley RW, Huang Y (2003)

Carboxyl-terminal-truncated apolipoprotein E4 causes Alzheimer's disease-like

neurodegeneration and behavioral deficits in transgenic mice. Proc Natl Acad Sci U S A

100:10966-10971.

Harris OA, Muh CR, Surles MC, Pan Y, Rozycki G, Macleod J, Easley K (2009) Discrete

cerebral hypothermia in the management of traumatic brain injury: a randomized

controlled trial. J Neurosurg 110:1256-1264.

Hatters DM, Peters-Libeu CA, Weisgraber KH (2006) Apolipoprotein E structure:

insights into function. Trends Biochem Sci 31:445-454.

Hatters DM, Budamagunta MS, Voss JC, Weisgraber KH (2005) Modulation of

apolipoprotein E structure by domain interaction: differences in lipid-bound and lipid-free

forms. J Biol Chem 280:34288-34295.

Henn H-W (1998) Crash Tests and the Head Injury Criterion. Teach Math Appl 17:162-

170.

268

Hertz E (1993) A Note on the Head Injury Criteria (HIC) as a predictor of the Risk of

Skull Fracture. In: Proceedings of the 37th Annual Conference of Association for the

Advancement of Automotive Medicine, pp 303-312. San Antonio, TX: Association for

the Advancement of Automotive Medicine.

Herz J, Chen Y (2006) Reelin, lipoprotein receptors and synaptic plasticity. Nat Rev

Neurosci 7:850-859.

Hiekkanen H, Kurki T, Brandstack N, Kairisto V, Tenovuo O (2009) Association of injury

severity, MRI-results and ApoE genotype with 1-year outcome in mainly mild TBI: a

preliminary study. Brain Inj 23:396-402.

Hirsch-Reinshagen V, Zhou S, Burgess BL, Bernier L, McIsaac SA, Chan JY, Tansley

GH, Cohn JS, Hayden MR, Wellington CL (2004) Deficiency of ABCA1 impairs

apolipoprotein E metabolism in brain. J Biol Chem 279:41197-41207.

Hirsch-Reinshagen V, Maia LF, Burgess BL, Blain JF, Naus KE, McIsaac SA, Parkinson

PF, Chan JY, Tansley GH, Hayden MR, Poirier J, Van Nostrand WE, Wellington CL

(2005) The absence of ABCA1 decreases soluble apoE levels but does not diminish

amyloid deposition in two murine models of Alzheimer's Disease. J Biol Chem

280:43243-43256.

Hobbs HH, Rader DJ (1999) ABC1: connecting yellow tonsils, neuropathy, and very low

HDL. J Clin Invest 104:1015-1017.

269

Hodgson VR, Gurdjian ES, Thomas LM (1966) Experimental skull deformation and

brain displacement demonstrated by flash x-ray technique. J Neurosurg 25:549-552.

Hoe HS, Freeman J, Rebeck GW (2006) Apolipoprotein E decreases tau kinases and

phospho-tau levels in primary neurons. Mol Neurodegener 1:18.

Hoge CW, McGurk D, Thomas JL, Cox AL, Engel CC, Castro CA (2008) Mild traumatic

brain injury in US soldiers returning from Iraq. New Eng J Med 358:453-463.

Holbourn AH (1943) Mechanics of head injury. The Lancet 2:438-441.

Holbourn AH (1945) Mechanics of brain injuries. Br Med Bull 3:147-149.

Holmin S, Hojeberg B (2004) In situ detection of intracerebral cytokine expression after

human brain contusion. Neurosci Lett 369:108-114.

Holmin S, Soderlund J, Biberfeld P, Mathiesen T (1998) Intracerebral inflammation after

human brain contusion. Neurosurgery 42:291-298; discussion 298-299.

Holtzman DM, Pitas RE, Kilbridge J, Nathan B, Mahley RW, Bu G, Schwartz AL (1995)

Low density lipoprotein receptor-related protein mediates apolipoprotein E-dependent

neurite outgrowth in a central nervous system-derived neuronal cell line. Proc Natl Acad

Sci U S A 92:9480-9484.

270

Horlein AJ, Naar AM, Heinzel T, Torchia J, Gloss B, Kurokawa R, Ryan A, Kamei Y,

Soderstrom M, Glass CK, Rosenfeld MG (1995) Ligand-independent repression by the

thyroid hormone receptor mediated by a nuclear receptor co-repressor. Nature 377:397-

404.

Houck KA, Borchert KM, Hepler CD, Thomas JS, Bramlett KS, Michael LF, Burris TP

(2004) T0901317 is a dual LXR/FXR agonist. Mol Genet Metab 83:184-187.

Hu X, Crick SL, Bu G, Frieden C, Pappu RV, Lee J-M (2009) Amyloid seeds formed by

cellular uptake, concentration, and aggregation of the amyloid-beta peptide. Proc Natl

Acad Sci U S A.

Huang Y, Liu XQ, Wyss-Coray T, Brecht WJ, Sanan DA, Mahley RW (2001)

Apolipoprotein E fragments present in Alzheimer's disease brains induce neurofibrillary

tangle-like intracellular inclusions in neurons. Proc Natl Acad Sci U S A 98:8838-8843.

Hubschmann OR, Kornhauser D (1983) Effects of intraparenchymal hemorrhage on

extracellular cortical potassium in experimental head trauma. J Neurosurg 59:289-293.

Huh JW, Franklin MA, Widing AG, Raghupathi R (2006) Regionally distinct patterns of

calpain activation and traumatic axonal injury following contusive brain injury in

immature rats. Dev Neurosci 28:466-476.

271

Hunter JV, Wilde EA, Tong KA, Holshouser BA (2012) Emerging imaging tools for use

with traumatic brain injury research. J Neurotrauma 29:654-671.

Huuskonen J, Fielding PE, Fielding CJ (2004) Role of p160 coactivator complex in the

activation of liver X receptor. Arterioscler Thromb Vasc Biol 24:703-708.

Hyder AA, Wunderlich CA, Puvanachandra P, Gururaj G, Kobusingye OC (2007) The

impact of traumatic brain injuries: a global perspective. NeuroRehabilitation 22:341-353.

Ikonomovic MD, Uryu K, Abrahamson EE, Ciallella JR, Trojanowski JQ, Lee VM, Clark

RS, Marion DW, Wisniewski SR, DeKosky ST (2004) Alzheimer's pathology in human

temporal cortex surgically excised after severe brain injury. Exp Neurol 190:192-203.

Inoue T, Lin A, Ma X, McKenna SL, Creasey GH, Manley GT, Ferguson AR, Bresnahan

JC, Beattie MS (2013) Combined SCI and TBI: recovery of forelimb function after

unilateral cervical spinal cord injury (SCI) is retarded by contralateral traumatic brain

injury (TBI), and ipsilateral TBI balances the effects of SCI on paw placement. Exp

Neurol 248:136-147.

Iqbal K, Liu F, Gong CX, Grundke-Iqbal I (2010) Tau in Alzheimer disease and related

tauopathies. Curr Alzheimer Res 7:656-664.

Irizarry MC, Cheung BS, Rebeck GW, Paul SM, Bales KR, Hyman BT (2000)

Apolipoprotein E affects the amount, form, and anatomical distribution of amyloid beta-

272

peptide deposition in homozygous APP(V717F) transgenic mice. Acta Neuropathol

100:451-458.

Iwata A, Browne KD, Chen XH, Yugushi T, Smith DH (2005) Traumatic brain injury

induces biphasic upregulation of ApoE and ApoJ protein in rats. J Neurosci Res 82:103-

114.

Iwata A, Stys PK, Wolf JA, Chen XH, Taylor AG, Meaney DF, Smith DH (2004)

Traumatic axonal injury induces proteolytic cleavage of the voltage-gated sodium

channels modulated by tetrodotoxin and protease inhibitors. J Neurosci 24:4605-4613.

Iwata N, Tsubuki S, Takaki Y, Shirotani K, Lu B, Gerard NP, Gerard C, Hama E, Lee H-

J, Saido TC (2001) Metabolic Regulation of Brain Aβ by Neprilysin. Science 292:1550-

1552.

Iwata N, Tsubuki S, Takaki Y, Watanabe K, Sekiguchi M, Hosoki E, Kawashima-

Morishima M, Lee HJ, Hama E, Sekine-Aizawa Y, Saido TC (2000) Identification of the

major Abeta1-42-degrading catabolic pathway in brain parenchyma: suppression leads

to biochemical and pathological deposition. Nat Med 6:143-150.

Iwatsubo T (2004) The gamma-secretase complex: machinery for intramembrane

proteolysis. Curr Opin Neurobiol 14:379-383.

273

Jamroz-Wisniewska A, W¢jcicka G, Horoszewicz K, Beltowski J (2007) Liver X

receptors (LXRs). Part II: Non-lipid effects, role in pathology, and therapeutic

implications. Postepy Hig Med Dosw 61:760-785.

Janowski BA, Willy PJ, Devi TR, Falck JR, Mangelsdorf DJ (1996) An oxysterol

signalling pathway mediated by the nuclear receptor LXR alpha. Nature 383:728-731.

Janowski BA, Grogan MJ, Jones SA, Wisely GB, Kliewer SA, Corey EJ, Mangelsdorf DJ

(1999) Structural requirements of ligands for the oxysterol liver X receptors LXRalpha

and LXRbeta. Proc Natl Acad Sci U S A 96:266-271.

Jha A, Weintraub A, Allshouse A, Morey C, Cusick C, Kittelson J, Harrison-Felix C,

Whiteneck G, Gerber D (2008) A randomized trial of modafinil for the treatment of

fatigue and excessive daytime sleepiness in individuals with chronic traumatic brain

injury. J Head Trauma Rehabil 23:52-63.

Ji ZS, Miranda RD, Newhouse YM, Weisgraber KH, Huang Y, Mahley RW (2002)

Apolipoprotein E4 potentiates amyloid beta peptide-induced lysosomal leakage and

apoptosis in neuronal cells. J Biol Chem 277:21821-21828.

Jiang JY, Xu W, Li WP, Gao GY, Bao YH, Liang YM, Luo QZ (2006) Effect of long-term

mild hypothermia or short-term mild hypothermia on outcome of patients with severe

traumatic brain injury. J Cereb Blood Flow Metab 26:771-776.

274

Jiang JY, Xu W, Li WP, Xu WH, Zhang J, Bao YH, Ying YH, Luo QZ (2005) Efficacy of

standard trauma craniectomy for refractory intracranial hypertension with severe

traumatic brain injury: a multicenter, prospective, randomized controlled study. J

Neurotrauma 22:623-628.

Jiang Q, Lee CY, Mandrekar S, Wilkinson B, Cramer P, Zelcer N, Mann K, Lamb B,

Willson TM, Collins JL, Richardson JC, Smith JD, Comery TA, Riddell D, Holtzman DM,

Tontonoz P, Landreth GE (2008a) ApoE promotes the proteolytic degradation of Abeta.

Neuron 58:681-693.

Jiang Y, Sun XC, Gui L, Tang WY, Zhen LP, Gu YJ, Wu HT (2008b) Lack of association

between apolipoprotein E promoters in epsilon-4 carriers and worsening on computed

tomography in early stage of traumatic brain injury. Acta Neurochir (Suppl) 105:233-

236.

Johnson VE, Stewart W, Smith DH (2010) Traumatic brain injury and amyloid-beta

pathology: a link to Alzheimer's disease? Nat Rev Neurosci 11:361-370.

Johnson VE, Stewart W, Smith DH (2012) Widespread tau and amyloid-beta pathology

many years after a single traumatic brain injury in humans. Brain Pathol 22:142-149.

Johnson VE, Stewart W, Smith DH (2013) Axonal pathology in traumatic brain injury.

Exp Neurol 246:35-43.

275

Johnson VE, Stewart W, Graham DI, Stewart JE, Praestgaard AH, Smith DH (2009) A

neprilysin polymorphism and amyloid-beta plaques after traumatic brain injury. J

Neurotrauma 26:1197-1202.

Jordan BD (2013) The clinical spectrum of sport-related traumatic brain injury. Nat Rev

Neurol 9:222-230.

Kakarieka A, Braakman R, Schakel EH (1994) Clinical significance of the finding of

subarachnoid blood on CT scan after head injury. Acta Neurochir (Wien) 129:1-5.

Kakuda N, Funamoto S, Yagishita S, Takami M, Osawa S, Dohmae N, Ihara Y (2006)

Equimolar production of amyloid beta-protein and amyloid precursor protein intracellular

domain from beta-carboxyl-terminal fragment by gamma-secretase. J Biol Chem

281:14776-14786.

Kamal A, Stokin GB, Yang Z, Xia CH, Goldstein LS (2000) Axonal transport of amyloid

precursor protein is mediated by direct binding to the kinesin light chain subunit of

kinesin-I. Neuron 28:449-459.

Kamal A, Almenar-Queralt A, LeBlanc JF, Roberts EA, Goldstein LS (2001) Kinesin-

mediated axonal transport of a membrane compartment containing beta-secretase and

presenilin-1 requires APP. Nature 414:643-648.

276

Kane MJ, Angoa-Perez M, Briggs DI, Viano DC, Kreipke CW, Kuhn DM (2012) A mouse

model of human repetitive mild traumatic brain injury. J Neurosci Methods 203:41-49.

Kanekiyo T, Xu H, Bu G ApoE and Aβ in Alzheimer’s Disease: Accidental Encounters or

Partners? Neuron 81:740-754.

Kanemitsu H, Tomiyama T, Mori H (2003) Human neprilysin is capable of degrading

amyloid beta peptide not only in the monomeric form but also the pathological

oligomeric form. Neurosci Lett 350:113-116.

Karran E, Mercken M, De Strooper B (2011) The amyloid cascade hypothesis for

Alzheimer's disease: an appraisal for the development of therapeutics. Nat Rev Drug

Discov 10:698-712.

