Structural Brain Abnormalities in Temporomandibular Disorders

346
Structural Brain Abnormalities in Temporomandibular Disorders by Massieh Moayedi A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Institute of Medical Science University of Toronto © Copyright by Massieh Moayedi 2012

Transcript of Structural Brain Abnormalities in Temporomandibular Disorders

Structural Brain Abnormalities in Temporomandibular Disorders

by

Massieh Moayedi

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Institute of Medical Science University of Toronto

© Copyright by Massieh Moayedi 2012

ii

Structural Brain Abnormalities in Temporomandibular Disorder

Massieh Moayedi

Doctor of Philosophy

Institute of Medical Science University of Toronto

2012

Abstract

Temporomandibular disorders (TMD) are a family of prevalent chronic pain disorders affecting

masticatory muscles and/or the temporomandibular joint. There is no unequivocally recognized

peripheral aetiology for idiopathic TMD. The central nervous system (CNS) may initiate and/or

maintain the pain in idiopathic TMD due to sustained or long-term nociceptive input that induces

maladaptive brain plasticity, and/or to inherent personality-related factors that may reduce the

brain's capacity to modulate nociceptive activity. The main aim of this thesis is to determine

whether there are structural neural abnormalities in patients with TMD, and whether these

abnormalities are related to TMD pain characteristics, or to neuroticism. The specific aims are to

delineate in TMD: (1) gray matter (GM) brain abnormalities and the contribution of pain and

neuroticism to abnormalities; (2) the contribution of abnormal brain GM aging in focal cortical

regions associated with nociceptive processes; and (3) abnormalities in brain white matter and

trigeminal nerve and the contribution of pain. In groups of 17 female patients with TMD and 17

age- and sex- matched controls, magnetic resonance imaging revealed that patients with TMD

had: (1) thicker cortex in the somatosensory, ventrolateral prefrontal and frontal polar cortices

iii

than controls, (2) cortical thickness in motor and cognitive areas that was negatively related to

pain intensity, orbitofrontal cortical thickness that was negatively correlated to pain

unpleasantness, and thalamic GM volume correlated to TMD duration, (3) an abnormal

relationship between neuroticism and orbitofrontal cortical thickness, (4) abnormal GM aging in

nociceptive, modulatory and motor areas, (5) widespread abnormalities in white matter tracts in

the brain related to sensory, motor and cognitive functions, (6) reduced trigeminal nerve integrity

related to pain duration, and (7) abnormal connectivity in cognitive and modulatory brain

regions. In sum, this thesis demonstrates for the first time abnormalities in both peripheral nerve

and CNS in patients with TMD.

Acknowledgments This thesis is dedicated to my father, who instilled in me a deep appreciation for learning, an

insatiable curiosity to understand the world, the drive to do so, and the humbleness to realize that

I may not understand everything. I owe my successes to his unwavering love and support.

I am grateful to Dr. Karen Davis for her mentorship. As a supervisor, Karen has a unique

understanding of her students’ needs and addressing them. Her training is unparalleled: the

emphasis and careful study of the body of literature is truly humbling. She has a deep

understanding of the scientific method, and her meticulous methodology maintains a level of

rigour and a high standard of quality in the work produced in her lab. Karen’s greatest skill is

transferring this deep appreciation to her students. Her patience and flexibility allow students to

find themselves a niche within the field, and to develop an expertise. Her rigour ensures that the

questions are addressed appropriately, and that studies are well-conducted. The greatest thing I

will take away from Karen’s lab is a strong sense of integrity. This is the greatest gift a young

scientist can be given – Thank you, Karen.

I am also very grateful to my committee members, Drs. Adrian Crawley, Barry Sessle and

Howard Tenenbaum, for their guidance and direction. They have helped me tremendously in the

development of the questions posed in this thesis, through helping me navigate the literature, and

understand complex concepts. I would like to especially thank Adrian for all of the time he spent

with me outside the lab, and the friendship we have developed. I would also like to thank Dr.

Mary Pat McAndrews – an unofficial committee member who has provided me with much

feedback, support and was always ready to lend an ear.

I also thank Dr. Bruce Freeman and Dr. Michael Goldberg for patient recruitment at the Mount

Sinai Dental Clinic, and Mr. Geoff Pope, Mr. Keith Ta and Mr. Eugen Hlasny for scanning our

subjects and helping us troubleshoot.

I would like to express my sincerest gratitude to my lab mates. I have had the opportunity to

work with several brilliant and hardworking colleagues. Udi Blankstein and Jerry Chen and I

joined the Davis lab in 2007, and we shared many experiences together. Working alongside them

created a very positive working environment, filled with ideas, and laughter. Keri Taylor and

Javeria Hashmi were PhD candidates when we joined the lab. They provided us with much

v

guidance, leadership and support. They also helped us stay on track and helped us develop a

deeper understanding of science. I am very grateful for time and the many conversations we had

together.

I would especially like to thank Irit Weissman-Fogel who was a post-doctoral fellow in our lab.

Without her hard work, dedication, meticulousness and patience, this thesis would have never

been completed. Her contributions to our data collection, study design and analysis were

enormous. Irit was also a unique role model – I will never forget her discipline at work, and her

amazing ability to maintain such a work-life balance. Her dedication to her work and her family

is nothing short of admirable.

During my training, the lab had a turnover, and the Davis lab was full of new faces and

colleagues. Drs. Nathalie Erpelding and Tim Salomons have been phenomenal colleagues and

are great friends. Both have helped me academically, and have become some of my closest

friends. Qi Wu, Danielle DeSouza, Aaron Kucyi, Gang Wang, and Ruma Goswami are great

additions to the lab, full of ideas. It’s been a real pleasure working alongside such brilliant minds

for the past five years, and I will take away many pleasant memories.

I would like to acknowledge the funding agencies that have made my research and this thesis

possible: The Ontario government for awarding me an Ontario Graduate Scholarship for the first

year of my graduate studies; the Canadian Institutes of Health Research (CIHR), for awarding

me a Banting and Best Canada Graduate Scholarship – Doctoral Research Award; the CIHR

Strategic Training Initiatives in Health Research programs: Pain: Molecules to Community and

Cell Signaling in Mucosal Inflammation and Pain for their support. This project was funded by

operating grants from the Canadian Institutes of Health Research and funds from Karen Davis’

Canada Research Chair.

I would like to thank my friends for their distractions, their support and their love. Your

encouragement got me through this!

I would also like to acknowledge Massey College – a truly unique place at the heart of the

University of Toronto. Heartfelt thanks to John Fraser and Elizabeth McCallum for making

Massey so special.

vi

Lastly, and most importantly, I thank my family. Albaloo, you are the most important person in

my life. You are the best role model. Thank you so much for everything. Chebli – what can I

say? No one is quite like you. You’re brilliant and I envy your creativity. Siamak – thank you for

being a groundstone, and a source of unwavering support, and always being the voice of reason. I

would like to thank my mom for all of her support, from close and afar. Mercedes and Debbie –

thank you both for your support. Lizbeth – merci pour ton support, ton amitié, et tes aiguilles. I

love you all!

Human beings are members of a whole, In creation of one essence and soul. If one member is afflicted with pain, Other members uneasy will remain. If you've no sympathy for human pain, The name of human you cannot retain!

-Sa’adi, Golestan, 1258

viii

It is a shame that we possess such insufficient knowledge concerning the character of pain – those symptoms which represent the essential part of all bodily suffering of man.

-A. Goldscheider, Ueber den Schmerz, 1894

ix

Table of Contents

Acknowledgments .......................................................................................................................... iv!

Table of Contents ........................................................................................................................... ix!

List of Tables ............................................................................................................................. xviii!

List of Figures .............................................................................................................................. xix!

List of Appendices ....................................................................................................................... xxi!

List of Abbreviations .................................................................................................................. xxii!

Chapter 1 Introduction and General Aims ...................................................................................... 1!

Chapter 2 Literature Review ........................................................................................................... 4!

2.1.! What is pain? ......................................................................................................................... 4!

2.1.1.! Historical perspectives and theories of pain ........................................................... 5!

2.1.1.1.! Specificity Theory ................................................................................... 5!

2.1.1.2.! Intensity theory of Pain ......................................................................... 11!

2.1.1.3.! Pattern Theory ....................................................................................... 12!

2.1.1.4.! Gate Control Theory of Pain ................................................................. 12!

2.1.2.! Multidimensional aspects of pain ......................................................................... 15!