Katayama Y, Becker DP, Tamura T, Hovda DA (1990) Massive increases in

extracellular potassium and the indiscriminate release of glutamate following concussive

brain injury. J Neurosurg 73:889-900.

Katsnelson A (2011) Gene tests for brain injury still far from the football field. Nat Med

17:638-638.

Katz A, Udata C, Ott E, Hickey L, Burczynski ME, Burghart P, Vesterqvist O, Meng X

(2009) Safety, pharmacokinetics and pharmacodynamics of single doses of LXR-623, a

novel liver-X-receptor agonist, in healthy participants. J Clin Pharmacol 49:643-649.

277

Katzman R, Aronson M, Fuld P, Kawas C, Brown T, Morgenstern H, Frishman W, Gidez

L, Eder H, Ooi WL (1989) Development of dementing illnesses in an 80-year-old

volunteer cohort. Ann Neurol 25:317-324.

Kay AD, Day SP, Kerr M, Nicoll JA, Packard CJ, Caslake MJ (2003) Remodeling of

cerebrospinal fluid lipoprotein particles after human traumatic brain injury. J

Neurotrauma 20:717-723.

Kay T, Harrington DE, Admas R, Anderson T, Berrol S, Cicerone K, Dahlberg C, Gerber

D, Goka R, Harley P, Hilt J, Horn L, Lehmkuhl D, Malec J (1993) Definition of mild

traumatic brain injury. In: J Head Trauma Rehabil, pp 86-87.

Kennedy MA, Venkateswaran A, Tarr PT, Xenarios I, Kudoh J, Shimizu N, Edwards PA

(2001) Characterization of the human ABCG1 gene: liver X receptor activates an

internal promoter that produces a novel transcript encoding an alternative form of the

protein. J Biol Chem 276:39438-39447.

Kennedy MA, Barrera GC, Nakamura K, Bald n A, Tarr P, Fishbein MC, Frank J,

Francone OL, Edwards PA (2005) ABCG1 has a critical role in mediating cholesterol

efflux to HDL and preventing cellular lipid accumulation. Cell Metabolism 1 121-131.

Khoshyomn S, Tranmer BI (2004) Diagnosis and management of pediatric closed head

injury. Semin Pediatr Surg 13:80-86.

278

Khuman J, Meehan WP, 3rd, Zhu X, Qiu J, Hoffmann U, Zhang J, Giovannone E, Lo

EH, Whalen MJ (2011) Tumor necrosis factor alpha and Fas receptor contribute to

cognitive deficits independent of cell death after concussive traumatic brain injury in

mice. J Cereb Blood Flow Metab 31:778-789.

Kiehn O, Kjaerulff O (1998) Distribution of central pattern generators for rhythmic motor

outputs in the spinal cord of limbed vertebrates. Ann N Y Acad Sci 860:110-129.

Kigerl KA, Gensel JC, Ankeny DP, Alexander JK, Donnelly DJ, Popovich PG (2009)

Identification of two distinct macrophage subsets with divergent effects causing either

neurotoxicity or regeneration in the injured mouse spinal cord. J Neurosci 29:13435-

13444.

Kilbourne M, Kuehn R, Tosun C, Caridi J, Keledjian K, Bochicchio G, Scalea T,

Gerzanich V, Simard JM (2009) Novel model of frontal impact closed head injury in the

rat. J Neurotrauma 26:2233-2243.

Kim J, Castellano JM, Jiang H, Basak JM, Parsadanian M, Pham V, Mason SM, Paul

SM, Holtzman DM (2009) Overexpression of low-density lipoprotein receptor in the

brain markedly inhibits amyloid deposition and increases extracellular A beta clearance.

Neuron 64:632-644.

King AI (2000) Fundamentals of impact biomechanics: Part I--Biomechanics of the

head, neck, and thorax. Annu Rev Biomed Eng 2:55-81.

279

King AI, Yang KH, Hardy WN (2011) Recent firsts in cadaveric impact biomechanics

research. Clin Anat 24:294-308.

King AI, Yang KH, Zhang L, Hardy W (2004) Is Rotational Acceleration More Injurious to

the Brain Than Linear Acceleration? In: Frontiers in Biomedical Engineering (Hwang

NHC, Woo SLY, eds), pp 135-147: Springer US.

King AI, Yang KH, Zhang L, Hardy W, Viano DC (2003) Is head injury caused by linear

or angular acceleration? In: Proceedings of the International Research Conference on

the Biomechanics of Impacts (IRCOBI), pp 1-12. Lisbon, Portugal.

Kitamura HW, Hamanaka H, Watanabe M, Wada K, Yamazaki C, Fujita SC, Manabe T,

Nukina N (2004) Age-dependent enhancement of hippocampal long-term potentiation in

knock-in mice expressing human apolipoprotein E4 instead of mouse apolipoprotein E.

Neurosci Lett 369:173-178.

Kitamura K (1981) Therapeutic effect of pyritinol on sequelae of head injuries. J Int Med

Res 9:215-221.

Kjaerulff O, Kiehn O (1996) Distribution of networks generating and coordinating

locomotor activity in the neonatal rat spinal cord in vitro: a lesion study. J Neurosci

16:5777-5794.

280

Koch S, Donarski N, Goetze K, Kreckel M, Sturernburg HJ, Buhmann C, Beisiegel U

(2001) Characterization of four lipoprotein classes in human cerebrospinal fluid. J Lipid

Res 42:1143-1151.

Koldamova R, Staufenbiel M, Lefterov I (2005a) Lack of ABCA1 considerably

decreased brain apoE level and increases amyloid deposition in APP23 mice. J Biol

Chem 280:43224-43235.

Koldamova R, Fitz NF, Lefterov I (2010) The role of ATP-binding cassette transporter

A1 in Alzheimer's disease and neurodegeneration. Biochim Biophys Acta 180:824-830.

Koldamova RP, Lefterov IM, Lefterova MI, Lazo JS (2001) Apolipoprotein A1 directly

interacts with amyloid precursor protein and inhibits A beta aggregation and toxicity.

Biochemistry 40:3553-3560.

Koldamova RP, Lefterov IM, Ikonomovic MD, Skoko J, Lefterov PI, Isanski BA,

DeKosky ST, Lazo JS (2003) 22R-hydroxycholesterol and 9-cis-retinoic acid induce

ATP-binding cassette transporter A1 expression and cholesterol efflux in brain cells and

decrease amyloid á secretion. J Biol Chem 278:13244-13256.

Koldamova RP, Lefterov IM, Staufenbiel M, Wolfe D, Huang S, Glorioso JC, Walter M,

Roth MG, Lazo JS (2005b) The liver X receptor ligand T0901317 decreases amyloid

beta production in vitro and in a mouse model of Alzheimer's disease. J Biol Chem

280:4079-4088.

281

Koo EH, Sisodia SS, Archer DR, Martin LJ, Weidemann A, Beyreuther K, Fischer P,

Masters CL, Price DL (1990) Precursor of amyloid protein in Alzheimer disease

undergoes fast anterograde axonal transport. Proc Natl Acad Sci U S A 87:1561-1565.

Korwek KM, Trotter JH, Ladu MJ, Sullivan PM, Weeber EJ (2009) ApoE isoform-

dependent changes in hippocampal synaptic function. Mol Neurodegener 4:21.

Koshinaga M, Katayama Y, Fukushima M, Oshima H, Suma T, Takahata T (2000)

Rapid and widespread microglial activation induced by traumatic brain injury in rat brain

slices. J Neurotrauma 17:185-192.

Krave U, Hojer S, Hansson HA (2005) Transient, powerful pressures are generated in

the brain by a rotational acceleration impulse to the head. Eur J Neurosci 21:2876-2882.

Krave U, Al-Olama M, Hansson HA (2011) Rotational acceleration closed head flexion

trauma generates more extensive diffuse brain injury than extension trauma. J

Neurotrauma 28:57-70.

Krimbou L, Denis B, Haidar M, Carrier M, Marcil M, Genest J, Jr. (2004) Molecular

interactions between apoE and ABCA1: impact on apoE lipidation. J Lipid Res 45:839-

848.

282

Ksiezak-Reding H, Morgan K, Mattiace LA, Davies P, Liu WK, Yen SH, Weidenheim K,

Dickson DW (1994) Ultrastructure and biochemical composition of paired helical

filaments in corticobasal degeneration. Am J Pathol 145:1496-1508.

Kuppa S (2004) Injury Criteria for Side Impact Dummies. In: National Transportation

Biomechanics Research Center.

Kwon BK, Hillyer J, Tetzlaff W (2010) Translational research in spinal cord injury: a

survey of opinion from the SCI community. J Neurotrauma 27:21-33.

LaDu MJ, Falduto MT, Manelli AM, Reardon CA, Getz GS, Frail DE (1994) Isoform-

specific binding of apolipoprotein E to beta-amyloid. J Biol Chem 269:23403-23406.

LaDu MJ, Pederson TM, Frail DE, Reardon CA, Getz GS, Falduto MT (1995)

Purification of apolipoprotein E attenuates isoform-specific binding to beta-amyloid. J

Biol Chem 270:9039-9042.

LaDu MJ, Gilligan SM, Lukens JR, Cabana VG, Reardon CA, Van Eldik LJ, Holtzman

DA (1998) Nascent astrocyte particles differ from lipoproteins in CSF. J Neurochem

70:2070-2081.

LaDu MJ, Reardon C, Van Eldik L, Fagan AM, Bu G, Holtzmann D, Getz GS (2000a)

Lipoproteins in the central nervous system. Ann New York Acad Sci 903:167-175.

283

LaDu MJ, Reardon C, Van Eldik L, Fagan AM, Bu G, Holtzman D, Getz GS (2000b)

Lipoproteins in the central nervous system. Ann N Y Acad Sci 903:167-175.

Langlois JA, Rutland-Brown W, Wald MM (2006) The Epidemiology and Impact of

Traumatic Brain Injury: A Brief Overview. J Head Trauma Rehabil 21:375-378.

Langmann T, Klucken J, Reil M, Liebisch G, Luciani M-Fo, Chimini G, Kaminski WE,

Schmitz G (1999) Molecular cloning of the human ATP-binding cassette transporter 1

(hABC1): Evidence for sterol-dependent regulation in macrophages. Biochem Biophys

Res Commun 257:29-33.

LaPlaca MC, Simon CM, Prado GR, Cullen DK (2007) CNS injury biomechanics and

experimental models. Prog Brain Res 161:13-26.

Laskowitz DT, Horsburgh K, Roses AD (1998) Apolipoprotein E and the CNS response

to injury. J Cereb Blood Flow Metab 18:465-471.

Lau IV, Viano DC, Gamero F (1989) Invalidity of speculated injury mechanism in

autopsy reports. Injury 20:16-21.

Laurer HL, Bareyre FM, Lee VM, Trojanowski JQ, Longhi L, Hoover R, Saatman KE,

Raghupathi R, Hoshino S, Grady MS, McIntosh TK (2001) Mild head injury increasing

the brain's vulnerability to a second concussive impact. J Neurosurg 95:859-870.

284

Lawn RM, Wade DP, Couse TL, Wilcox JN (2001) Localization of human ATP-Binding

cassette transporter 1 (ABC1) in normal and atherosclerotic tissues. Arterioscler

Thromb Vasc Biol 21:378-385.

Lee G, Neve RL, Kosik KS (1989) The microtubule binding domain of tau protein.

Neuron 2:1615-1624.

Lee Y-H, Jung MG, Kang HB, Choi K-C, Haam S, Jun W, Kim Y-J, Cho HY, Yoon H-G

(2008) Effect of anti-histone acetyltransferase activity from Rosa rugosa Thunb.

(Rosaceae) extracts on androgen receptor-mediated transcriptional regulation. J

Ethnopharmacol 118:412-417.

Lefterov I, Fitz NF, Cronican AA, Fogg A, Lefterov P, Kodali R, Wetze R, Koldamova R

(2010) Apolipoprotein A-I deficiency increases cerebral amyloid angiopathy and

cognitive deficits in APP/PS1E9 mice. J Biol Chem 285 36945-36957.

Lehmann JM, Kliewer SA, Moore LB, Smith-Oliver TA, Sundseth SS, Winegar DA,

Blanchard DE, Spencer TA, Wilson TM (1997) Activation of the nuclear receptor LXR by

oxysterols defines a new hormone response pathway. J Biol Chem 272:3137-3140.

Leissring MA, Farris W, Chang AY, Walsh DM, Wu X, Sun X, Frosch MP, Selkoe DJ

(2003) Enhanced proteolysis of beta-amyloid in APP transgenic mice prevents plaque

formation, secondary pathology, and premature death. Neuron 40:1087-1093.

285

Levy-Lahad E, Wasco W, Poorkaj P, Romano DM, Oshima J, Pettingell WH, Yu CE,

Jondro PD, Schmidt SD, Wang K, et al. (1995) Candidate gene for the chromosome 1

familial Alzheimer's disease locus. Science 269:973-977.

Lewen A, Li GL, Nilsson P, Olsson Y, Hillered L (1995) Traumatic brain injury in rat

produces changes of beta-amyloid precursor protein immunoreactivity. Neuroreport

6:357-360.

Lewis TL, Cao D, Lu H, Mans RA, Su YR, Jungbauer L, Linton MF, Fazio S, Ladu MJ, Li

L (2010) Overexpression of human apolipoprotein A-I preserves cognitive function and

attenuates neuroinflammation and cerebral amyloid angiopathy in a mouse model of

Alzheimer's Disease. J Biol Chem 285:36958-36968.

Li G, Shen YC, Li YT, Chen CH, Zhau YW, Silverman JM (1992) A case-control study of

Alzheimer's disease in China. Neurology 42:1481-1488.