2.1.3.! Chronification of pain ........................................................................................... 16!

2.1.3.1.! Sensitization, Hyperalgesia and Allodynia ........................................... 16!

2.1.3.2.! The role of glia in sensitization ............................................................. 18!

2.1.4.! Functional pain syndromes ................................................................................... 18!

2.1.4.1.! Central abnormalities in functional pain syndromes ............................. 19!

2.1.5.! Pain-cognition interactions ................................................................................... 19!

2.1.6.! Salience and pain .................................................................................................. 22!

2.1.7.! Pain-motor interactions ......................................................................................... 23!

2.1.8.! Neuroticism and pain ............................................................................................ 24!

x

2.2.! Temporomandibular disorders ............................................................................................ 25!

2.2.1.! Historical perspectives .......................................................................................... 25!

2.2.2.! Prevalence and social costs ................................................................................... 25!

2.2.3.! Clinical aspects ..................................................................................................... 26!

2.2.3.1.! Sensory abnormalities ........................................................................... 27!

2.2.3.2.! Motor abnormalities .............................................................................. 34!

2.2.3.3.! Cognitive abnormalities ........................................................................ 34!

2.3.! Anatomy and physiology of orofacial pain and sensation .................................................. 35!

2.3.1.! Peripheral receptors and sensory afferents ........................................................... 35!

2.3.2.! Orofacial nervous system: the trigeminal nerve ................................................... 37!

2.3.3.! White matter pathways ......................................................................................... 40!

2.3.4.! Nociceptive pathways and mechanisms ................................................................ 43!

2.3.4.1.! Peripheral nociceptive mechanisms ...................................................... 43!

2.3.4.2.! Spinal nociceptive pathways and mechanisms ...................................... 44!

2.3.4.3.! Trigeminal nociceptive pathways and mechanisms .............................. 47!

2.3.4.4.! Trigeminal motor pathways ................................................................... 49!

2.3.4.5.! Supraspinal nociceptive and pain regions ............................................. 52!

Thalamus ...................................................................................................... 54!

Primary somatosensory cortex .................................................................... 55!

The parasylvian cortex: S2 and insula ......................................................... 57!

The cingulate cortex ..................................................................................... 59!

Prefrontal cortex .......................................................................................... 62!

Motor regions .............................................................................................. 63!

2.3.5.! Descending modulation ........................................................................................ 66!

xi

2.3.5.1.! Descending modulation pathways ......................................................... 68!

2.3.5.2.! Diffuse noxious inhibitory controls ....................................................... 69!

2.4.! Structural brain imaging ..................................................................................................... 71!

2.4.1.! MRI ....................................................................................................................... 71!

2.4.1.1.! What is magnetic resonance? ................................................................ 71!

2.4.1.2.! MR signals ............................................................................................. 74!

Magnetization time (T1) ............................................................................... 74!

Signal decay time (T2) ................................................................................. 75!

Spatial Encoding .......................................................................................... 75!

Contrast ....................................................................................................... 76!

2.4.2.! Gray matter imaging ............................................................................................. 77!

2.4.2.1.! Voxel-based morphometry .................................................................... 77!

2.4.2.2.! Cortical thickness .................................................................................. 78!

2.4.3.! White matter imaging ........................................................................................... 79!

2.4.3.1.! Diffusion tensor imaging ....................................................................... 80!

2.4.3.2.! Between Group Comparisons of White Matter: Tract-based spatial statistics 85!

2.4.3.3.! Tractography ......................................................................................... 87!

Streamline (deterministic) tractography ...................................................... 88!

Probabilistic tractography ........................................................................... 89!

2.5.! Brain imaging of pain ......................................................................................................... 91!

2.5.1.! Functional neuroimaging of pain .......................................................................... 91!

2.5.1.1.! Imaging acute pain ................................................................................ 92!

S1 ................................................................................................................. 93!

xii

S2/Parietal operculum ................................................................................. 94!

Insula ........................................................................................................... 95!

Cingulate cortex ........................................................................................... 96!

Prefrontal cortex .......................................................................................... 97!

Cortical and subcortical motor regions ....................................................... 98!

2.5.1.2.! Imaging salience and pain ..................................................................... 98!

2.5.1.3.! Chronic pain .......................................................................................... 99!

2.5.2.! Structural imaging of pain .................................................................................. 101!

2.5.2.1.! Acute Pain ........................................................................................... 101!

2.5.2.2.! Chronic Pain ........................................................................................ 102!

Gray matter abnormalities ......................................................................... 102!

White matter abnormalities ........................................................................ 105!

2.6.! Neuroplasticity .................................................................................................................. 111!

2.6.1.! Use-dependent neuroplasticity ............................................................................ 112!

2.6.2.! Neuroplasticity and pain ..................................................................................... 113!

2.6.3.! Age-related neuroplasticity ................................................................................. 115!

2.6.4.! Microstructural basis of neuroplasticity .............................................................. 116!

Chapter 3 Rationale, Specific Aims and Hypotheses ................................................................. 118!

3.1.! Study I. Contribution of chronic pain and neuroticism to abnormal forebrain gray matter in TMD patients ............................................................................................................... 118!

3.2.! Study II. Age-related gray matter abnormalities in patients with TMD ........................... 120!

3.3.! Study III. White matter brain abnormalities in temporomandibular disorder ................... 121!

Chapter 4 General Methods ........................................................................................................ 123!

4.1.! Overview of Project .......................................................................................................... 123!

xiii

4.2.! Subject Recruitment .......................................................................................................... 123!

4.3.! Questionnaires ................................................................................................................... 124!

4.3.1.! Edinburgh Handedness Inventory ....................................................................... 124!

4.3.2.! NEO-Five Factor Inventory-Short Form ............................................................ 125!

4.3.3.! Pain Catastrophizing Scale ................................................................................. 125!

4.3.4.! Patient Interview ................................................................................................. 125!

4.4.! MR Imaging ...................................................................................................................... 126!

4.4.1.! Study design ........................................................................................................ 126!

4.4.2.! Imaging parameters ............................................................................................. 126!

4.4.3.! Gray Matter Analysis .......................................................................................... 127!

4.4.3.1.! Cortical Thickness Analysis ................................................................ 127!

4.4.3.2.! Voxel-Based Morphometry ................................................................. 130!

4.4.4.! White matter analysis .......................................................................................... 130!

4.4.4.1.! Tract-based spatial statistics ................................................................ 131!

4.4.4.2.! TBSS with other DTI metrics .............................................................. 131!

4.4.4.3.! Probabilistic tractography .................................................................... 132!

Chapter 5 STUDY I: Contributions of chronic pain and neuroticism to abnormal forebrain gray matter in TMD ............................................................................................................... 133!

5.1.! Introduction ....................................................................................................................... 133!

5.2.! Methods ............................................................................................................................. 135!

5.2.1.! Subjects ............................................................................................................... 135!

5.2.2.! Questionnaires ..................................................................................................... 135!

5.2.3.! Imaging parameters ............................................................................................. 136!

5.2.4.! Structural brain imaging analysis ........................................................................ 136!

5.2.4.1.! Cortical thickness analysis .................................................................. 136!

5.2.4.2.! Subcortical analysis with VBM ........................................................... 137!

xiv

5.2.5.! Statistical analyses .............................................................................................. 137!

5.2.5.1.! Demographics ...................................................................................... 137!

5.2.5.2.! Gray matter group differences ............................................................. 138!

CTA ............................................................................................................ 138!

VBM ........................................................................................................... 139!

5.2.5.3.! Clinical correlates ................................................................................ 139!

5.2.5.4.! Neuroticism effects .............................................................................. 139!

5.3.! Results ............................................................................................................................... 140!

5.3.1.! Patient demographics .......................................................................................... 140!

5.3.2.! Group differences: S1 and frontal thickening in TMD ....................................... 140!

5.3.3.! Effect of chronic pain intensity and unpleasantness ........................................... 141!

5.3.4.! Effect of TMD chronicity ................................................................................... 141!

5.3.5.! Effect of neuroticism ........................................................................................... 141!

5.4.! Discussion ......................................................................................................................... 154!

5.4.1.! S1 and thalamic gray matter increases in TMD .................................................. 154!

5.4.2.! Cortical thickening in cognitive and modulatory regions in TMD ..................... 154!

5.4.3.! Relationship between chronic pain intensity and cortical thickness in M1 and aMCC .................................................................................................................. 155!

5.4.4.! Interaction between neuroticism and group in prefrontal gray matter ................ 157!

5.4.5.! Study limitations ................................................................................................. 157!

5.5.! Conclusions ....................................................................................................................... 158!

Chapter 6 STUDY II: Age-related gray matter abnormalities in temporomandibular disorder . 159!

6.1.! Introduction ....................................................................................................................... 159!

6.2.! Methods ............................................................................................................................. 160!

6.2.1.! Subjects ............................................................................................................... 160!

6.2.2.! Imaging & Analysis ............................................................................................ 161!

xv

6.2.3.! Global age effects ............................................................................................... 161!

6.2.4.! Age-by-group effects in gray matter ................................................................... 161!

CTA ...................................................................................................................... 161!

VBM ..................................................................................................................... 162!

6.2.5.! Contribution of TMD duration to age-related gray matter abnormalities ........... 162!

6.3.! Results ............................................................................................................................... 163!

6.3.1.! Patient characteristics .......................................................................................... 163!

6.3.2.! Global age effects ............................................................................................... 163!

6.3.3.! Focal age effects ................................................................................................. 163!

6.3.4.! The contribution of TMD duration to gray matter age effects ............................ 164!

6.4.! Discussion ......................................................................................................................... 174!

6.4.1.! Whole brain gray matter atrophy ........................................................................ 174!

6.4.2.! Use-dependent plasticity ..................................................................................... 174!

6.4.3.! Cellular basis of changes in gray matter ............................................................. 176!

6.4.4.! Study limitations ................................................................................................. 177!

6.5.! Conclusion ........................................................................................................................ 177!

Chapter 7 STUDY III: White matter brain and trigeminal abnormalities in TMD .................... 178!

7.1.! Introduction ....................................................................................................................... 178!

7.2.! Methods ............................................................................................................................. 179!

7.2.1.! Subjects ............................................................................................................... 179!

7.2.2.! Questionnaires ..................................................................................................... 180!

7.2.3.! Imaging parameters ............................................................................................. 180!

7.2.4.! DTI pre-processing ............................................................................................. 181!

7.2.4.1.! Other DTI metrics ............................................................................... 181!

7.2.5.! Assessment of CNV ............................................................................................ 181!

7.2.6.! Tract-based spatial statistics ............................................................................... 182!

xvi

7.2.7.! Probabilistic tractography ................................................................................... 183!

7.2.8.! Statistical analyses .............................................................................................. 183!

7.2.8.1.! Whole brain white matter .................................................................... 183!

7.2.8.2.! Mask analysis ...................................................................................... 184!

7.2.8.3.! Quantitative Tractography: connection probability ............................ 185!

7.3.! Results ............................................................................................................................... 185!

7.3.1.! Patient characteristics .......................................................................................... 185!

7.3.2.! Trigeminal nerve FA ........................................................................................... 186!

7.3.3.! Patients have lower white matter FA .................................................................. 186!

7.3.4.! Probabilistic tractography ................................................................................... 187!

7.3.5.! White matter FA related to TMD pain characteristics ........................................ 188!

7.4.! Discussion ......................................................................................................................... 201!

7.4.1.! Trigeminal nerve abnormalities .......................................................................... 201!

7.4.2.! Abnormal sensorimotor tracts in TMD ............................................................... 202!

7.4.3.! Cognitive interference of pain and decreased modulation in TMD .................... 202!

7.4.4.! Abnormal connectivity in TMD .......................................................................... 203!

7.4.5.! Intense and prolonged TMD may drive plasticity in white matter ..................... 203!

7.4.6.! Cellular basis of changes in FA .......................................................................... 204!

7.4.7.! Study limitations ................................................................................................. 205!

7.5.! Conclusion ........................................................................................................................ 205!

Chapter 8 General Discussion ..................................................................................................... 206!

8.1.! Pain-driven abnormalities along nociceptive pathways .................................................... 207!

8.2.! Abnormalities in motor regions ........................................................................................ 209!

8.3.! Abnormalities in cognitive-modulatory brain regions ...................................................... 211!

8.4.! Neuroticism and TMD ...................................................................................................... 214!

8.5.! Basis of structural brain abnormalities ............................................................................. 216!

xvii

8.6.! Study limitations ............................................................................................................... 217!

8.7.! Future directions ............................................................................................................... 218!

8.8.! Conclusions ....................................................................................................................... 220!

References ................................................................................................................................... 221!

Appendices .................................................................................................................................. 305!

Copyright Acknowledgements .................................................................................................... 319!

xviii

List of Tables

Table 2-1: Summary of quantitative sensory testing studies in TMD .......................................... 30

Table 2-2: Number of quantitative sensory testing studies testing sensory thresholds and tolerance in TMD .......................................................................................................................... 33!

Table 2-3: Summary of regions with gray matter abnormalities in chronic pain ....................... 107

Table 5-1: Patient demographics ................................................................................................ 143!

Table 5-2: Group differences in cortical thickness ..................................................................... 145!

Table 5-3: Cortical thickness negatively correlates with TMD pain intensity or unpleasantness ............................................................................................................................. 146!

Table 6-1 Distribution of subject ages ........................................................................................ 166

Table 6-2: Age-related group differences in cortical thickness and subcortical gray matter

volume ......................................................................................................................................... 167!

Table 6-3: Contributions of TMD duration to the age-gray matter relationships ....................... 168

Table 7-1: Group differences in whole brain skeletonised white matter. ................................... 189!

Table 7-2: White matter regions in patients with TMD with significantly lower fractional anisotropy compared to controls ................................................................................................. 190!

Table 7-3: Group differences in mean diffusivity (MD), radial diffusivity (RD) and axial diffusivity (λ1) in clusters with significant group differences in FA .......................................... 191!

xix

List of Figures

Figure 2-1: Descartes’ line drawing of the pain system. ................................................................ 7

Figure 2-2: Schematic of the Gate Control Theory of Pain mechanisms ..................................... 14

Figure 2-3: Schematic of diffusion tensor imaging, and summary of eigenvectors and metrics .. 42

Figure 2-4: Diffusion tensor imaging reconstruction of Meynert’s classification of white matter tracts in the brain ........................................................................................................................... 84

Figure 2-5: White matter skeleton from tract-based spatial statistics ........................................... 87

Figure 4-1: Template deformation process for skull-stripping in FreeSurfer ............................ 128

Figure 4-2: Cutting planes overlaid on sections through white matter-labeled volume ............. 128

Figure 4-3 White surface overlaid on original T1 scan .............................................................. 129

Figure 4-4 Pial surface overlaid on original T1 scan .................................................................. 129

Figure 4-5 Original (left), gray/white boundary (middle), and pial surface (right) reconstructions of a left hemisphere ..................................................................................................................... 129

Figure 5-1: Masks used to restrict analyses ................................................................................ 147!

Figure 5-2: Cortical thickening in TMD ..................................................................................... 149!

Figure 5-3: TMD pain intensity and unpleasantness are negatively correlated to modulatory regions ......................................................................................................................................... 150 !

Figure 5-4: TMD duration is related to plasticity in the thalamus .............................................. 152!

Figure 5-5: Patients show an abnormal positive relationship between OFC thickness and neuroticism .................................................................................................................................. 153!

Figure 6-1: Total gray matter volume and correlations with age and duration ........................... 169!

Figure 6-2: Age-by-group interactions in cortical thickness ...................................................... 171!

Figure 6-3: Group differences in age effects within subcortical gray matter volume ................ 172!

Figure 6-4: Summary of age-related gray matter abnormalities ................................................. 173!

Figure 7-1: Schematic of diffusion tensor imaging, and summary of eigenvectors and metrics 182!

Figure 7-2: Colour orientation maps of the trigeminal nerves .................................................... 192!

Figure 7-3: Trigeminal nerve fractional anisotropy abnormalities in TMD ............................... 193!

xx

Figure 7-4: Group differences in mean fractional anisotropy between patients with TMD and controls ........................................................................................................................................ 195!

Figure 7-5: Group differences in fractional anisotropy between patients with TMD and controls ........................................................................................................................................ 196!

Figure 7-6: Clusters with significant group differences in fractional anisotropy ....................... 197!

Figure 7-7: Abnormal white matter connectivity in TMD .......................................................... 198!