Li J, Kanekiyo T, Shinohara M, Zhang Y, LaDu MJ, Xu H, Bu G (2012) Differential

regulation of amyloid-beta endocytic trafficking and lysosomal degradation by

apolipoprotein E isoforms. J Biol Chem 287:44593-44601.

Li XY, Li J, Feng DF, Gu L (2010) Diffuse axonal injury induced by simultaneous

moderate linear and angular head accelerations in rats. Neuroscience 169:357-369.

286

Li Y, Zhang L, Kallakuri S, Zhou R, Cavanaugh JM (2011) Quantitative relationship

between axonal injury and mechanical response in a rodent head impact acceleration

model. J Neurotrauma 28:1767-1782.

Liang Y, Lin S, Beyer TP, Zhang Y, Wu X, Bales KR, DeMattos RB, May PC, Li SD,

Jiang XC, Eacho PI, Cao G, Paul SM (2004) A liver X receptor and retinoid X receptor

heterodimer mediates apolipoprotein E expression, secretion and cholesterol

homeostasis in astrocytes. J Neurochem 88 623-634.

Liao H, Langmann T, Schmitz G, Zhu Y (2002) Native LDL upregulation of ATP-binding

cassette transporter-1 in human vascular endothelial cells. Arterioscler Thromb Vasc

Biol 22:127-132.

Lichtman SW, Seliger G, Tycko B, Marder K (2000) Apolipoprotein E and functional

recovery from brain injury following postacute rehabilitation. Neurology 55:1536-1539.

Lighthall JW, Dixon CE, Anderson TE (1989) Experimental models of brain injury. J

Neurotrauma 6:83-97.

Lin Y, Wen L (2013) Inflammatory response following diffuse axonal injury. Int J Med Sci

10:515-521.

Lindenberg R, Freytag E (1960) The mechanism of cerebral contusions. A pathologic-

anatomic study. Arch Pathol 69:440-469.

287

Lissner HR, Lebow M, Evans FG (1960) Experimental studies on the relation between

acceleration and intracranial pressure changes in man. Surg Gynecol Obstet 111:329-

338.

Loane DJ, Byrnes KR (2010) Role of microglia in neurotrauma. Neurotherapeutics

7:366-377.

Loane DJ, Pocivavasek A, Moussa CE, Thompson R, Matsuoka Y, Faden AI, Rebeck

GW, Burns MP (2009) Amyloid precursor protein secretases as therapeutic targets for

traumatic brain injury. Nat Med 15:377-379.

Loane DL, Washington PM, Vardanian L, Pocivavsek A, Hoe HS, Duff KE, Cernak I,

Rebek GW, Faden AI, Burns MP (2011) Modulation of ABCA1 by an LXR agonist

reduces beta-amyloid levels and improves outcome after traumatic brain injury. J

Neurotrauma 28:225-236.

Lopez-Guerrero AL, Martinez-Lage JF, Gonzalez-Tortosa J, Almagro MJ, Garcia-

Martinez S, Reyes SB (2012) Pediatric crushing head injury: biomechanics and clinical

features of an uncommon type of craniocerebral trauma. Childs Nerv Syst 28:2033-

2040.

Lorent K, Overbergh L, Moechars D, De Strooper B, Van Leuven F, Van den Berghe H

(1995) Expression in mouse embryos and in adult mouse brain of three members of the

amyloid precursor protein family, of the alpha-2-macroglobulin receptor/low density

288

lipoprotein receptor-related protein and of its ligands apolipoprotein E, lipoprotein lipase,

alpha-2-macroglobulin and the 40,000 molecular weight receptor-associated protein.

Neuroscience 65:1009-1025.

Lu J, Gary KW, Neimeier JP, Ward J, Lapane KL (2012) Randomized controlled trials in

adult traumatic brain injury. Brain Inj 26:1523-1548.

Lu KT, Wang YW, Yang JT, Yang YL, Chen HI (2005) Effect of interleukin-1 on

traumatic brain injury-induced damage to hippocampal neurons. J Neurotrauma 22:885-

895.

Lu LQ, Jiang JY, Yu MK, Hou LJ, Chen ZG, Zhang GJ, Zhu C (2003) Standard large

trauma craniotomy for severe traumatic brain injury. Chin J Traumatol 6:302-304.

Lund EG, Guileyardo JM, Russell DW (1999) cDNA cloning of cholesterol 24-

hydroxylase, a mediator of cholesterol homeostasis in the brain. Proc Natl Acad Sci U S

A 96:7238-7243.

Lynch JR, Pineda JA, Morgan D, Zhang L, Warner DS, Benveniste H, Laskowitz DT

(2002) Apolipoprotein E affects the central nervous system response to injury and the

development of cerebral edema. Ann Neurol 51:113-117.

289

Lynch JR, Tang W, Wang H, Vitek MP, Bennett ER, Sullivan PM, Warner DS, Laskowitz

DT (2003) APOE genotype and an ApoE-mimetic peptide modify the systemic and

central nervous system inflammatory response. J Biol Chem 278:48529-48533.

Macciocchi S, Seel RT, Thompson N, Byams R, Bowman B (2008) Spinal cord injury

and co-occurring traumatic brain injury: assessment and incidence. Arch Phys Med

Rehabil 89:1350-1357.

Maeda J, Higuchi M, Inaji M, Ji B, Haneda E, Okauchi T, Zhang MR, Suzuki K, Suhara

T (2007) Phase-dependent roles of reactive microglia and astrocytes in nervous system

injury as delineated by imaging of peripheral benzodiazepine receptor. Brain Res

1157:100-111.

Maezawa I, Nivison M, Montine KS, Maeda N, Montine TJ (2006) Neurotoxicity from

innate immune response is greatest with targeted replacement of E4 allele of

apolipoprotein E gene and is mediated by microglial p38MAPK. FASEB J 20:797-799.

Magnoni S, Brody DL (2010) New perspectives on amyloid-beta dynamics after acute

brain injury: moving between experimental approaches and studies in the human brain.

Arch Neurol 67:1068-1073.

Magnoni S, Esparza TJ, Conte V, Carbonara M, Carrabba G, Holtzman DM, Zipfel GJ,

Stocchetti N, Brody DL (2012) Tau elevations in the brain extracellular space correlate

290

with reduced amyloid-beta levels and predict adverse clinical outcomes after severe

traumatic brain injury. Brain 135:1268-1280.

Mahley RW (1998) Apolipoprotein E: cholesterol transport protein with expanding role in

cell biology. Science 240:622-630.

Mahley RW, Huang Y (2006) Apolipoprotein (apo) E4 and Alzheimer's disease: unique

conformational and biophysical properties of apoE4 can modulate neuropathology. Acta

Neurologica Scandinavica 114:8-14.

Marder E, Bucher D (2001) Central pattern generators and the control of rhythmic

movements. Current Biology 11:R986-R996.

Marion DW, Penrod LE, Kelsey SF, Obrist WD, Kochanek PM, Palmer AM, Wisniewski

SR, DeKosky ST (1997) Treatment of traumatic brain injury with moderate hypothermia.

N Engl J Med 336:540-546.

Marklund N, Hillered L (2011) Animal modelling of traumatic brain injury in preclinical

drug development: where do we go from here? Br J Pharmacol 164:1207-1229.

Marmarou A, Foda MA, van den Brink W, Campbell J, Kita H, Demetriadou K (1994) A

new model of diffuse brain injury in rats. Part I: Pathophysiology and biomechanics. J

Neurosurg 80:291-300.

291

Martinez LO, Agerholm-Larsen B, Wang N, Chen W, Tall AR (2003) Phosphorylation of

a Pest sequence in ABCA1 promotes calpain degradation and is reversed by apoA-I. J

Biol Chem 278:37368-37374.

Martland HS (1928) Punch drunk. J Am Med Assoc 91:1103-1107.

Masson F, Thicoipe M, Aye P, Mokni T, Senjean P, Schmitt V, Dessalles PH,

Cazaugade M, Labadens P (2001) Epidemiology of severe brain injuries: a prospective

population-based study. J Trauma 51:481-489.

Masters CL, Selkoe DJ (2012) Biochemistry of amyloid beta-protein and amyloid

deposits in Alzheimer disease. Cold Spring Harb Perspect Med 2:a006262.

Mattei TA, Bond BJ, Goulart CR, Sloffer CA, Morris MJ, Lin JJ (2012) Performance

analysis of the protective effects of bicycle helmets during impact and crush tests in

pediatric skull models. J Neurosurg Pediatr 10:490-497.

Matuseviciene G, Borg J, Stalnacke BM, Ulfarsson T, de Boussard C (2013) Early

intervention for patients at risk for persisting disability after mild traumatic brain injury: a

randomized, controlled study. Brain Inj 27:318-324.

May S, Martin G, Wellington C (2011) Brain injury and dementia: Is there a connection?

J Neurology Neurosci 2:1-7.

292

McAllister TW (2011) Neurobiological consequences of traumatic brain injury. Dialogues

Clin Neurosci 13:287-300.

McGinn MJ, Kelley BJ, Akinyi L, Oli MW, Liu MC, Hayes RL, Wang KK, Povlishock JT

(2009) Biochemical, structural, and biomarker evidence for calpain-mediated

cytoskeletal change after diffuse brain injury uncomplicated by contusion. J Neuropathol

Exp Neurol 68:241-249.

McIntosh TK, Noble L, Andrews B, Faden AI (1987) Traumatic brain injury in the rat:

characterization of a midline fluid-percussion model. Cent Nerv Syst Trauma 4:119-134.

McIntosh TK, Smith DH, Meaney DF, Kotapka MJ, Gennarelli TA, Graham DI (1996)

Neuropathological sequelae of traumatic brain injury: relationship to neurochemical and

biomechanical mechanisms. Lab Invest 74:315-342.

McIntosh TK, Vink R, Noble L, Yamakami I, Fernyak S, Soares H, Faden AL (1989)

Traumatic brain injury in the rat: characterization of a lateral fluid-percussion model.

Neuroscience 28:233-244.

McKee AC, Cantu RC, Nowinski CJ, Hedley-Whyte ET, Gavett BE, Busdon AE, Santini

VE, Lee HS, Kubilus CA, Stern RA (2009) Chronic traumatic encephalopathy in

athletes: Progressive tauopathy following repetitive head injury. J Neuropathol Exp

Neurol 68:709-735.

293

McKee AC, Gavett BE, Stern RA, Nowinski CJ, Cantu RC, Kowall NW, Perl DP, Hedley-

Whyte ET, Price B, Sullivan C, Morin P, Lee HS, Kubilus CA, Daneshvar DH, Wulff M,

Budson AE (2010) TDP-43 proteinopathy and motor neuron disease in chronic

traumatic encephalopathy. J Neuropathol Exp Neurol 69:918-929.

McKee AC et al. (2013) The spectrum of disease in chronic traumatic encephalopathy.

Brain 136:43-64.

McKenzie KJ, McLellan DR, Gentleman SM, Maxwell WL, Gennarelli TA, Graham DI

(1996) Is beta-APP a marker of axonal damage in short-surviving head injury? Acta

Neuropathol 92:608-613.

McKissock W, Taylor JC, Bloom WH, Till K (1960) Extradural haematoma. Observations

on 125 cases. Lancet 2:167–172.

McLean AJ (1995) Brain injury without head impact? J Neurotrauma 12:621-625.

McLean AJ, Anderson RWG (1997) Biomechanics of closed head injury. In: Head Injury

- Pathophysiology and Management of Severe Closed Head Injury (Reilly P, Bullock R,

eds), pp 25-37. London: Chapman and Hall.

Meaney DF, Smith DH (2011) Biomechanics of concussion. Clin Sports Med 30:19-31,

vii.

294

Meaney S, Hassan M, Sakinis A, Lutjohann D, von Bergmann K, Wennmalm U,

Diczfalusy U, Bjorkhem I (2001) Evidence that the major oxysterols in human circulation

originate from distinct pools of cholesterol: a stable isotope study. J Lipid Res 42:70-78.

Meehan WP, 3rd, Zhang J, Mannix R, Whalen MJ (2012) Increasing recovery time

between injuries improves cognitive outcome after repetitive mild concussive brain

injuries in mice. Neurosurgery 71:885-891.

Menke JG, Macnaul KL, Hayes NS, Baffic J, Chao YS, Elbrecht A, Kelly LJ, Lam MH,

Schmidt A, Sahoo S, Wang J, Wright SD, Xin P, Zhou G, Moller DE, Sparrow CP (2002)

A novel liver X receptor agonist establishes species differences in the regulation of

cholesterol 7alpha-hydroxylase (CYP7a). Endocrinology 143:2548-2558.

Menon DK, Schwab K, Wright DW, Maas AI (2010) Position statement: definition of

traumatic brain injury. Arch Phys Med Rehabil 91:1637-1640.

Miller BC, Eckman EA, Sambamurti K, Dobbs N, Chow KM, Eckman CB, Hersh LB,

Thiele DL (2003) Amyloid-β peptide levels in brain are inversely correlated with insulysin

activity levels in vivo. Proc Natl Acad Sci U S A 100:6221-6226.

Millspaugh JA (1937) Dementia pugilistica. US Naval Med Bull 35:297-303.

295

Miners JS, Van Helmond Z, Chalmers K, Wilcock G, Love S, Kehoe PG (2006)

Decreased expression and activity of neprilysin in Alzheimer disease are associated

with cerebral amyloid angiopathy. J Neuropathol Exp Neurol 65:1012-1021.

Mitro N, Vargas L, Romeo R, Koder A, Saez E (2007) T0901317 is a potent PXR ligand:

Implications for the biology ascribed to LXR. FEBS Lett 581:1721-1726.

Moein H, Khalili HA, Keramatian K (2006) Effect of methylphenidate on ICU and

hospital length of stay in patients with severe and moderate traumatic brain injury. Clin

Neurol Neurosurg 108:539-542.