Figure 7-8: Regions with group differences in fractional anisotropy are also correlated with TMD characteristics .................................................................................................................... 200

Figure 8-1 Proposed model of neural basis of sensory abnormalities in TMD .......................... 209

Figure 8-2 Proposed model of neural basis of motor abnormalities in TMD ............................. 211

Figure 8-3 Proposed model of the neural basis for cognitive abnormalities in TMD ................ 213

Figure 8-4 Proposed model for the neural basis of modulatory system abnormalities in TMD . 216

xxi

List of Appendices

Appendix I: Recruitment Letter .................................................................................................. 293

Appendix II: Consent forms ........................................................................................................ 294!

Appendix III: Edinburgh Handedness Inventory ........................................................................ 299

Appendix IV: Pain Catastrophizing scale ................................................................................... 300

Appendix IV: TMD pain assessment .......................................................................................... 301

Appendix V: McGill Pain Questionnaire .................................................................................... 304!

Appendix VI: McGill Pain Questionnaire results for patients with TMD .................................. 305!

Appendix VIII: Making Sense of Gray Matter Abnormalities in Chronic Orofacial Pain ......... 306!

List of Abbreviations

AAOP American Association of Orofacial Pain ACC Anterior cingulate cortex ADC Apparent diffusion coefficient aMCC Anterior mid-cingulate cortex AMH Aδ-mechano-heat sensitive afferent Amyg Amygdala ASSET Array Spatial Sensitivity Encoding Technique B0 Magnetic field BA Brodmann’s Area BCE Before Common Era BET Brain Extraction Tool BG Basal ganglia Cb Cerebellum CBP Chronic back pain cCMA Caudal cingulate motor area CH Cluster headache CL Centrolateral nucleus of the thalamus CM Centromedian nucleus of the thalamus CMA Cingulate motor area CMH C-mechano-heat sensitive afferent CNS Central nervous system CNV Cranial nerve five (trigeminal nerve) CNVII Facial nerve CNIX Glossopharyngeal nerve CNX Vagus nerve CPP Chronic pelvic pain CPTH Chronic posttraumatic headache CRPS Complex regional pain syndrome CSF Cerebrospinal fluid CTA Cortical thickness analysis CTTH Chronic tension-type headache DBM Deformation-based morphometry DCML Dorsal-column medial-leminiscal pathway dlPFC Dorsolateral prefrontal cortex DLPT Dorsolateral pontomesencephalic tegmentum dmPFC Dorsomedial prefrontal cortex DNIC Diffuse noxious inhibitory controls

xxiii

dODF Diffusion orientation density functions dPCC Dorsal posterior cingulate cortex DTI Diffusion tensor imaging DWI Diffusion-weighted imaging EC/ExC External/Extreme capsules EEG Electroencephalography FA Fractional anisotropy FDR False discovery rate FDT FSL Diffusion Toolbox fMRI Functional magnetic resonance imaging FMS Fibromyalgia fODF Fibre orientation density functions FSL FMRIB's Software Library FSPGR Fast spoiled gradient echo FWE Family-wise error FWHM Full-width half-maximum GLM General linear model GMV Gray matter volume GP Globus pallidus HC Hippocampus Hip OA Hip osteoarthritis HPC Heat-pinch-cold HT Hypothalamus HTM High-threshold mechanoreceptors HypH Hypnic headache IASP International Association for the Study of Pain IBS Irritable bowel syndrome IC Internal capsule ICAL Anterior limb of the internal capsule ICBM International Consortium for Brain Mapping iTL Inferior temporal lobe LC Locus coeruleus LEP Laser-evoked potentials LTM Low-threshold mechanoreceptors M1 Primary motor cortex MCC Mid-cingulate cortex MCS Motor cortex stimulation

xxiv

MD Mean diffusivity MD Medial dorsal thalamus MDvc Ventral caudal portion of the medial dorsal nucleus of the thalamus MEG Magnetoencephalography MeT Mesencephalic nucleus of the trigeminal nerve MNI Montreal Neurological Institute MOH Medication-overuse headache mPFC Medial prefrontal cortex MR Magnetic resonance MRI Magnetic resonance imaging MSN Main sensory nucleus NC Caudate nucleus NEO-FFI NEO-Five Factor Inventory NRM Nucleus raphe magnus NS Nociceptive-specific ODF Orientation density functions OFC Orbitofrontal cortex PAG Periaqueductal gray PCS Pain catastrophizing scale PDF Probability density function PET Positron-emission tomography Pf Parafascicular nucleus of the thalamus PFC Prefrontal cortex pgACC Pregenual anterior cingulate cortex PHG Parahippocampal gyrus PIFP Persistent idiopathic facial pain plIC Posterior limb of the internal capsule PMC Premotor cortex pMCC Posterior mid-cingulate cortex PMv Ventral premotor cortex PNS Peripheral nervous system PO Posterior nucleus of the thalamus PPC Posterior parietal cortex pTL Posterior temporal lobe Put Putamen PVD Provoked vestibulodynia QST Quantitative sensory testing RA Relative anisotropy

xxv

rCBF Regional cerebral blood flow rCMA Rostral cingulate motor area RCZ Rostral caudal zone RD Radial diffusivity rf Radiofrequency RF Reticular formation of the brainstem RFT Random field theory Rheum Arth Rheumatoid arthritis ROI Region of interest RVM Rostral ventromedial medulla S Signal S1 Primary somatosensory cortex S2 Secondary somatosensory cortex SC Subcoeruleus sgACC Subgenual cingulate cortex SM Submedian nucleus of the thalamus SMA Supplementary motor area SN Substantia nigra SPECT Single-photon emission computed tomography SPM Statistical parametric mapping STG Superior temporal gyrus STN Spinal trigeminal nucleus STT Spinothalamic tract SVC Small-volume correction T1 Spin-lattice relaxation time or Magnetization time T2 Spin-spin relaxation time or Signal decay time TBSS Tract-based spatial statistics TE Echo time Thal Thalamus TI Inversion time TIV Total intracranial volume TL Temporal lobe TMD Temporomandibular disorder TMD-RDC Temporomandibular disorder research diagnostic criteria TMJ Temporomandibular joint TMS Transcranial magnetic stimulation TN Trigeminal neuralgia TNP Trigeminal neuropathic pain TR Repetition time TTT Trigeminothalamic tract uODF Uncertainty orientation density functions

xxvi

V1 Ophthalmic branch of the trigeminal nerve V1 Primary visual cortex V2 Maxillary branch of the trigeminal nerve V3 Mandibular branch of the trigeminal nerve VBM Voxel-based morphometry VBSNC Trigeminal brainstem nuclear complex Vc Subnucleus caudalis Vi Subnucleus interpolaris VL Ventrolateral nucleus of the thalamus vlPFC Ventrolateral prefrontral cortex vmPFC Ventromedial prefrontal cortex vmPO Posterior region of the ventromedial nucleus of the thalamus Vo Subnucleus oralis vPCC Ventral posterior cingulate cortex VPI Ventroposterioinferior nucleus of the thalamus VPL Ventroposteriolateral nucleus of the thalamus VPM Ventroposteriomedial nucleus of the thalamus vSTR Ventral striatum WDR Wide-dynamic range γ Gyromagnetic ratio λ1 Longitudinal diffusivity ρ Proton density τ Diffusion time ω0 Precession rate

1

Chapter 1 Introduction and General Aims

Temporomandibular disorders (TMD) comprise clinical problems involving the structures of and

around the temporomandibular joint (TMJ), the masticatory musculature, or both [1,418]. TMD

represent the most common orofacial chronic pain disorder [285], and are primarily characterized

by spontaneous pain, or pain associated with jaw function, which can affect mandibular range of

motion and produce TMJ sounds during jaw function. Pain from TMD can arise from the

muscles of mastication, the TMJ, or both, and, in general, when combined (i.e., muscular and

TMJ pain) the chief complaint tends to be muscular in nature [402].

There have not yet been any comprehensive epidemiological studies of TMD in the Canadian

population, and so this thesis will refer to data from the United States. TMD are estimated to

affect between 3-20 percent of the United States’ adult population [269,539,540,562]. Of these,

only a small proportion of persons suffering from TMD pain seek treatment, and are seen at

tertiary and quaternary medical facilities, such as pain clinics [269,350]. It has been estimated

that TMD cost $4 billion annually due to lost-wages and medical treatment. TMD 1.5-9 times

more prevalent in women [113,201,284,285,312,721] and a nationwide epidemiological study in

the United States reported that 84% of persons with TMD were women [284].

In some cases, there is no clear aetiological evidence for TMD pain, and the pain is considered

idiopathic [283,286,287,653]. In this group of patients, it has been suggested that abnormal

function in the central nervous system (CNS) may initiate or maintain TMD pain [757]. One line

of evidence for abnormal CNS function is that TMD symptomatology, including persistent pain,

allodynia, and hyperalgesia occurs not only in the orofacial region, but also in other body sites

[305,314,390,565,756,883]. Patients with TMD also show greater temporal summation of pain to

repetitive noxious heat stimuli [563], and dysfunctional diffuse noxious inhibitory controls

(DNIC) [113,491,756]. These centrally-mediated processes provide further evidence for CNS

dysfunction in TMD and provide evidence for the involvement of ascending nociceptive

pathways and/or descending pain-modulatory pathways [515]. Additionally, patients with TMD

can exhibit cognitive [370,379,380] and motor dysfunction [837] possibly related to

abnormalities in brain regions associated with these functions [799,800,839,928].

2

There are two main routes by which the CNS may contribute to the development and/or

maintenance of chronic pain conditions such as TMD. One possibility is that long-term

nociceptive input into the brain induces maladaptive brain plasticity, which may play a role in

maintaining pain [12,213,576,950]. This maladaptive plasticity suggests that patients with TMD

may be unable to habituate to increased nociceptive activity, which may be related to a reduced

capacity of the brain to dampen pain by descending (top-down) controls [95].

The second route by which the CNS may contribute to the development and/or maintenance of

chronic pain relates to inherent personality-related factors that reduce the brain’s capacity to

modulate nociceptive input. This poor control of pain might lead to the development of

vulnerability towards the development of chronic pain. For example, there is evidence that

neuroticism may be associated with pain-related suffering [408], pain sensitivity [177,373,907],

nerve injury outcomes as well as neuropathic pain [848], inhibition of negative thoughts [178],

and the development of TMD [315].

Structural magnetic resonance imaging (MRI) and diffusion-weighted imaging (DWI) can be

used to investigate the structure of CNS gray and white matter, respectively. Indeed, many

structural MRI studies have demonstrated differences in gray matter in populations of patients

with chronic pain associated with pain-related characteristics (e.g., intensity, unpleasantness, or

duration) (see [98,206,303,356,575,764,772]). Furthermore, structural MRI studies have reported

accelerated loss of gray matter in the brains of patients with chronic pain [27,502]. These age-

related changes in chronic pain may be related to the cumulative effect of pain over time (i.e.,

pain duration) or other pre-existing factors that may alter the normal pattern of aging in TMD. In

addition to abnormalities in central gray matter found in the presence of chronic pain, previous

studies examining measures of white matter integrity in other clinical conditions with sensory

abnormalities and/or chronic pain reported abnormalities in white matter tracts related to sensory,

modulatory and cognitive functions [156,356,558,847]. Furthermore, these studies reported

white matter abnormalities correlated with clinical findings in the patients. Therefore, studying

the correlation between TMD characteristics and measures of white matter integrity can provide

insight into whether chronic pain drives changes in white matter microstructure.

With regard to the second route by which the CNS may contribute to the development and/or

maintenance of chronic pain, pre-existing abnormalities may be a factor for the development of

3

TMD. These structural abnormalities in the CNS could be related to personality traits that are

related to chronic pain, such as neuroticism [315]. However, not all patients with chronic pain

have high neuroticism scores, and not all persons with neuroticism have chronic pain [179].

Therefore, neuroticism alone is not sufficient to develop chronic pain. Rather, there may exist an

abnormal relationship between neuroticism and brain structure, which may impair function in

pain-modulatory brain regions, such as the orbitofrontal cortex (OFC) [954] or the medial

prefrontal cortex (mPFC) [245,386] and this could facilitate or maintain chronic pain. However,

the precise mechanism by which neuroticism can contribute to TMD development remains to be

elucidated.

Given the poor understanding of the mechanisms underlying TMD, it is challenging to develop

rational treatment approaches for this chronic pain condition. Moreover, it is becoming clearer

that TMD may be mediated by central mechanisms. However, while there are some data

demonstrating neuropsychological and cognitive, as well as psychosocial contributors to TMD,

there are little data demonstrating whether there are underlying neuroanatomical changes that

could support the notion that TMD is a centrally-mediated pain condition.

Therefore, the main aims of this thesis are:

1. To examine gray matter abnormalities in patients with idiopathic TMD and to

determine the contribution of pain-related characteristics and neuroticism;

2. To determine whether the cumulative effect of idiopathic TMD interacts with the

normal aging process in focal cortical regions associated with nociceptive processes;

3. To evaluate white matter brain and trigeminal nerve (cranial nerve five: CNV)

abnormalities in patients with idiopathic TMD and to determine the contribution of

pain-related characteristics.

4

Chapter 2 Literature Review

2.1. What is pain?

The current definition of pain, established by International Association for the Study of Pain

(IASP) in 1986, defines pain as “an unpleasant sensory and emotional experience associated with

actual or potential tissue damage, or described in terms of tissue damage, or both.” It is important

to recognize that pain and nociception are not synonymous. Nociception, based on the 2011

IASP taxonomy [110], is “the neural process of encoding noxious stimuli” and is initiated by the

activation of peripheral receptors that preferentially respond to stimuli in the noxious range,

referred to as nociceptors. It includes the neural activity in the peripheral nervous systems (PNS)

and CNS, elicited by a noxious stimulus. In contrast to nociception, pain requires a conscious

perception.

Pain is a complex, multidimensional experience involving sensory, affective, cognitive,

attentional, motivational, and motor components [593]. The sensorial dimension of pain relates

to the discriminability of pain qualities, location, intensity and duration of pain. Pain is inherently

salient. Based on the context, our psychological and affective state, we initiate a response to

noxious stimulus. Therefore, different people respond to pain differently.

Pain can be categorized based on its temporal profile: short-term pain is known as acute pain,

and lasts from seconds to days. Acute pain plays an essential role in signaling actual or potential

tissue damage [442]. However, chronic pain persists more than three to six months, and is not

necessarily related with tissue damage [791]. The aetiology of chronic pain is not as evident as it

is in acute pain. Often, chronic pain persists beyond the time necessary for the tissue to heal

normally following injury and, in these cases, it is believed that nociceptive input and/or the CNS

maintain pain [576]. The IASP has devised a classification scheme for chronic pain [600], based

on five axes: (1) the region of the pain, (2) the system affected, (3) temporal characteristics of the

pain, (4) time since onset, and (5) the aetiology of the pain.

5

2.1.1. Historical perspectives and theories of pain

A number of theories have been postulated to describe mechanisms underlying pain perception.

These theories date back several centuries, and even millennia [483,729]. This thesis will

primarily focus on theories that were postulated since the 17th century.

2.1.1.1. Specificity Theory

Specificity theory refers to the presence of dedicated pathways for each somatosensory modality.

The fundamental tenet of the specificity theory is that each modality has a specific receptor and

associated sensory fibre (primary afferent) that is sensitive to one specific stimulus [273]. For

instance, the model proposes that non-noxious mechanical stimuli are encoded by low-threshold

mechanorecpetors (LTM), which are associated with dedicated primary afferents that project to

mechanoreceptive second-order neurones in the spinal cord or brainstem (depending on the

source of the input). These second-order neurones project to higher mechanoreceptive areas in

the brain. Similarly, noxious stimuli would activate a nociceptor, which would project to higher

pain centres through a pain fibre. These ideas have been emerging over several millennia, but

were experimentally tested, and formally postulated as a theory in the nineteenth century by

physiologists in Western Europe. This thesis will briefly review some of the key ideas that have

led to the development of specificity theory.