Moore KJ, Rayner KJ, Suarez Y, Fernandez-Hernando C (2010) microRNAs and

cholesterol metabolism. Trends Endocrinol Metab 21:699-706.

Moran LM, Taylor HG, Ganesalingam K, Gastier-Foster JM, Frick J, Bangert B, Dietrich

A, Nuss KE, Rusin J, Wright M, Yeates KO (2009) Apolipoprotein E4 as a predictor of

outcomes in pediatric mild traumatic brain injury. J Neurotrauma 26:1489-1495.

Morrow JA, Hatters DM, Lu B, Höchtl P, Oberg KA, Rupp B, Weisgraber KH (2002)

Apolipoprotein E4 forms a molten globule: a potential basis for its association with

disease. J Biol Chem 277:50380-50385.

Mortimer JA, French LR, Hutton JT, Schuman LM (1985) Head injury as a risk factor for

Alzheimer's disease. Neurology 35:264-267.

296

Mortimer JA, van Duijn CM, Chandra V, Fratiglioni L, Graves AB, Heyman A, Jorm AF,

Kokmen E, Kondo K, Rocca WA (1991) Head trauma as a risk factor for Alzheimer's

disease: a collaborative re-analysis of case-control studies. EURODEM Risk Factors

Research Group. Int J Epidemiol 20 Suppl 2:S28-S35.

Mouzon B, Chaytow H, Crynen G, Bachmeier C, Stewart J, Mullan M, Stewart W,

Crawford F (2012) Repetitive mild traumatic brain injury in a mouse model produces

learning and memory deficits accompanied by histological changes. J Neurotrauma

29:2761-2773.

Muizelaar JP, Marmarou A, Ward JD, Kontos HA, Choi SC, Becker DP, Gruemer H,

Young HF (1991) Adverse effects of prolonged hyperventilation in patients with severe

head injury: a randomized clinical trial. J Neurosurg 75:731-739.

Murthy S, Born E, Mathur SN, Field FJ (2002) LXR/RXR activation enhances

basolateral efflux of cholesterol in CaCo-2 cells. J Lipid Res 43:1054-1064.

Myburgh JA, Cooper DJ, Finfer SR, Venkatesh B, Jones D, Higgins A, Bishop N, Higlett

T, Australian Traumatic Brain Injury Study Investigators for the Australian:New Zealand

Intensive Care Society Clinical Trials G (2008) Epidemiology and 12-month outcomes

from traumatic brain injury in Australia and New Zealand. J Trauma 64:854-862.

297

Nagao K, Takahashi K, Azuma Y, Takada M, Kimura Y, Matsuo M, Kioka N, Ueda K

(2012) ATP hydrolysis-dependent conformational changes in the extracellular domain of

ABCA1 are associated with apoA-I binding. J Lipid Res 53:126-136.

Nagata KO, Nakada C, Kasai RS, Kusumi A, Ueda K (2013) ABCA1 dimer–monomer

interconversion during HDL generation revealed by single-molecule imaging. Proc Natl

Acad Sci U S A 110:5034-5039.

Nakagawa H, Ueno M, Itokazu T, Yamashita T (2013) Bilateral movement training

promotes axonal remodeling of the corticospinal tract and recovery of motor function

following traumatic brain injury in mice. Cell Death Dis 4:e534.

Nakamura K, Kennedy MA, Bald n A, Bohanic DD, Lyons K, Edwards PA (2004)

Expression and regulation of multiple murine ATP-binding cassette transporter G1

mRNAs/isoforms that stimulate cellular cholesterol efflux to high density lipoprotein. J

Biol Chem 279 45980-45989.

Namba Y, Tomonaga M, Kawasaki H, Otomo E, Ikeda K (1991) Apolipoprotein E

immunoreactivity in cerebral amyloid deposits and neurofibrillary tangles in Alzheimer's

disease and kuru plaque amyloid in Creutzfeldt-Jakob disease. Brain Res 541:163-166.

Namjoshi DR, Good C, Cheng WH, Panenka W, Richards D, Cripton PA, Wellington CL

(2013a) Towards clinical management of traumatic brain injury: a review of models and

mechanisms from a biomechanical perspective. Dis Model Mech 6:1325-1338.

298

Namjoshi DR, Martin G, Donkin J, Wilkinson A, Stukas S, Fan J, Carr M, Tabarestani S,

Wuerth K, Hancock RE, Wellington CL (2013b) The liver X receptor agonist GW3965

improves recovery from mild repetitive traumatic brain injury in mice partly through

apolipoprotein E. PLoS One 8:e53529.

Naslund J, Jensen M, Tjernberg LO, Thyberg J, Terenius L, Nordstedt C (1994) The

metabolic pathway generating p3, an A beta-peptide fragment, is probably non-

amyloidogenic. Biochem Biophys Res Commun 204:780-787.

Nathan BP, Bellosta S, Sanan DA, Weisgraber KH, Mahley RW, Pitas RE (1994)

Differential effects of apolipoproteins E3 and E4 on neuronal growth in vitro. Science

264:850-852.

Nathan BP, Jiang Y, Wong GK, Shen F, Brewer GJ, Struble RG (2002) Apolipoprotein

E4 inhibits, and apolipoprotein E3 promotes neurite outgrowth in cultured adult mouse

cortical neurons through the low-density lipoprotein receptor-related protein. Brain Res

928:96-105.

Nathan BP, Chang KC, Bellosta S, Brisch E, Ge N, Mahley RW, Pitas RE (1995) The

inhibitory effect of apolipoprotein E4 on neurite outgrowth is associated with microtubule

depolymerization. J Biol Chem 270:19791-19799.

Neubuerger KT, Sinton DW, Denst J (1959) Cerebral atrophy associated with boxing.

AMA Arch Neurol Psychiatry 81:403-408.

299

Neve RL, Harris P, Kosik KS, Kurnit DM, Donlon TA (1986) Identification of cDNA

clones for the human microtubule-associated protein tau and chromosomal localization

of the genes for tau and microtubule-associated protein 2. Mol Brain Res 1:271-280.

Newman JA (2002) Biomechanics of Head Trauma: Head Protection. In: Accidental

Injury: Biomechanics and Prevention (Nahum AM, Melvin JW, eds), pp 303–323. New

York, USA: Springer.

Novack TA, Banos JH, Brunner R, Renfroe S, Meythaler JM (2009) Impact of early

administration of sertraline on depressive symptoms in the first year after traumatic

brain injury. J Neurotrauma 26:1921-1928.

O'Neill RA et al. (2006) Isoelectric focusing technology quantifies protein signaling in 25

cells. Proc Natl Acad Sci U S A 103:16153-16158.

Oberkofler H, Schraml E, Krempler F, Patsch W (2003) Potentiation of liver X receptor

transcriptional activity by peroxisome-proliferator-activated receptor gamma co-activator

1 alpha. Biochem J 371:89-96.

Ogata M, Tsujita M, Hossain M, Akita N, Gonzales FJ, Staels B, Suzuki S, Fukutomi T,

Kimura G, Yokoyama S (2009) On the mechanism for PPAR agonists to enhance

ABCA1 expression. Atherosclerosis 205:413-419.

300

Ohkubo N, Lee YD, Morishima A, Terashima T, Kikkawa S, Tohyama M, Sakanaka M,

Tanaka J, Maeda N, Vitek MP, Mitsuda N (2003) Apolipoprotein E and reelin ligands

modulate tau phosphorylation through an apolipoprotein E receptor/disabled-1/glycogen

synthase kinase-3beta cascade. FASEB J 17:295-297.

Olsson A, Csajbok L, Ost M, Hoglund K, Nylen K, Rosengren L, Nellgard B, Blennow K

(2004) Marked increase of beta-amyloid(1-42) and amyloid precursor protein in

ventricular cerebrospinal fluid after severe traumatic brain injury. J Neurol 251:870-876.

Omalu B, Hammers JL, Bailes J, Hamilton RL, Kamboh MI, Webster G, Fitzsimmons

RP (2011) Chronic traumatic encephalopathy in an Iraqi war veteran with posttraumatic

stress disorder who committed suicide. Neurosurg Focus 31:E3.

Omalu BI, DeKosky ST, Hamilton RL, Minster RL, Kamboh MI, Shakir AM, Wecht CH

(2006) Chronic traumatic encephalopathy in a national football league player: part II.

Neurosurgery 59:1086-1093.

Ommaya AK, Boretos JW, Beile EE (1969) The Lexan Calvarium: An Improved Method

for Direct Observation of the Brain. Journal of Neurosurgery 30:25-29.

Ono K, Kikuchi A, Nakamura M, Kobayashi H, Nakamura N (1980) Human Head

Tolerance to Sagittal Impact Reliable Estimation Deduced from Experimental Head

Injury Using Subhuman Primates and Human Cadaver Skulls. SAE Technical Paper

801303.

301

Onyszchuk G, Al-Hafez B, He Y-Y, Bilgen M, Berman NEJ, Brooks WM (2007) A mouse

model of sensorimotor controlled cortical impact: Characterization using longitudinal

magnetic resonance imaging, behavioral assessments and histology. J Neurosci

Methods 160:187-196.

Oram JF, Vaughan AM (2000) ABCA1-mediated transport of cellular cholesterol and

phosopholipids to HDL apoplipoproteins. Curr Opin Lipidol 11:253-260.

Oram JF, Lawn RM (2001) ABCA1. The gatekeeper for eliminating excess tissue

cholesterol. J Lipid Res 42:1173-1179.

Oram JF, Lawn RM, Garvin MR, Wade DP (2000) ABCA1 is the cAMP-inducible

apolipoprotein receptor that mediates cholesterol secretion from macrophages. J Biol

Chem 275:34508-34511.

Ou J, Tu H, Shan B, Luk A, DeBose-Boyd RA, Bashmakov Y, Goldstein JL, Brown MS

(2001) Unsaturated fatty acids inhibit transcription of the sterol regulatory element-

binding protein-1c (SREBP-1c) gene by antagonizing ligand-dependent activation of the

LXR. Proc Natl Acad Sci U S A 98:6027-6032.

Pan W, Kastin AJ, Rigai T, McLay R, Pick CG (2003) Increased hippocampal uptake of

tumor necrosis factor alpha and behavioral changes in mice. Exp Brain Res 149:195-

199.

302

Park E, Bell JD, Baker AJ (2008) Traumatic brain injury: can the consequences be

stopped? CMAJ 178:1163-1170.

Patrick LM, Lissner HR, Gurdjian ES (1963) Survival by design: Head protection. In:

Proc. 7th STAPP Car Crash Conference, pp 483-499. Warren dale, PA: Association for

the Advancement of Automotive Medicine.

Paxinos G, Franklin KBJ (2001) The Mouse Brain In Stereotaxic Coordinates. San

Diego: Elsevier/Academic Press.

Pei JJ, Tanaka T, Tung YC, Braak E, Iqbal K, Grundke-Iqbal I (1997) Distribution,

levels, and activity of glycogen synthase kinase-3 in the Alzheimer disease brain. J

Neuropathol Exp Neurol 56:70-78.

Pellman EJ, Viano DC, Tucker AM, Casson IR (2003a) Concussion in professional

football: location and direction of helmet impacts-Part 2. Neurosurgery 53:1328-1340;

discussion 1340-1321.

Pellman EJ, Viano DC, Tucker AM, Casson IR, Waeckerle JF (2003b) Concussion in

professional football: reconstruction of game impacts and injuries. Neurosurgery

53:799-812; discussion 812-794.

303

Peng Y, Yang J, Deck C, Otte D, Willinger R (2013) Development of head injury risk

functions based on real-world accident reconstruction. Int J Crashworthiness 19:105-

114.

Perez-Polo JR, Rea HC, Johnson KM, Parsley MA, Unabia GC, Xu G, Infante SK,

Dewitt DS, Hulsebosch CE (2013) Inflammatory consequences in a rodent model of

mild traumatic brain injury. J Neurotrauma 30:727-740.

Perez A, Morelli L, Cresto JC, Castano EM (2000) Degradation of soluble amyloid beta-

peptides 1-40, 1-42, and the Dutch variant 1-40Q by insulin degrading enzyme from

Alzheimer disease and control brains. Neurochem Res 25:247-255.

Petit-Turcotte C, Aumont N, Blain JF, Poirier J (2007) Apolipoprotein E receptors and

amyloid expression are modulated in an apolipoprotein E-dependent fashion in

response to hippocampal deafferentation in rodent. Neuroscience 150:58-63.

Petrlova J, Hong HS, Bricarello DA, Harishchandra G, Lorigan GA, Jin LW, Voss JC

(2011) A differential association of Apolipoprotein E isoforms with the amyloid-beta

oligomer in solution. Proteins 79:402-416.

Pierce JE, Trojanowski JQ, Graham DI, Smith DH, McIntosh TK (1996)

Immunohistochemical characterization of alterations in the distribution of amyloid

precursor proteins and beta-amyloid peptide after experimental brain injury in the rat. J

Neurosci 16:1083-1090.

304

Pinteaux E, Parker LC, Rothwell NJ, Luheshi GN (2002) Expression of interleukin-1

receptors and their role in interleukin-1 actions in murine microglial cells. J Neurochem

83:754-763.

Pittella JE, Gusmao SN (2003) Diffuse vascular injury in fatal road traffic accident

victims: its relationship to diffuse axonal injury. J Forensic Sci 48:626-630.

Plassman BL, Havlik RJ, Steffens DC, Helms MJ, Newman TN, Drosdick D, Phillips C,

Gau BA, Welsh-Bohmer KA, Burke JR, Guralnik JM, Breitner JC (2000) Documented

head injury in early adulthood and risk of Alzheimer's disease and other dementias.

Neurology 55:1158-1166.

Plotz EJ, Kabara JJ, Davis ME, LeRoy GV, Gould RG (1968) Studies on the synthesis

of cholesterol in the brain of the human fetus. Am J Obstet Gynecol 101:534-538.