René Descartes was one of the first western philosophers to describe a detailed pain pathway in

humans. Descartes’ manuscript Treatise of Man (originally written in French) was illustrated,

edited and published posthumously, first in Latin in 1662 [241], and then in French in 1664

[242]. In Treatise of Man, based on the French edition by Louis La Forge (who was also one of

the illustrators), Descartes describes pain as a perception that exists in the brain and makes the

distinction between the neural phenomenon of sensory transduction (today known as

nociception) and the perceptual experience of pain. What is essential to the development of

Descartes’ theory is his description of nerves, which he perceived as hollow tubules that convey

both sensory and motor information. This understanding of neural function was by no means

novel. In the 3rd century BCE, Herophilus demonstrated the existence of sensory and motor

nerves, and Erasistratus demonstrated that the brain influenced motor activity [729]. Half a

millennium later, Galen demonstrated that sectioning the spinal cord caused sensory and motor

deficits [651]. Within the spirit of scientific enquiry that resurfaced in the renaissance,

6

anatomical studies by Vesalius published in 1543 reiterated and confirmed Galen’s findings

[651]. In relation to this, Galen had postulated that three conditions be met for perception: (1) an

organ must to be able to receive the stimulus, (2) a connection from the organ to the brain, and

(3) a processing centre that converts the sensation to a conscious perception [729]. Descartes

contributed to Galen’s model by postulating that a gate existed between the brain and the tubular

structures (the connections), which was opened by a sensory cue [242]. A sensory cue would

“tug” on the tube, which would then open a gate between the tube and the brain. The opening of

this gate would then allow “Animal Spirits” (an extension of the Greek pneuma1) to flow through

these tubes and within the muscles to move them. Although this sensory system was not specific

to pain, La Forge’s drawing (based on Descartes’ concept) of a foot near a flame is one of the

most ubiquitous figures in neuroscience (see Figure 2-1). This example describes the pathway for

promptly moving one’s foot away from a hot flame. In the figure (and its description in the text)

the heat of the flame near the foot activates a fibril (or fibre) within the nerve tubule that

traverses up the leg, to the spinal cord and finally to the brain. Descartes compared this fibre to a

cord attached to a bell – by pulling on the other end of cord, the bell will ring. The proverbial

bells in this case are the pores that line the ventricles in the brain. Once these pores open in

response to the sensory input, the Animal Spirits were thought to flow through the tubule and

elicit a motor response. This motor response included turning the head and the eyes to see the

flame, and raising the hands and folding the body away from the flame for protection. Descartes

conceived that there are many of these fibrils and that their movements elicit the sensations. For

example, the perception of pain would be felt in the brain when there is a significant “tug” on the

fibre, which caused it to sever. In contrast, a tug of the same magnitude that doesn’t cause the

fibre to break would evoke a tickling (or tingling – Descartes uses the French word

chatouillement) perception. Although Descartes’ figure of the boy and the flame suggests that

there is a dedicated pain pathway, a closer read of the text indicates that Descartes believed that

the pattern and rate of firing (intensity of tugging) of a fibre provided the adequate information to

the brain about the stimulus intensity and quality.

1In ancient Greek medicine, the concept of pneuma represents air, breath, or motion. It was also called the breath of

life and the spirit of man. [87] Bernard C, Magendie F. Leçon d’ouverture du Cours de Médecine du Collège de France. Paris: Baillière, 1856.

7

Figure 2-1: Line drawing of the pain system by Florentius Schuyl (left) and Louis La Forge (right), based on Descartes description in Treatise of Man: The fire (A) is close to the foot (B). Particles from the fire move and press the skin, and tug on the fibril (C), which opens the pore (d, e) where the fibril terminates. Opening the pore is akin to tugging on a rope attached to a bell, thus ringing the bell. The open pore allows the “animal spirits” to flow from the cavity (F) into the fibril and, part of them activate the muscles to move the foot away from the fire, and part of them activate the muscles to turn the eyes and the head toward the fire to look at it, and part of them are used to bring forth the hands and fold the body to protect it. (Image on the left is reproduced from and the image on the right and text is reproduced from [242], out of copyright; translated by Massieh Moayedi).

8

The concept of a dedicated pain pathway (also known as specificity theory) was developed by

Charles Bell in his landmark essay Idea of a new anatomy of the brain; submitted for the

observation of his friends, first published as a conference proceeding in 1811, and later

reproduced in a journal [79]. In this essay, Bell provided an alternative perspective about the

organization of the nervous system. First he suggested that the brain is not “common sensorium”

as suggested by Descartes, which was the accepted model of the brain at the time. Instead, he

provided anatomical evidence that the brain was a heterogeneous structure, a theory first

postulated by Willis in the seventeenth century [729]. He then suggested that nerves were

bundles of heterogeneous neurones that have specialized functions, and that their bundling was

only for ease of distribution. Thus, Bell spoke of different sensory neurones for different types of

stimuli, motor neurones, and so-called vital neurones that are wired to the mind, rather than the

brain. He did, however, maintain that perception of stimulus (such as vision and nociception) is

different than the perceptual experience (e.g., colour and pain, respectively). Importantly for

specificity theory, Bell states:

…that while each organ of sense is provided with a capacity of receiving certain changes; to be played upon it, as it were, yet each is utterly incapable of receiving the impressions destined for another organ of sensation. It is also very remarkable that an impression made on two different nerves of sense, though with the same instrument, will produce two distinct sensations; and the ideas resulting will only have relation to the organ affected [79].

This is the fundamental tenet of specificity theory, which postulates that there is a dedicated fibre

that leads up a dedicated pain pathway to the sensory modality’s region of the brain. This model,

therefore, suggests that a pathway specific to pain exists.

François Magendie was a French physician considered by some as the father of experimental

physiology [87,773,827]. Magendie made substantial contributions to neurophysiology,

including reiterating Bell’s findings regarding the existence of both motor and sensory nerves,

and that these have separate paths to and from the spinal cord (the ventral and dorsal roots,

respectively) [827]. This differentiation of spinal nerves is known as the Bell-Magendie Law,

which is a fundamental aspect of the organization of the nervous system.

9

Concurrently in Germany, Johannes Müller published a Manual of Physiology, which echoed

Charles Bonnet’s manual published a century earlier [729]. Müller’s manual, published in 1840,

sought to summarize and synthesize findings in physiology. The purpose of this synthesis was to

understand how different stimuli were so clearly sensed and how the brain could distinguish

them from one another. He, like Bonnet, concluded that specific receptors have must have

specific energy of stimulation, and that there were infinite numbers and types of fibres, each to a

specific sensory stimulus; e.g., there is a specific fibre for the smell of bananas, and another for

the scent of an apple, and yet another for the scent of an orange. Furthermore, because of a sense

organ’s specific energy, the sensory neurone will only encode a single perceptual quality. For

example, if a warm fibre is artificially stimulated by electric current, the brain will perceive

warmth. In line with these findings, Erasmus Darwin (Charles Darwin’s grandfather) provided

the first evidence for a set of specific nerves for the perception of heat [204].

The discovery of specific cutaneous touch receptors such as Pacinian corpuscles [148],

Meissner’s corpuscles [149], Merkel’s discs [394], and Ruffini’s end-organs [393] in the latter

half of the nineteenth century provided further evidence that specific sensory qualia were

encoded by dedicated nerve fibres (see Section 2.3.1). However, there remained a debate about

the nature of pain as part of the five senses. An end-organ specific to pain stimuli (nociceptor)

had not yet been discovered. In contrast to the idea of a dedicated pain pathway, it was argued

that pain was different than the other senses in that it is inherently unpleasant [106,199]. These

ideas persisted from Plato and Aristotle’s writings of pain as an emotion [767]. This inherently

makes pain the antithesis of pleasure, and, because pleasure is a characteristic of the mind (i.e.,

an emotion), it was inferred that pain was also a characteristic of the mind, and not a percept of

the body.

Further evidence for the specificity theory came from Schiff’s and Woroschiloff’s findings of a

pain pathway in the spinal cord in a series of experiments between 1854 and 1859. These

findings built upon Charles-Édouard Brown-Séquard’s observations that sensory fibres decussate

in the spinal cord [18,199]. Schiff and Woroschiloff established the presence of two pathways

through observations of the effect of incisions at different levels of the spinal cord: the

anterolateral pathway for pain and temperature, and the posterior bundles for tactile sensibility

[199,729]. However, Schiff and Woroschiloff noted that the tactile pathway did not decussate at

the level of the spinal cord. These findings were supported by a case study reported by William

10

Richard Gowers, a physician in London. In this report, Gowers reported that a patient with a

bullet wound to the gray matter of the spinal cord lost the sense of pain and temperature, but not

touch [729]. He concluded that there were specific pathways for pain and temperature, separate

from that of touch. However, those who held onto the Aristotelian dogma argued strenuously

against specificity theory. They insisted that pain is a quality of all senses, a percept of the mind.

Only when Blix and Goldscheider published their findings of sensory spots on the skin

independently did specificity theory gain momentum, and did pain become a recognized sense

[199]. Sensory spots were defined as tiny areas of the skin that elicit a specific sensation when

touched. These sensory spots were specific to warmth, cold, pressure or pain. However, both

Blix and Goldscheider moved away from specificity theory some years later, and moved towards

the intensity theory of pain, a concurrent theory (see section 2.1.1.2).

Between 1894 and 1896, Max von Frey’s carried out experiments that advanced the specificity

theory. Von Frey indicated that there were four somatosensory modalities: cold, heat, pain and

touch and that all of the other skin senses were derivatives of these four modalities. To test this

idea, he developed his now well known von Frey hairs (termed an aesthesiometer) that consisted

of a hair, usually from a human, but sometimes he used a horsehair or a hog bristle, attached to a

wooden stick [682]. By measuring the hair’s diameter, length and the maximal weight precisely

it could support (maximal tension) without breaking off of the stick, it was possible to measure

the force applied to a very specific spot. Today, von Frey hairs are made of fine nylon filaments

of varying thicknesses (and hence stiffness) to deliver different forces and pressures upon

bending. Using these hairs, he could carefully determine the pressure required to elicit a

sensation at each of the skin spots identified by Blix and Goldscheider. Further, his experimental

setup allowed him to determine which spots responded to innocuous pressure, and which ones

responded noxious pressure. Von Frey demonstrated that there were distinct spots for innocuous

pressure and for noxious pressure. He presented a model of the skin that comprised of a “mosaic

of distinct tactile, cold, warm, and pain spots distributed across the skin with distinctive regional

variation” [683]. Von Frey related the distribution of the pressure points to the distribution of

Meissner’s corpuscles (see section 2.3.1), whereas pain points were related to the distribution of

free nerve endings in the skin (see section 2.3.3.1). Despite these remarkable findings, the

specificity theory made a number of assumptions about the anatomical basis of somesthesis. For

instance, when von Frey postulated the theory, pain receptors had yet to be identified, nor were

11

the peripheral pathways and brain centers specific to pain sensation established, as well as other

factors (for a review, see [199,729]).

Charles Scott Sherrington addressed some of the assumptions of the specificity theory in his

proposed framework of nociception. He applied a Darwinian (i.e., evolutionary) approach to

describe the specificity of neurones, which included the four basic modalities recognized by von

Frey. He stated that “the main function of the receptor is […] to lower the excitability threshold

of the arc for one kind of stimulus and heighten it for all others” [796,797]. This approach

resolved the divide between the intensity theory (see below) and specificity theory [729]. He also

coined the term “nocicipient” [795] to describe the specificity of the cutaneous end-organ for

pain, later termed nociceptor [796]. Sherrington developed a framework that advanced the

specificity theory of pain even further. However, the nociceptor had yet to be identified

definitively.

The discovery of myelinated primary afferent fibres that respond only to mechanical noxious

stimuli occurred much later – in 1967 [124]. Soon thereafter, Bessou and Perl [89] discovered

nociceptive unmyelinated afferent fibres: polymodal nociceptors and high-threshold

mechanoreceptors (HTM). These findings revolutionized the field of pain research, and helped

advance and develop a number of theories of pain. Since Sherrington’s endorsement of the

specificity theory of pain, this became the dominant theory at the time. However, its popularity

waned with the postulation of the Gate Control Theory of pain (see Section 2.1.1.4) postulated

by Melzack and Wall [595].

2.1.1.2. Intensity theory of Pain

The intensive (or summation) theory of pain (now referred to as the intensity theory) has been

postulated at several different times throughout history. First postulated in the 4th century BCE

by Plato in his œuvre Timaeus [698], the theory defines pain not as a unique sensory experience,

but rather as an emotion that occurs when a stimulus is stronger than usual. Centuries later,

Erasmus Darwin reiterated this concept in Zoonomia [204]. One hundred years after Darwin,

Wilhelm Erb also suggested that pain occurred in any sensory system when sufficient intensity

was reached, rather than being a stimulus modality in its own right [199]. Arthur Goldscheider

further advanced the intensity theory, based on an experiment performed by Bernhard Naunyn in

1859 (cited in: [199]). These experiments showed that repeated tactile stimulation (below the

12

threshold for tactile perception) produced pain in patients with syphilis who had degenerating

dorsal columns. When this stimulus was presented to patients 60-600 times per second, they

rapidly developed what they described as unbearable pain. Naunyn reproduced these results in a

series of experiments with different types of stimuli, including electrical stimuli. It was

concluded that there must be some form of summation that occurs, in order for the subthreshold

stimuli to become unbearably painful. Goldscheider’s suggested a neurophysiological model to

describe this summation effect: repeated subthreshold stimulation or suprathreshold hyper-

intensive stimulation could cause pain. Goldscheider suggested that the increased sensory input

would converge and summate in the gray matter of the spinal cord. This theory competed with

the specificity theory of pain, which was championed by von Frey. However, the intensity theory

lost support with Sherrington’s evolutionary framework for specificity theory; and postulated the

existence of sensory receptors that are specialized to respond to noxious stimuli, for which he

coined the term “nociceptor”.

2.1.1.3. Pattern Theory

In an attempt to overhaul theories of somaesthesis (including pain), J. P. Nafe postulated a

“quantitative theory of feeling” [636]. This theory ignored findings of specialized nerve endings

and many of the observations supporting the specificity and/or intensive theory of pain. The

pattern theory stated that any somaesthetic sensation occurred by a specific and particular pattern

of neural firing, and that the spatial and temporal profile of firing of the peripheral nerves

encoded the stimulus type and intensity. Lele, Sinclair and Weddell [531] championed this

theory, and added that cutaneous sensory nerve fibres, with the exception of those innervating

hair cells, are the same. To support this claim, they cited work that had shown that distorting a

nerve fibre would cause action potentials to discharge in any nerve fibre, whether encapsulated

or not. Furthermore, intense stimulation of any of these nerve fibres would cause the percept of

pain [807,924].

2.1.1.4. Gate Control Theory of Pain

In 1965, Ronald Melzack and Patrick (Pat) David Wall proposed a theory that would

revolutionize pain research: the Gate Control theory of pain. The Gate Control theory recognized

the experimental evidence that supported specificity theory and pattern theory, and provided a

model that could explain these seemingly opposed findings. In a landmark paper, Melzack and

13

Wall [595] carefully discussed the shortcomings of the specificity and pattern theories – the two

dominant theories of the era. The Gate Control theory attempted to bridge the gap between

intensity and specificity theory with a framework based on the aspects of each theory that had

been corroborated by physiological data. Specifically, Melzack and Wall accepted that there are

nociceptors (pain fibres) and touch fibres, and proposed that that these fibres synapse in two

different locations within dorsal horn of the spinal cord: the substantia gelatinosa and the

“transmission” cells. The model (see Figure 2-2) proposed that signals produced in primary

afferents from stimulation of the skin were transmitted to three regions within the spinal cord: (1)

the substantia gelatinosa, (2) the dorsal column, and (3) a group of cells that they called

“transmission” cells. They proposed that the gate in the spinal cord is the substantia gelatinosa in

the dorsal horn, which modulates the transmission of sensory information from the primary

afferent neurones to “transmission cells” in the spinal cord. This gating mechanism is controlled

by the activity in the large and small fibers. Large fiber activity inhibits (or closes) the gate,

whereas small fiber activity facilitates (or opens) the gate. Activity from descending fibres from

supraspinal regions to the dorsal horn could also modulate this gate. When nociceptive

information reaches a threshold that exceeds the inhibition elicited, it “opens the gate” and

activates pathways that lead to the experience of pain and its related behaviours. Therefore, the

Gate control theory of pain provided a neural basis for the findings that supported and in fact

helped to reconcile the apparent differences between the pattern and specificity theories of pain.

14

Figure 2-2: Schematic of the Gate Control Theory of Pain mechanisms. L – large diameter fibres; S – small diameter fibres. The fibres project to the substantia gelatinosa (SG) and first central transmission (T) cells. The inhibitory effect exerted by SG on the afferent fibre terminals is increased by activity in L fibres and decreased by activity in S fibres. The central control trigger is represented by a line running from the large-fibre systerm to the central control mechanisms; these mechanisms, in turn, project back to the gate control system. The T cells project to the entry cells of the action system. +, Excitation; –, inhibition. (Figure and legend are reproduced with permission from: [595]).