Poirier J (1994) Apolipoprotein E in animal models of CNS injury and Alzheimer's

disease. Trends Neurosci 17:525-530.

Poirier J, Hess M, May PC, Finch CE (1991) Astrocytic apolipoprotein E mRNA and

GFAP mRNA in hippocampus after entorhinal cortex lesioning. Brain Res Mol Brain Res

11:97-106.

305

Poirier J, Baccichet A, Dea D, Gauthier S (1993a) Cholesterol synthesis and lipoprotein

reuptake during synaptic remodelling in hippocampus in adult rats. Neuroscience 55:81-

90.

Poirier J, Davignon J, Bouthillier D, Kogan S, Bertrand P, Gauthier S (1993b)

Apolipoprotein E polymorphism and Alzheimer's disease. The Lancet 342:697-699.

Povlishock JT, Kontos HA (1985) Continuing axonal and vascular change following

experimental brain trauma. Cent Nerv Syst Trauma 2:285-298.

Povolny MA, Kaplan SP (1993) Traumatic brain injury occuring with spinal cord injury:

Significance for rehabilitation. J Rehabil:23-37.

Prange MT, Margulies SS (2002) Regional, directional, and age-dependent properties

of the brain undergoing large deformation. J Biomech Eng 124:244-252.

Prange MT, Meaney DF, Margulies SS (2000) Defining brain mechanical properties:

effects of region, direction, and species. Stapp Car Crash J 44:205-213.

Prasad P, Mertz H (1985) The Position of the United States Delegation to the ISO

Working Group 6 on the Use of HIC in the Automotive Environment. SAE Technical

Paper 851246.

306

Prevention CfDCa (2007) Nonfatal traumatic brain injuries from sports and recreation

activities-United States, 2001-2005. MMWR Morb Mortal Wkly Rep 56:733-737.

Pruthi N, Chandramouli BA, Kuttappa TB, Rao SL, Subbakrishna DK, Abraham MP,

Mahadevan A, Shankar SK (2010) Apolipoprotein E polymorphism and outcome after

mild to moderate traumatic brain injury: a study of patient population in India. Neurol

India 58:264-269.

Pudenz RH, Shelden CH (1946) The lucite calvarium; a method for direct observation of

the brain; cranial trauma and brain movement. J Neurosurg 3:487-505.

Putukian M (2011) The acute symptoms of sport-related concussion: diagnosis and on-

field management. Clin Sports Med 30:49-61, viii.

Puvanachandra P, Hyder AA (2008) Traumatic brain injury in Latin America and the

Caribbean: a call for research. Salud Pública de México 50:s3-s5.

Qiu W, Zhang Y, Sheng H, Zhang J, Wang W, Liu W, Chen K, Zhou J, Xu Z (2007)

Effects of therapeutic mild hypothermia on patients with severe traumatic brain injury

after craniotomy. J Crit Care 22:229-235.

Qiu WQ, Walsh DM, Ye Z, Vekrellis K, Zhang J, Podlisny MB, Rosner MR, Safavi A,

Hersh LB, Selkoe DJ (1998) Insulin-degrading enzyme regulates extracellular levels of

amyloid beta-protein by degradation. J Biol Chem 273:32730-32738.

307

Qiu WS, Liu WG, Shen H, Wang WM, Hang ZL, Zhang Y, Jiang SJ, Yang XF (2005)

Therapeutic effect of mild hypothermia on severe traumatic head injury. Chin J

Traumatol 8:27-32.

Quinet EM, Savio DA, Halpern AR, Chen L, Schuster GU, Gustafsson JA, Basso MD,

Nambi P (2006) Liver X receptor (LXR)-beta regulation in LXRalpha-deficient mice:

implications for therapeutic targeting. Mol Pharmacol 70:1340-1349.

Raby CA, Morganti-Kossmann MC, Kossmann T, Stahel PF, Watson MD, Evans LM,

Mehta PD, Spiegel K, Kuo YM, Roher AE, Emmerling MR (1998) Traumatic brain injury

increases beta-amyloid peptide 1-42 in cerebrospinal fluid. J Neurochem 71:2505-2509.

Rader DJ, deGoma EM (2012) Approach to the Patient with Extremely Low HDL-

Cholesterol. J Clin Endocrinol Metab 97:3399-3407.

Rafaels KA, Bass CR, Panzer MB, Salzar RS, Woods WA, Feldman SH, Walilko T,

Kent RW, Capehart BP, Foster JB, Derkunt B, Toman A (2012) Brain injury risk from

primary blast. J Trauma Acute Care Surg 73:895-901.

Raffai RL, Dong LM, Farese RV, Jr., Weisgraber KH (2001) Introduction of human

apolipoprotein E4 "domain interaction" into mouse apolipoprotein E. Proc Natl Acad Sci

U S A 98:11587-11591.

308

Raghupathi R (2004) Cell death mechanisms following traumatic brain injury. Brain

Pathol 14:215-222.

Ramlackhansingh AF, Brooks DJ, Greenwood RJ, Bose SK, Turkheimer FE, Kinnunen

KM, Gentleman S, Heckemann RA, Gunanayagam K, Gelosa G, Sharp DJ (2011)

Inflammation after trauma: microglial activation and traumatic brain injury. Ann Neurol

70:374-383.

Rasmusson DX, Brandt J, Martin DB, Folstein MF (1995) Head injury as a risk factor in

Alzheimer's disease. Brain Inj 9:213-219.

Rayner KJ, Suarez Y, Davalos A, Parathath S, Fitzgerald ML, Tamehiro N, Fisher EA,

Moore KJ, Fernandez-Hernando C (2010) MiR-33 contributes to the regulation of

cholesterol homeostasis. Science 328:1570-1573.

Reilly P (2007) The impact of neurotrauma on society: an international perspective.

Prog Brain Res 161:3-9.

Ren Z, Iliff JJ, Yang L, Yang J, Chen X, Chen MJ, Giese RN, Wang B, Shi X,

Nedergaard M (2013) 'Hit & Run' model of closed-skull traumatic brain injury (TBI)

reveals complex patterns of post-traumatic AQP4 dysregulation. J Cereb Blood Flow

Metab 33:834-845.

309

Repa JJ, Berge KE, Pomajzl C, Richardson JA, Hobbs H, Mangelsdorf DJ (2002)

Regulation of ATP-binding cassette sterol transporters ABCG5 and ABCG8 by the liver

X receptors alpha and beta. J Biol Chem 277:18793-18800.

Riddell DR et al. (2007) The LXR agonist TO901317 selectively lowers hippocampal

Abeta42 and improves memory in the Tg2576 mouse model of Alzheimer's disease. Mol

Cell Neurosci 34:621-628.

Rigamonti E, Helin L, Lestavel S, Mutka AL, Lepore M, Fontaine C, Bouhlel MA, Bultel

S, Fruchart JC, Ikonen E, Clavey V, Staels B, Chinetti-Gbaguidi G (2005) Liver X

receptor activation controls intracellular cholesterol trafficking and esterification in

human macrophages. Circ Res 97 682-689.

Roberts GW, Gentleman SM, Lynch A, Graham DI (1991) beta A4 amyloid protein

deposition in brain after head trauma. The Lancet 338:1422-1423.

Roberts GW, Gentleman SM, Lynch A, Murray L, Landon M, Graham DI (1994) Beta

amyloid protein deposition in the brain after severe head injury: implications for the

pathogenesis of Alzheimer's disease. J Neurol Neurosurg Psychiatry 57:419-425.

Rogaev EI et al. (1995) Familial Alzheimer's disease in kindreds with missense

mutations in a gene on chromosome 1 related to the Alzheimer's disease type 3 gene.

Nature 376:775-778.

310

Rogers J, Strohmeyer R, Kovelowski CJ, Li R (2002) Microglia and inflammatory

mechanisms in the clearance of amyloid beta peptide. Glia 40:260-269.

Rooker S, Jander S, Van Reempts J, Stoll G, Jorens PG, Borgers M, Verlooy J (2006)

Spatiotemporal pattern of neuroinflammation after impact-acceleration closed head

injury in the rat. Mediators Inflamm 2006:90123.

Roozenbeek B, Maas AI, Menon DK (2013) Changing patterns in the epidemiology of

traumatic brain injury. Nat Rev Neurol 9:231-236.

Roselaar SE, Daugherty A (1998) Apolipoprotein E-deficient mice have impaired innate

immune responses to Listeria monocytogenes in vivo. J Lipid Res 39:1740-1743.

Rothwell N (2003) Interleukin-1 and neuronal injury: mechanisms, modification, and

therapeutic potential. Brain Behav Immun 17:152-157.

Rowson S, Duma SM (2011) Development of the STAR evaluation system for football

helmets: integrating player head impact exposure and risk of concussion. Ann Biomed

Eng 39:2130-2140.

Rowson S, Duma SM (2013) Brain Injury Prediction: Assessing the Combined

Probability of Concussion Using Linear and Rotational Head Acceleration. Ann Biomed

Eng.

311

Rowson S, Duma SM, Beckwith JG, Chu JJ, Greenwald RM, Crisco JJ, Brolinson PG,

Duhaime AC, McAllister TW, Maerlender AC (2012) Rotational head kinematics in

football impacts: an injury risk function for concussion. Ann Biomed Eng 40:1-13.

Rubovitch V, Ten-Bosch M, Zohar O, Harrison CR, Tempel-Brami C, Stein E, Hoffer BJ,

Balaban CD, Schreiber S, Chiu WT, Pick CG (2011) A mouse model of blast-induced

mild traumatic brain injury. Exp Neurol 232:280-289.

Rudelli R, Strom JO, Welch PT, Ambler MW (1982) Posttraumatic premature

Alzheimer's disease. Neuropathologic findings and pathogenetic considerations. Arch

Neurol 39:570-575.

Runnerstam M, Bao F, Huang Y, Shi J, Gutierrez E, Hamberger A, Hansson HA, Viano

D, Haglid K (2001) A new model for diffuse brain injury by rotational acceleration: II.

Effects on extracellular glutamate, intracranial pressure, and neuronal apoptosis. J

Neurotrauma 18:259-273.

Rusnak M (2013) Traumatic brain injury: Giving voice to a silent epidemic. Nat Rev

Neurol 9:186-187.

Russel WR (1932) Cerebral involvement in head injury: a study based on the

examination of two hundred cases. Brain 55:549-603.

312

Russo R, Borghi R, Markesbery W, Tabaton M, Piccini A (2005) Neprylisin decreases

uniformly in Alzheimer’s disease and in normal aging. FEBS Letters 579:6027-6030.

Rutherford GW, Bazarian JJ, Cernak I, Corrigan JD, Dikmen SS, Grady SM, Hesdorffer

DC, Kraus JF, Levin HS, Noble L, Potolicchio SJ, Rauch SL, Stiers W, Tamminga CA,

Temkin N, Weisskopf MG (2008) Gulf War and Health: Long-Term Consequences of

Traumatic Brain Injury: The National Academies Press.

Saatman KE, Feeko KJ, Pape RL, Raghupathi R (2006) Differential behavioral and

histopathological responses to graded cortical impact injury in mice. J Neurotrauma

23:1241-1253.

Saatman KE, Duhaime AC, Bullock R, Maas AI, Valadka A, Manley GT (2008)

Classification of traumatic brain injury for targeted therapies. J Neurotrauma 25:719-

738.

Sabet AA, Christoforou E, Zatlin B, Genin GM, Bayly PV (2008) Deformation of the

human brain induced by mild angular head acceleration. J Biomech 41:307-315.

Saido T, Leissring MA (2012) Proteolytic degradation of amyloid beta-protein. Cold

Spring Harb Perspect Med 2:a006379.

Saunders AM, Strittmatter WJ, Schmechel D, George-Hyslop PH, Pericak-Vance MA,

Joo SH, Rosi BL, Gusella JF, Crapper-MacLachlan DR, Alberts MJ (1993) Association

313

of apolipoprotein E allele epsilon 4 with late-onset familial and sporadic Alzheimer's

disease. Neurology 43:1467-1472.

Scalfani MT, Dhar R, Zazulia AR, Videen TO, Diringer MN (2012) Effect of osmotic

agents on regional cerebral blood flow in traumatic brain injury. J Crit Care

27:526.e527-512.

Scheuner D et al. (1996) Secreted amyloid beta-protein similar to that in the senile

plaques of Alzheimer's disease is increased in vivo by the presenilin 1 and 2 and APP

mutations linked to familial Alzheimer's disease. Nat Med 2:864-870.

Schley D, Carare-Nnadi R, Please CP, Perry VH, Weller RO (2006) Mechanisms to

explain the reverse perivascular transport of solutes out of the brain. J Theor Biol

238:962-974.

Schmidt ML, Zhukareva V, Newell KL, Lee VM, Trojanowski JQ (2001) Tau isoform

profile and phosphorylation state in dementia pugilistica recapitulate Alzheimer's

disease. Acta Neuropathol 101:518-524.

Schmitz G, Langmann T (2001) Structure, function and regulation of the ABC1 gene

product. Curr Opin Lipidol 12:129-140.

Schmitz G, Langmann T (2005) Transcriptional regulatory networks in lipid metabolism

control ABCA1 expression. Biochim Biophys Acta 1735:1-19.

314

Schroepfer GJ, Jr. (2000) Oxysterols: modulators of cholesterol metabolism and other

processes. Physiol Rev 80:361-554.

Schultz JR, Tu H, Luk A, Repa JJ, Medina JC, Li L, Schwendner S, Wang S, Thoolen

M, Mangelsdorf DJ, Lustig KD, Shan B (2000) Role of LXRs in control of lipogenesis.

Genes & Development 14:2831-2838.

Schwartz K, Lawn RM, Wade DP (2000) ABC1 gene expression and ApoA-1-mediated

cholesterol efflux are regulated by LXR. Biochem Biophys Res Commun 274:794-802.

Schwetye KE, Cirrito JR, Esparza TJ, Mac Donald CL, Holtzman DM, Brody DL (2010)

Traumatic brain injury reduces soluble extracellular amyloid-β in mice: A

methodologically novel combined microdialysis-controlled cortical impact study.

Neurobiol Disease 40:555-564.

Seelig JM, Becker DP, Miller JD, Greenberg RP, Ward JD, Choi SC (1981) Traumatic

acute subdural hematoma: major mortality reduction in comatose patients treated within

four hours. N Engl J Med 304:1511-1518.

Seitz A, Kragol M, Aglow E, Showe L, Heber-Katz E (2003) Apolipoprotein E expression

after spinal cord injury in the mouse. J Neurosci Res.

Selkoe DJ (1994) Alzheimer's disease: a central role for amyloid. J Neuropathol Exp

Neurol 53:438-447.

315

Selkoe DJ, Wolfe MS (2007) Presenilin: Running with Scissors in the Membrane. Cell

131:215-221.

Servadei F, Murray GD, Teasdale GM, Dearden M, Iannotti F, Lapierre F, Maas AJ,

Karimi A, Ohman J, Persson L, Stocchetti N, Trojanowski T, Unterberg A (2002)

Traumatic subarachnoid hemorrhage: demographic and clinical study of 750 patients

from the European brain injury consortium survey of head injuries. Neurosurgery

50:261-267; discussion 267-269.

Sevagan G, Zhu F, Jiang B, Yang KH (2013) Numerical simulations of the occupant

head response in an infantry vehicle under blunt impact and blast loading conditions.

Proc Inst Mech Eng H 227:778-787.

Sganzerla EP, Tomei G, Rampini P, Ceretti L, Gaini SM, Giovanelli M, Villani R (1989)

A peculiar intracerebral hemorrhage: The gliding contusion, it s relationship to diffuse

brain damage. Neurosurgical Review 12:215-218.

Shatsky SA, Alter WA, Evans DE, Armbrustmacher VW, Clark G, Earle KM (1974)

Traumatic distortions of the primate head and chest: correlation of biomechanical,

radiological and pathological data. Proceedings: Stapp Car Crash Conference 18:-.

Shein NA, Doron H, Horowitz M, Trembovler V, Alexandrovich AG, Shohami E (2007)

Altered cytokine expression and sustained hypothermia following traumatic brain injury

in heat acclimated mice. Brain Res 1185:313-320.

316

Shelden CH, Pudenz RH, Restarski JS, Craig WM (1944) The Lucite Calvarium—A

Method for Direct Observation of the Brain. J Neurosurg 1:67-75.

Sherriff FE, Bridges LR, Sivaloganathan S (1994a) Early detection of axonal injury after

human head trauma using immunocytochemistry for beta-amyloid precursor protein.

Acta Neuropathol 87:55-62.

Sherriff FE, Bridges LR, Gentleman SM, Sivaloganathan S, Wilson S (1994b) Markers

of axonal injury in post mortem human brain. Acta Neuropathol (Berl) 88:433-439.

Shibata M, Yamada S, Kumar SR, Calero M, Bading J, Frangione B, Holtzman DM,

Miller CA, Strickland DK, Ghiso J, Zlokovic BV (2000) Clearance of Alzheimer's

amyloid-{beta}1-40 peptide from brain by LDL receptor-related protein-1 at the blood-

brain barrier. J Clin Invest 106:1489-1499.

Shiozaki T, Hayakata T, Taneda M, Nakajima Y, Hashiguchi N, Fujimi S, Nakamori Y,

Tanaka H, Shimazu T, Sugimoto H (2001) A multicenter prospective randomized

controlled trial of the efficacy of mild hypothermia for severely head injured patients with

low intracranial pressure. Mild Hypothermia Study Group in Japan. J Neurosurg 94:50-

54.

Shirotani K, Tsubuki S, Iwata N, Takaki Y, Harigaya W, Maruyama K, Kiryu-Seo S,

Kiyama H, Iwata H, Tomita T, Iwatsubo T, Saido TC (2001) Neprilysin degrades both

317

amyloid beta peptides 1-40 and 1-42 most rapidly and efficiently among thiorphan- and

phosphoramidon-sensitive endopeptidases. J Biol Chem 276:21895-21901.

Shitaka Y, Tran HT, Bennett RE, Sanchez L, Levy MA, Dikranian K, Brody DL (2011)

Repetitive closed-skull traumatic brain injury in mice causes persistent multifocal axonal

injury and microglial reactivity. J Neuropathol Exp Neurol 70:551-567.

Shohami E, Gallily R, Mechoulam R, Bass R, Ben-Hur T (1997) Cytokine production in

the brain following closed head injury: dexanabinol (HU-211) is a novel TNF-alpha

inhibitor and an effective neuroprotectant. J Neuroimmunol 72:169-177.

Simon P, Dupuis R, Costentin J (1994) Thigmotaxis as an index of anxiety in mice.

Influence of dopaminergic transmissions. Behav Brain Res 61:59-64.

Slunt HH, Thinakaran G, Von Koch C, Lo AC, Tanzi RE, Sisodia SS (1994) Expression

of a ubiquitous, cross-reactive homologue of the mouse beta-amyloid precursor protein

(APP). J Biol Chem 269:2637-2644.

Smith C, Graham DI, Murray LS, Stewart J, Nicoll JA (2006) Association of APOE e4

and cerebrovascular pathology in traumatic brain injury. J Neurol Neurosurg Psychiatry

77:363-366.

Smith DH, Meaney DF (2000) Axonal Damage in Traumatic Brain Injury. Neuroscientist

6:483-495.

318

Smith DH, Meaney DF, Shull WH (2003a) Diffuse axonal injury in head trauma. J Head

Trauma Rehabil 18:307-316.

Smith DH, Johnson VE, Stewart W (2013) Chronic neuropathologies of single and

repetitive TBI: substrates of dementia? Nat Rev Neurol 9:211-221.

Smith DH, Chen XH, Iwata A, Graham DI (2003b) Amyloid beta accumulation in axons

after traumatic brain injury in humans. J Neurosurg 98:1072-1077.

Smith DH, Wolf JA, Lusardi TA, Lee VM, Meaney DF (1999a) High tolerance and

delayed elastic response of cultured axons to dynamic stretch injury. J Neurosci

19:4263-4269.

Smith DH, Chen XH, Xu BN, McIntosh TK, Gennarelli TA, Meaney DF (1997)

Characterization of diffuse axonal pathology and selective hippocampal damage

following inertial brain trauma in the pig. J Neuropathol Exp Neurol 56:822-834.

Smith DH, Nonaka M, Miller R, Leoni M, Chen XH, Alsop D, Meaney DF (2000)

Immediate coma following inertial brain injury dependent on axonal damage in the

brainstem. J Neurosurg 93:315-322.

Smith DH, Soares HD, Pierce JS, Perlman KG, Saatman KE, Meaney DF, Dixon CE,

McIntosh TK (1995) A model of parasagittal controlled cortical impact in the mouse:

cognitive and histopathologic effects. J Neurotrauma 12:169-178.

319

Smith DH, Chen XH, Nonaka M, Trojanowski JQ, Lee VM, Saatman KE, Leoni MJ, Xu

BN, Wolf JA, Meaney DF (1999b) Accumulation of amyloid beta and tau and the

formation of neurofilament inclusions following diffuse brain injury in the pig. J

Neuropathol Exp Neurol 58:982-992.

Smith HP, Kelly DL, Jr., McWhorter JM, Armstrong D, Johnson R, Transou C, Howard

G (1986) Comparison of mannitol regimens in patients with severe head injury

undergoing intracranial monitoring. J Neurosurg 65:820-824.

Smith KR, Jr., Goulding PM, Wilderman D, Goldfader PR, Holterman-Hommes P, Wei F

(1994) Neurobehavioral effects of phenytoin and carbamazepine in patients recovering

from brain trauma: a comparative study. Arch Neurol 51:653-660.

Soloniuk D, Pitts LH, Lovely M, Bartkowski H (1986) Traumatic intracerebral

hematomas: timing of appearance and indications for operative removal. J Trauma

26:787-794.

Song C, Hiipakka RA, Liao S (2001) Auto-oxidized cholesterol sulfates are antagonistic

ligands of liver X receptors: implications for the development and treatment of

atherosclerosis. Steroids 66:473-479.

Song C, Kokontis JM, Hiipakka RA, Liao S (1994) Ubiquitous receptor: a receptor that

modulates gene activation by retinoic acid and thyroid hormone receptors. Proc Natl

Acad Sci U S A 91:10809-10813.

320

Sontag E, Hladik C, Montgomery L, Luangpirom A, Mudrak I, Ogris E, White CL, 3rd

(2004a) Downregulation of protein phosphatase 2A carboxyl methylation and

methyltransferase may contribute to Alzheimer disease pathogenesis. J Neuropathol

Exp Neurol 63:1080-1091.

Sontag E, Luangpirom A, Hladik C, Mudrak I, Ogris E, Speciale S, White CL, 3rd

(2004b) Altered expression levels of the protein phosphatase 2A ABalphaC enzyme are

associated with Alzheimer disease pathology. J Neuropathol Exp Neurol 63:287-301.

Sorbi S, Nacmias B, Piacentini S, Repice A, Latorraca S, Forleo P, Amaducci L (1995)

ApoE as a prognostic factor for post-traumatic coma. Nat Med 1:852.

Spillantini MG, Goedert M (2013) Tau pathology and neurodegeneration. Lancet Neurol

12:609-622.

Spillantini MG, Goedert M, Crowther RA, Murrell JR, Farlow MR, Ghetti B (1997)

Familial multiple system tauopathy with presenile dementia: a disease with abundant

neuronal and glial tau filaments. Proc Natl Acad Sci U S A 94:4113-4118.

Stalhammar D, Galinat BJ, Allen AM, Becker DP, Stonnington HH, Hayes RL (1987) A

new model of concussive brain injury in the cat produced by extradural fluid volume

loading: I. Biomechanical properties. Brain Inj 1:73-91.

321

Statler KD, Alexander H, Vagni V, Holubkov R, Dixon CE, Clark RS, Jenkins L,

Kochanek PM (2006) Isoflurane exerts neuroprotective actions at or near the time of

severe traumatic brain injury. Brain Res 1076:216-224.

Statler KD, Kochanek PM, Dixon CE, Alexander HL, Warner DS, Clark RS, Wisniewski

SR, Graham SH, Jenkins LW, Marion DW, Safar PJ (2000) Isoflurane improves long-

term neurologic outcome versus fentanyl after traumatic brain injury in rats. J

Neurotrauma 17:1179-1189.

Stern RA, Riley DO, Daneshvar DH, Nowinski CJ, Cantu RC, McKee AC (2011) Long-

term consequences of repetitive brain trauma: chronic traumatic encephalopathy.

PM&R 3:S460-S467.

Strich SJ (1956) Diffuse degeneration of the cerebral white matter in severe dementia

following head injury. J Neurol Neurosurg Psychiat 19:163-185.

Strittmatter WJ, Weisgraber KH, Huang DY, Dong LM, Salvesen GS, Pericak-Vance M,

Schmechel D, Saunders AM, Goldgaber D, Roses AD (1993) Binding of human

apolipoprotein E to synthetic amyloid beta peptide: isoform-specific effects and

implications for late-onset Alzheimer disease. Proc Natl Acad Sci U S A 90:8098-8102.

Stukas S, Robert J, Wellington CL (2014) High-density lipoproteins and cerebrovascular

integrity in Alzheimer's disease. Cell Metab 19:574-591.

322

Stukas S, May S, Wilkinson A, Chan J, Donkin J, Wellington CL (2011) The LXR

agonist GW3965 increases apoA-I protien levels in the central nervous system

independent of ABCA1. Biochim Biophys Acta doi:10.1016.

Suh YH, Checler F (2002) Amyloid precursor protein, presenilins, and alpha-synuclein:

molecular pathogenesis and pharmacological applications in Alzheimer's disease.

Pharmacol Rev 54:469-525.

Sun Y, Yao J, Kim T-W, Tall AR (2003) Expression of Liver X Receptor Target Genes

Decreases Cellular Amyloid β Peptide Secretion. J Biol Chem 278:27688-27694.

Surgeons AAoN (2011) Sports-related head injury. In.

Taber KH, Warden DL, Hurley RA (2006) Blast-related traumatic brain injury: what is

known? J Neuropsychiatry Clin Neurosci 18:141-145.

Tachikawa M, Watanabe M, Hori S, Fukaya M, Ohtsuki S, Asashima T, Terasaki T

(2005) Distinct spatio-temporal expression of ABCA and ABCG transporters in the

developing and adult mouse brain. J Neurochem 95 294-304.

Tagliaferri F, Compagnone C, Korsic M, Servadei F, Kraus J (2006) A systematic review

of brain injury epidemiology in Europe. Acta Neurochir (Wien) 148:255-268.

323

Tai LM, Bilousova T, Jungbauer L, Roeske SK, Youmans KL, Yu C, Poon WW,

Cornwell LB, Miller CA, Vinters HV, Van Eldik LJ, Fardo DW, Estus S, Bu G, Gylys KH,

Ladu MJ (2013) Levels of soluble apolipoprotein E/amyloid-beta (Abeta) complex are

reduced and oligomeric Abeta increased with APOE4 and Alzheimer disease in a

transgenic mouse model and human samples. J Biol Chem 288:5914-5926.

Takahashi H, Manaka S, Sano K (1981) Changes in extracellular potassium

concentration in cortex and brain stem during the acute phase of experimental closed

head injury. J Neurosurg 55:708-717.

Takahashi Y, Smith JD (1999) Cholesterol efflux to apolipoprotein A1 involves

endocytosis and resecretion in a calcium-dependent pathway. Proc Natl Acad Sci USA

96 11358-11363.

Takhounts EG, Crandall JR, Darvish K (2003) On the importance of nonlinearity of brain

tissue under large deformations. Stapp Car Crash J 47:79-92.

Tall AR (2008) Cholesterol efflux pathways and other potential mechanisms involved in

the athero-protective effect of high density lipoproteins. J Intern Med 263:256-273.

Tanaka AR, Ikeda Y, Abe-Dohmae S, Arakawa R, Sadanami K, Kidera A, Nakagawa S,

Nagase T, Aoki R, Kioka N, Amachi T, Yokoyama S, Ueda K (2001) Human ABCA1

Contains a Large Amino-Terminal Extracellular Domain Homologous to an Epitope of

Sjögren's Syndrome. Biochem Biophys Res Commun 283:1019-1025.

324

Taneda M, Kataoka K, Akai F, Asai T, Sakata I (1996) Traumatic subarachnoid

hemorrhage as a predictable indicator of delayed ischemic symptoms. J Neurosurg

84:762-768.

Tang-Schomer MD, Patel AR, Baas PW, Smith DH (2010) Mechanical breaking of

microtubules in axons during dynamic stretch injury underlies delayed elasticity,

microtubule disassembly, and axon degeneration. FASEB J 24:1401-1410.

Tang-Schomer MD, Johnson VE, Baas PW, Stewart W, Smith DH (2012) Partial

interruption of axonal transport due to microtubule breakage accounts for the formation

of periodic varicosities after traumatic axonal injury. Exp Neurol 233:364-372.

Tang YP, Noda Y, Hasegawa T, Nabeshima T (1997) A concussive-like brain injury

model in mice (I): impairment in learning and memory. J Neurotrauma 14:851-862.

Taupin V, Toulmond S, Serrano A, Benavides J, Zavala F (1993) Increase in IL-6, IL-1

and TNF levels in rat brain following traumatic lesion. Influence of pre- and post-

traumatic treatment with Ro5 4864, a peripheral-type (p site) benzodiazepine ligand. J

Neuroimmunol 42:177-185.

Teasdale G, Jennett B (1974) Assessment of coma and impaired consciousness: a

practical scale. The Lancet 304:81-84.

325

Teasdale GM, Nicoll JA, Murray G, Fiddes M (1997) Association of apolipoprotein E

polymorphism with outcome after head injury. The Lancet 350:1069-1071.

Tenovuo O, Alin J, Helenius H (2009) A randomized controlled trial of rivastigmine for

chronic sequels of traumatic brain injury-what it showed and taught? Brain Inj 23:548-

558.

Thibault LE, Meaney DF, Anderson BJ, Marmarou A (1992) Biomechanical aspects of a

fluid percussion model of brain injury. J Neurotrauma 9:311-322.

Thinakaran G, Koo EH (2008) Amyloid precursor protein trafficking, processing, and

function. J Biol Chem 283:29615-29619.

Thomas L (1997) Retrospective Power Analysis. Conserv Biol 11:276-280.

Thompson HJ, Lifshitz J, Marklund N, Grady MS, Graham DI, Hovda DA, McIntosh TK

(2005) Lateral fluid percussion brain injury: a 15-year review and evaluation. J

Neurotrauma 22:42-75.

Tokuda T, Calero M, Matsubara E, Vidal R, Kumar A, Permanne B, Zlokovic B, Smith

JD, Ladu MJ, Rostagno A, Ghiso J (2000) Lipidation of apolipoprotein E influences its

isoform-specific interaction with Alzheimer's amyloid beta peptides. Biochem J 348:359-

365.

326

Tolnay M, Sergeant N, Ghestem A, Chalbot S, De Vos RA, Jansen Steur EN, Probst A,

Delacourte A (2002) Argyrophilic grain disease and Alzheimer's disease are

distinguished by their different distribution of tau protein isoforms. Acta Neuropathol

104:425-434.

Tomlinson BE (1970) Brain-stem lesions after head injury. J Clin Pathol Suppl (R Coll

Pathol) 4:154-165.

Tran HT, LaFerla FM, Holtzman DM, Brody DL (2011a) Controlled cortical impact

traumatic brain injury in 3xTg-AD mice causes acute intra-axonal amyloid-á

accumulation and independently accelerates the development of tau abnormalities. J

Neurosci 31:9513-9525.

Tran HT, LaFerla FM, Holtzman DM, Brody DL (2011b) Controlled cortical impact

traumatic brain injury in 3xTg-AD mice causes acute intra-axonal amyloid-β

accumulation and independently accelerates the development of tau abnormalities. J

Neurosci 31:9513-9525.

Tran HT, Sanchez L, Esperza TJ, Brody DL (2011c) Distint temporal and anatomical

distributions of amyloid-β and tau abnormalities following controlled cortical impact in

transgenic mice. PLoS One 6:e2575.

327

Trommer BL, Shah C, Yun SH, Gamkrelidze G, Pasternak ES, Ye GL, Sotak M,

Sullivan PM, Pasternak JF, LaDu MJ (2004) ApoE isoform affects LTP in human

targeted replacement mice. Neuroreport 15:2655-2658.

Tsenter J, Beni-Adani L, Assaf Y, Alexandrovich AG, Trembovler V, Shohami E (2008)

Dynamic changes in the recovery after traumatic brain injury in mice: effect of injury

severity on T2-weighted MRI abnormalities, and motor and cognitive functions. J

Neurotrauma 25:324-333.

Tsoi LM, Wong KY, Liu YM, Ho YY (2007) Apoprotein E isoform-dependent expression

and secretion of pro-inflammatory cytokines TNF-alpha and IL-6 in macrophages. Arch

Biochem Biophys 460:33-40.

Ucar T, Tanriover G, Gurer I, Onal MZ, Kazan S (2006) Modified experimental mild

traumatic brain injury model. J Trauma 60:558-565.

Uchihara T, Shibuya K, Nakemura A, Yagishita S (2005) Silver stains distinguish tau-

positive structures in corticobasal degeneration/progressive supranuclear palso and in

Alzheimer's disease - comparison between Gallyas and Campbell-Switzer methods.

Acta Neuropathol 109:299-305.

United Sates Departments of Defense and Veterans Affairs (2008) Common definition

of TBI. In: U.S. Departments of Defense and Veterans Affairs.

328

Uryu K, Chen XH, Martinez D, Browne KD, Johnson VE, Graham DI, Lee VM,

Trojanowski JQ, Smith DH (2007) Multiple proteins implicated in neurodegenerative

diseases accumulate in axons after brain trauma in humans. Exp Neurol 208:185-192.

Valastro B, Ghribi O, Poirier J, Krzywkowski P, Massicotte G (2001) AMPA receptor

regulation and LTP in the hippocampus of young and aged apolipoprotein E-deficient

mice. Neurobiol Aging 22:9-15.

Van Den Heuvel C, Thornton E, Vink R (2007) Traumatic brain injury and Alzheimer's

Disease: a review. Prog Brain Res 161:297-310.

van Dommelen JAW, Hrapko M, Peters GWM (2009) Mechanical Properties of Brain

Tissue: Characterisation and Constitutive Modelling. In: Mechanosensitivity of the

Nervous System (Kamkim A, Kiseleva I, eds), pp 249-279: Springer Netherlands.

van Duijn CM, Tanja TA, Haaxma R, Schulte W, Saan RJ, Lameris AJ, Antonides-

Hendriks G, Hofman A (1992) Head trauma and the risk of Alzheimer's disease. Am J

Epidemiol 135:775-782.

Van Oosten M, Rensen PCN, Van Amersfoort ES, Van Eck M, Van Dam A-M, Brevé

JJP, Vogel T, Panet A, Van Berkel TJC, Kuiper J (2001) Apolipoprotein E protects

against bacterial lipopolysaccharide-induced lethality: a new therapeutic approach to

treat gram-negative sepsis. J Biol Chem 276:8820-8824.

329

Vance JE, Hayashi H, Karten B (2005) Cholesterol homeostasis in neurons and glial

cells. Semin Cell Dev Biol 16:193-212.

Vassar R (2004) BACE1: the beta-secretase enzyme in Alzheimer's disease. J Mol

Neurosci 23:105-114.

Vedhachalam C, Duong PT, Nickel M, Nguyen D, Dhanasekaran P, Saito H, Rothblat

GH, Lund-Katz S, Phillips MC (2007) Mechanism of ATP-binding cassette transporter

A1 (ABCA1)-mediated cellular lipid efflux to apolipoprotein A-I and formation of high

density lipoprotein particles. J Biol Chem 282:25123-25130.

Vekrellis K, Ye Z, Qiu WQ, Walsh D, Hartley D, Chesneau V, Rosner MR, Selkoe DJ

(2000) Neurons regulate extracellular levels of amyloid beta-protein via proteolysis by

insulin-degrading enzyme. J Neurosci 20:1657-1665.

Venkateswaran A, Repa JJ, Lobaccaro JM, Bronson A, Mangelsdorf DJ, Edwards PA

(2000) Human white/murine ABC8 mRNA levels are highly induced in lipid-loaded

macrophages. A transcriptional role for specific oxysterols. J Biol Chem 275:14700-

14707.

Venteclef N, Jakobsson T, Ehrlund A, Damdimopoulos A, Mikkonen L, Ellis E, Nilsson

L-M, Parini P, Jänne OA, Gustafsson J-Å, Steffensen KR, Treuter E (2010) GPS2-

dependent corepressor/SUMO pathways govern anti-inflammatory actions of LRH-1

and LXRβ in the hepatic acute phase response. Genes & Development 24:381-395.

330

Verghese PB, Castellano JM, Garai K, Wang Y, Jiang H, Shah A, Bu G, Frieden C,

Holtzman DM (2013) ApoE influences amyloid-beta (Abeta) clearance despite minimal

apoE/Abeta association in physiological conditions. Proc Natl Acad Sci U S A

110:E1807-1816.

Versace J National Highway Traffic Safety Adm. In: 15th Stapp Car Crash Conference,

SAE Paper #710881, pp 771-796. 1971.

Versace J (1971) A review of the severity index. In: Proc. of the 15th Stapp Car Crash

Conference, pp 771-796.

Viano DC, Hamberger A, Bolouri H, Saljo A (2009) Concussion in professional football:

animal model of brain injury--part 15. Neurosurgery 64:1162-1173; discussion 1173.

Viano DC, Hamberger A, Bolouri H, Saljo A (2012) Evaluation of three animal models

for concussion and serious brain injury. Ann Biomed Eng 40:213-226.

Viano DC, Casson IR, Pellman EJ, Bir CA, Zhang L, Sherman DC, Boitano MA (2005)

Concussion in professional football: comparison with boxing head impacts--part 10.

Neurosurgery 57:1154-1172; discussion 1154-1172.

Vos PE, Battistin L, Birbamer G, Gerstenbrand F, Potapov A, Prevec T, Stepan Ch A,

Traubner P, Twijnstra A, Vecsei L, von Wild K (2002) EFNS guideline on mild traumatic

brain injury: report of an EFNS task force. Eur J Neurol 9:207-219.

331

Wahrle S, Jiang H, Parsadanian M, Hartman RE, Bales KR, Paul SM, Holtzman DM

(2005) Deletion of Abca1 increases Abeta deposition in the PDAPP transgenic mouse

model of Alzheimer disease. J Biol Chem 280 43236-43242.

Wahrle SE, Jiang H, Parsadanian M, Legleiter J, Han X, Fryer JD, Kowalewski T,

Holtzman DM (2004) ABCA1 is required for normal CNS apoE levels and for lipidation

of astrocyte-secreted apoE. J Biol Chem 279 40987-40993.

Wahrle SE, Jiang H, Parsadanian M, Kim J, Li A, Knoten A, Jain S, Hirsch-Reinshagen

V, Wellington CL, Bales KR, Paul SM, Holtzman DM (2008) Overexpression of ABCA1

reduces amyloid deposition in the PDAPP mouse model of Alzheimer disease. J Clin

Invest 118:671-682.

Wang DS, Lipton RB, Katz MJ, Davies P, Buschke H, Kuslansky G, Verghese J,

Younkin SG, Eckman C, Dickson DW (2005) Decreased neprilysin immunoreactivity in

Alzheimer disease, but not in pathological aging. J Neuropathol Exp Neurol 64:378-385.

Wang H, Gao J, Lassiter T, McDonagh D, Sheng H, Warner D, Lynch J, Laskowitz D

(2006) Levetiracetam is neuroprotective in murine models of closed head injury and

subarachnoid hemorrhage. Neurocrit Care 5:71-78.

Wang HC, Duan ZX, Wu FF, Xie L, Zhang H, Ma YB (2010) A new rat model for diffuse

axonal injury using a combination of linear acceleration and angular acceleration. J

Neurotrauma 27:707-719.

332

Wang JZ, Grundke-Iqbal I, Iqbal K (2007) Kinases and phosphatases and tau sites

involved in Alzheimer neurofibrillary degeneration. Eur J Neurosci 25:59-68.

Wang L, Schuster GU, Hultenby K, Zhang Q, Andersson S, Gustafsson JA (2002) Liver

X receptors in the central nervous system: from lipid homeostasis to neuronal

degeneration. Proc Natl Acad Sci U S A 99:13878-13883.

Wang N, Tall AR (2003) Regulation and mechanisms of ATP-binding cassette

transporter A1-mediated cellular cholesterol efflux. Arterioscler Thromb Vasc Biol

23:1178-1184.

Wang N, Silver DL, Costet P, Tall AR (2000) Specific binding of ApoA-I, enhanced

cholesterol efflux, and altered plasma membrane morphology in cells expressing ABC1

[In Process Citation]. J Biol Chem 275:33053-33058.

Wang N, Chen W, Linsel-Nitschke P, Martinez LO, Agerholm-Larsen B, Silver DL, Tall

AR (2003) A PEST sequence in ABCA1 regulates degradation by calpain protease and

stabilization of ABCA1 by apoA-I. J Clin Invest 111:99-107.

Wang YJ, Shimura T, Kobayashi S, Teramoto A, Nakazawa S (1997) A lateral fluid

percussion model for the experimental severe brain injury and a morphological study in

the rats. Nihon Ika Daigaku Zasshi 64:172-175.

333

Weingarten MD, Lockwood AH, Hwo SY, Kirschner MW (1975) A protein factor

essential for microtubule assembly. Proc Natl Acad Sci U S A 72:1858-1862.

Weisgraber KH (1990) Apolipoprotein E distribution among human plasma lipoproteins:

role of the cysteine-arginine interchange at residue 112. J Lipid Res 31:1503-1511.

Weisgraber KH (1994) Apolipoprotein E: structure-function relationships. Adv Protein

Chem 45:249-302.

Weisgraber KH, Innerarity TL, Mahley RW (1982) Abnormal lipoprotein receptor-binding

activity of the human E apoprotein due to cysteine-arginine interchange at a single site.

J Biol Chem 257:2518-2521.

Weller RO, Yow HY, Preston SD, Mazanti I, Nicoll JA (2002) Cerebrovascular disease

is a major factor in the failure of elimination of Abeta from the aging human brain:

implications for therapy of Alzheimer's disease. Ann N Y Acad Sci 977:162-168.

Weller RO, Subash M, Preston SD, Mazanti I, Carare RO (2008) SYMPOSIUM:

Clearance of Aβ from the Brain in Alzheimer's Disease: Perivascular Drainage of

Amyloid-β Peptides from the Brain and Its Failure in Cerebral Amyloid Angiopathy and

Alzheimer's Disease. Brain Pathology 18:253-266.

Wellington CL, Walker EK, Suarez A, Kwok A, Bissada N, Singaraja R, Yang YZ, Zhang

LH, James E, Wilson JE, Francone O, McManus BM, Hayden MR (2002) ABCA1

334

mRNA and protein distribution patterns predict multiple different roles and levels of

regulation. Lab Invest 82:273-283.

Werner C, Engelhard K (2007) Pathophysiology of traumatic brain injury. Br J Anaesth

99:4-9.

Wetterau JR, Aggerbeck LP, Rall SC, Jr., Weisgraber KH (1988) Human apolipoprotein

E3 in aqueous solution. I. Evidence for two structural domains. J Biol Chem 263:6240-

6248.

White F, Nicoll JA, Horsburgh K (2001a) Alterations in apoE and apoJ in relation to

degeneration and regeneration in a mouse model of entorhinal cortex lesion. Exp

Neurol 169:307-318.

White F, Nicoll JA, Roses AD, Horsburgh K (2001b) Impaired neuronal plasticity in

transgenic mice expressing human apolipoprotein E4 compared to E3 in a model of

entorhinal cortex lesion. Neurobiol Dis 8:611-625.

Whitney KD, Watson MA, Collins JL, Benson WG, Stone TM, Numerick MJ, Tippin TK,

Wilson JG, Winegar DA, Kliewer SA (2002) Regulation of cholesterol homeostasis by

the liver X receptors in the central nervous system. Molec Endocrinol 16:1378-1385.

335

Whyte J, Hart T, Vaccaro M, Grieb-Neff P, Risser A, Polansky M, Coslett HB (2004)

Effects of methylphenidate on attention deficits after traumatic brain injury: a

multidimensional, randomized, controlled trial. Am J Phys Med Rehabil 83:401-420.

Wiebel FF, Gustafsson JA (1997) Heterodimeric interaction between retinoid X receptor

alpha and orphan nuclear receptor OR1 reveals dimerization-induced activation as a

novel mechanism of nuclear receptor activation. Mol Cell Biol 17:3977-3986.

Williams DJ, Tannenberg AE (1996) Dementia pugilistica in an alcoholic

achondroplastic dwarf. Pathology 28:102-104.

Willmott C, Ponsford J (2009) Efficacy of methylphenidate in the rehabilitation of

attention following traumatic brain injury: a randomised, crossover, double blind,

placebo controlled inpatient trial. J Neurol Neurosurg Psychiatry 80:552-557.

Wilson AC, Dugger BN, Dickson DW, Wang DS (2011) TDP-43 in aging and

Alzheimer's disease - a review. Int J Clin Exp Pathol 4:147-155.

Wilson C, Wardell MR, Weisgraber KH, Mahley RW, Agard DA (1991) Three-

dimensional structure of the LDL receptor-binding domain of human apolipoprotein E.

Science 252:1817-1822.

336

Winkler K, Scharnagl H, Tisljar U, Hoschutzky H, Friedrich I, Hoffmann MM, Huttinger

M, Wieland H, Marz W (1999) Competition of Abeta amyloid peptide and apolipoprotein

E for receptor-mediated endocytosis. J Lipid Res 40:447-455.

Winter CD, Iannotti F, Pringle AK, Trikkas C, Clough GF, Church MK (2002) A

microdialysis method for the recovery of IL-1beta, IL-6 and nerve growth factor from

human brain in vivo. J Neurosci Methods 119:45-50.

Wisniewski T, Frangione B (1992) Apolipoprotein E: A pathological chaperone protein in

patients with cerebral and systemic amyloid. Neurosci Lett 135:235-238.

Wojcicka G, Jamroz-Wisniewska A, Horoszewicz K, Beltowski J (2007) Liver X

receptors (LXRs). Part I: structure, function, regulation of activity, and role in lipid

metabolism. Postepy Hig Med Dosw (Online) 61:736-759.

Wolf G, Cifu D, Baugh L, Carne W, Profenna L (2012) The effect of hyperbaric oxygen

on symptoms after mild traumatic brain injury. J Neurotrauma 29:2606-2612.

Wolf JA, Stys PK, Lusardi T, Meaney D, Smith DH (2001) Traumatic axonal injury

induces calcium influx modulated by tetrodotoxin-sensitive sodium channels. J Neurosci

21:1923-1930.

Woodroofe MN, Sarna GS, Wadhwa M, Hayes GM, Loughlin AJ, Tinker A, Cuzner ML

(1991) Detection of interleukin-1 and interleukin-6 in adult rat brain, following

337

mechanical injury, by in vivo microdialysis: evidence of a role for microglia in cytokine

production. J Neuroimmunol 33:227-236.

Wright DW, Kellermann AL, Hertzberg VS, Clark PL, Frankel M, Goldstein FC,

Salomone JP, Dent LL, Harris OA, Ander DS, Lowery DW, Patel MM, Denson DD,

Gordon AB, Wald MM, Gupta S, Hoffman SW, Stein DG (2007) ProTECT: a

randomized clinical trial of progesterone for acute traumatic brain injury. Ann Emerg

Med 49:391-402.

Wright RM, Post A, Hoshizaki B, Ramesh KT (2013) A Multiscale Computational

Approach to Estimating Axonal Damage under Inertial Loading of the Head. J

Neurotrauma 30:102-118.

Wu X, Hu J, Zhuo L, Fu C, Hui G, Wang Y, Yang W, Teng L, Lu S, Xu G (2008)

Epidemiology of traumatic brain injury in eastern China, 2004: a prospective large case

study. J Trauma 64:1313-1319.

Xiao-Sheng H, Sheng-Yu Y, Xiang Z, Zhou F, Jian-ning Z (2000) Diffuse axonal injury

due to lateral head rotation in a rat model. J Neurosurg 93:626-633.

Xiao G, Wei J, Yan W, Wang W, Lu Z (2008) Improved outcomes from the

administration of progesterone for patients with acute severe traumatic brain injury: a

randomized controlled trial. Crit Care 12:R61.

338

Xiong Y, Mahmood A, Chopp M (2013) Animal models of traumatic brain injury. Nat Rev

Neurosci 14:128-142.

Xu PT, Gilbert JR, Qiu HL, Ervin J, Rothrock-Christian TR, Hulette C, Schmechel DE

(1999) Specific regional transcription of apolipoprotein E in human brain neurons. Am J

Pathol 154:601-611.

Xu Q, Bernardo A, Walker D, Kanegawa T, Mahley RW, Huang Y (2006) Profile and

regulation of apolipoprotein E (ApoE) expression in the CNS in mice with targeting of

green fluorescent protein gene to the ApoE locus. J Neurosci 26:4985-4994.

Yasojima K, McGeer EG, McGeer PL (2001a) Relationship between beta amyloid

peptide generating molecules and neprilysin in Alzheimer disease and normal brain.

Brain Res 919:115-121.

Yasojima K, Akiyama H, McGeer EG, McGeer PL (2001b) Reduced neprilysin in high

plaque areas of Alzheimer brain: a possible relationship to deficient degradation of β-

amyloid peptide. Neurosci Lett 297:97-100.

Yoganandan N, Gennarelli TA, Zhang J, Pintar FA, Takhounts E, Ridella SA (2009)

Association of contact loading in diffuse axonal injuries from motor vehicle crashes. J

Trauma 66:309-315.

339

Yokobori S, Bullock MR (2013) Pathobiology of primary traumatic brain injury. In: Brain

Injury Medicine: Principles and Practice, 2nd Edition (Zasler ND, Katz DI, Zafonte RD,

eds), pp 137-147. New York, NY: Demos Medical Publishing , LLC.

Yoshikawa T, Shimano H, Yahagi N, Ide T, Amemiya-Kudo M, Matsuzaka T, Nakakuki

M, Tomita S, Okazaki H, Tamura Y, Iizuka Y, Ohashi K, Takahashi A, Sone H, Osuga Ji

J, Gotoda T, Ishibashi S, Yamada N (2002) Polyunsaturated fatty acids suppress sterol

regulatory element-binding protein 1c promoter activity by inhibition of liver X receptor

(LXR) binding to LXR response elements. J Biol Chem 277:1705-1711.

Young B, Ott L, Kasarskis E, Rapp R, Moles K, Dempsey RJ, Tibbs PA, Kryscio R,

McClain C (1996) Zinc supplementation is associated with improved neurologic

recovery rate and visceral protein levels of patients with severe closed head injury. J

Neurotrauma 13:25-34.

Yu C, Youmans KL, LaDu MJ (2011) Proposed mechanism for lipoprotein remodelling

in the brain. Biochim Biophys Acta 1801:19-23.

Yu L, York J, von Bergmann K, Lutjohann D, Cohen JC, Hobbs HH (2003) Stimulation

of cholesterol excretion by the liver X receptor agonist required ATP-binding cassette

transporters G5 and G8. J Biol Chem 278:15565-15570.

340

Yu L, Li-Hawkins J, Hammer RE, Berge KE, Horton JD, Cohen JC, Hobbs H (2002)

Overexpression of ABCG5 and ABCG8 promotes biliary cholesterol secretion and

reduces fractional absorption of dietary cholesterol. J Clin Invest 110:671-680.

Zafonte RD, Bagiella E, Ansel BM, Novack TA, Friedewald WT, Hesdorffer DC,

Timmons SD, Jallo J, Eisenberg H, Hart T, Ricker JH, Diaz-Arrastia R, Merchant RE,

Temkin NR, Melton S, Dikmen SS (2012) Effect of citicoline on functional and cognitive

status among patients with traumatic brain injury: Citicoline Brain Injury Treatment Trial

(COBRIT). JAMA 308:1993-2000.

Zelcer N, Khanlou N, Clare R, Jiang Q, Reed-Geaghan EG, Landreth GE, Vinters HV,

Tontonoz P (2007) Attenuation of neuroinflammation and Alzheimer's disease pathology

by liver x receptors. Proc Natl Acad Sci U S A 104:10601-10606.

Zhang L, Yang KH, King AI (2004) A proposed injury threshold for mild traumatic brain

injury. J Biomech Eng 126:226-236.

Zhao YD, Wang W (2001) Neurosurgical trauma in People's Republic of China. World J

Surg 25:1202-1204.

Zhi D, Zhang S, Lin X (2003) Study on therapeutic mechanism and clinical effect of mild

hypothermia in patients with severe head injury. Surg Neurol 59:381-385.

341

Zhong N, Weisgraber KH (2009) Understanding the Association of Apolipoprotein E4

with Alzheimer Disease: Clues from Its Structure. J Biol Chem 284:6027-6031.

Zhou W, Xu D, Peng X, Zhang Q, Jia J, Crutcher KA (2008) Meta-analysis of APOE4

allele and outcomes after traumatic brain injury. J Neurotrauma 25:279-290.

Ziebell JM, Taylor SE, Cao T, Harrison JL, Lifshitz J (2012) Rod microglia: elongation,

alignment, and coupling to form trains across the somatosensory cortex after

experimental diffuse brain injury. J Neuroinflammation 9:247.

Zink BJ, Szmydynger-Chodobska J, Choobski A (2010) Emerging concepts in the

pathophysiology of trauamtic brain injury. Psychiatr Clin North Am 33:741-756.

Zlokovic BV (2004) Clearing amyloid through the blood–brain barrier. J Neurochem

89:807-811.

Zohar O, Rubovitch V, Milman A, Schreiber S, Pick CG (2011) Behavioral

consequences of minimal traumatic brain injury in mice. Acta Neurobiol Exp 71:36-45.

Zohar O, Schreiber S, Getslev V, Schwartz JP, Mullins PG, Pick CG (2003) Closed-

head minimal traumatic brain injury produces long-term cognitive deficits in mice.

Neuroscience 118:949-955.

342

Zuccarello M, Iavicoli R, Pardatscher K, Cervellini P, Fiore D, Mingrino S, Gerosa M

(1981) Posttraumatic intfaventricular haemorrhages. Acta Neurochirurgica 55:283-293.

Zweckberger K, Eros C, Zimmermann R, Kim SW, Engel D, Plesnila N (2006) Effect of

early and delayed decompressive craniectomy on secondary brain damage after

controlled cortical impact in mice. J Neurotrauma 23:1083-1093.

343