15

2.1.2. Multidimensional aspects of pain

Melzack and Casey [593] described pain as being multidimensional and complex, with sensory-

discriminative, affective-motivational and a cognitive-evaluative components. Additionally,

recent work has shown that pain can also affect and interact with motor systems [23]. The

concept of pain as a multidimensional experience has been described in ancient texts, dating as

far back as the Syriac Empire (circa 200 BCE). In the Book of Medicines [122], it is suggested

that pain is the product of bile and phlegm mingled with cold and heat. These simple

combinations occur in the brain, and, according to Syriac medicine, pain is a product of the brain

(a concept which has passed the test of time, and that we still hold true today). Different types of

pains would thus arise from differential combinations of these substances affecting the type of

pain. It is noteworthy that the concepts of bile and phlegm, and even those of cold and hot were

understood in a different paradigm of philosophical thought – these are not the simple

compounds we know today, but are rather used as a classification of the world. For instance,

certain foods make the body “cool”, whereas others make the body “warm.” These concepts are

not unique to the Syrians, since they follow a long tradition of ancient medicine passed down

from the Egyptians (who were the first to record medical texts, e.g., The Papyrus Ebers [120]), to

the Greeks (e.g., most-famously, Hippocrates and Galen), to the Babylonians and to the

Assyrians.

The contemporary definition used by the IASP is based on the divisional (multidimensional)

definition proposed by Melzack and Casey, in 1968 [593]. These dimensions include the

sensory-discriminative (intensity, location, quality, and duration), the affective-motivational

(unpleasantness and the subsequent flight response) and the cognitive-evaluative (appraisal,

cultural values, context and cognitive state) dimensions of pain. These three dimensions are not

independent, but rather, interact with one another. They are, however, partially dissociable: the

cognitive state of a person can modulate one or both of these dimensions of pain perception. In

general, the more intense a noxious stimulus is, the more unpleasant it will be [279]. However,

there are exceptions to this rule: hypnosis has been shown to modulate pain unpleasantness

without affecting intensity – that is, the person felt the pain, but was not as bothered by the

sensation [500,590,714-716,937]. This is an example of how the cognitive state can modulate the

percept of the affective-motivational component of pain and can be referred to as cognitive

16

modulation. Cognitive modulation of pain is reflected in the effects of placebo, nocebo

[171,172,295,908], cognitive behavioural therapy [752,753,793] and other treatments for chronic

pain.

Theories about chronic pain have continued to evolve as knowledge accumulates concerning the

structure and function of pathways underlying pain perception and pain modulation. Recent

advances in neuroimaging and cellular and molecular medicine have vastly expanded our

understanding of pain, and some of these key findings relevant to this thesis will be discussed

further when anatomical pathways and brain regions implicated in pain are described (see

Sections 2.3).

2.1.3. Chronification of pain

An important issue that is not fully understood is the way in which acute pain becomes chronic.

For example, how does pain change from being a beneficial signal of a potential threat that

engages protective mechanisms, to become a state where pain itself becomes a threat,

dysfunction, despair, and much suffering.

2.1.3.1. Sensitization, Hyperalgesia and Allodynia

When the skin is injured and pain is evoked, the tissue can in some situations become tender and

inflamed. The tenderness can be associated with two processes: allodynia and hyperalgesia.

Allodynia is defined as a “pain due to a stimulus that does not normally provoke pain” [110]. The

definition further specifies that allodynia is not related to any specific mechanism; rather it

specifies an unexpected behavioural response (pain) to a normally innocuous stimulus.

Furthermore, to demonstrate allodynic responses, it is preferred that responses be tested in

unaffected body areas to show normal responses, where possible. Because the normative

response to the stimulus is not painful, and the allodynic response is painful, “allodynia involves

a change in the quality of the sensation” [110]. Hyperalgesia is defined as a “increased pain from

a stimulus that normally provokes pain” [110]. There are two types of hyperalgesia: primary (at

the injury site) and secondary (outside the site of injury). These types of hyperalgesia can be

described with the following example. When glabrous skin (non-hairy skin, like the palm of the

hand) is burned, the burn spot (primary zone) becomes more sensitive to heat and mechanical

stimuli – a stimulus that was mildly painful before the burn is subsequently perceived as being

17

much more painful [609]. In normal cases, it is believed that allodynia and hyperalgesia provide

warning signs, and are likely related to nocifensive behaviours – that is, behaviours that are

meant to protect the damaged limb, prevent further damage, and ultimately promote healing.

The perceptual states of hyperalgesia and allodynia are related to the physiological process

known as sensitization, which is defined as “increased responsiveness of nociceptive neurones to

their normal input, and/or recruitment of a response to normally subthreshold inputs” [110].

There are different mechanisms of sensitization, depending on the type of tissue and/or injury

(chemical, mechanical or heat) [610]. This sensitization process can result from threshold related

changes in nociceptors (i.e., lower firing threshold to sensory stimuli and/or an increased firing

rate to suprathreshold stimuli, spontaneous firing and/or an expansion of the receptive field).

Evidence suggests that primary hyperalgesia occurs as a result of nociceptors becoming

sensitized [609,853]. Similarly, evidence supports that allodynia in the primary zone is also

related to sensitization. Peripheral sensitization occurs as a result of the prolonged activation of

peripheral nociceptors, or exposure to the algesic chemicals, which leads to the release of a

number of chemical messengers that reduce nociceptive neurones’ threshold (for a

comprehensive review, see: [515,951]). This process of sensitization causes amplification of the

response to a noxious stimulus as well as to non-noxious thermal and mechanical stimuli. In the

example of the burn described above, the region around the burn (secondary zone) is more

sensitive to mechanical stimuli, but not to heat stimuli [718]. In fact, the secondary zone is less

sensitive to heat stimuli than normal skin. Peripheral sensitization does not fully explain

secondary hyperalgesia and there is now much evidence that secondary hyperalgesia is largely

mediated by central mechanisms [512], although there may also be some peripheral contribution.

Specifically, neurones in the CNS show enhanced responses to mechanical and thermal stimuli in

the secondary zone [806], i.e. a central sensitization. The IASP defines central sensitization as an

“increased responsiveness of nociceptive neurones in the central nervous system to their normal

or subthreshold afferent input.” These abnormal neuronal states are thought to be, in part, the

neural basis for hyperalgesia and allodynia. This state of hyperalgesia is likely related to

dysfunctional endogenous pain-modulatory mechanisms (see section 2.3.6).

18

2.1.3.2. The role of glia in sensitization

There are many more glial cells in the nervous system than there are neurones. Classically, it was

believed that glia only served as support or structural cells. However, recent evidence has

suggested that these cells play a much more significant role in both the PNS and CNS

processing. In the context of pain, recent evidence suggests that glia are important to the

pathogenesis of pain [159]. Specifically, satellite glial cells in dorsal root ganglia and trigeminal

ganglia are implicated in nociceptive conduction and transmission, and microglia and astroglia in

the CNS are implicated in sensitization and chronification of pain [159]. However, microglia and

astroglia have no effect on normal nociceptive processing, they are only implicated in

pathological states [614,921]. Specific mechanistic details of the role of glia in sensitization are

outside of the scope of this thesis (for a comprehensive review, see: [159]). Microglia are also

activated after inflammation, and are implicated in CNS inflammatory processes [159].

Taken together, it is possible that prolonged nociceptive input may induce central sensitization, a

process that is mediated by neurones and glia. Over time, this sensitization can lead to changes in

the function and structure of the nervous system (i.e., plasticity – see section 2.6). Normally,

there is a delicate balance between nociceptive and antinociceptive signaling in the central

nervous system. For instance, persistent nociceptive input is accompanied by an increase in

activity in pain-modulatory pathways. This functional change is considered a form of beneficial

plasticity, or change in the nervous system (see Section 2.6). However, in some cases plasticity

can be harmful, or maladaptive. For instance, it is believed that in persons who develop chronic

pain, the supraspinal modulatory pathways become dysfunctional or ineffective. This

maladaptive plasticity can contribute to the development and/or maintenance of chronic pain,

even in the absence of a noxious stimulus [576] (see Section 2.6.2). Furthermore, pain sensitivity

and a reduction in the ability to modulate pain may be related to pre-existing abnormalities in the

CNS.

2.1.4. Functional pain syndromes

Pain without a known physiological or “organic” cause is diagnosed as functional pain disorders.

Functional pain syndromes include the most prevalent of all pain conditions, which can be

divided into two categories: somatic or visceral pain syndromes. Somatic pain syndromes include

TMD, vulvodynia, low-back pain and fibromyalgia (FMS). Visceral pain syndromes include

19

irritable bowel syndrome (IBS), chronic cardiac pain syndrome, and internal cystitis (or painful

bladder syndrome). Functional pain syndromes are prevalent in 15-20% of adults worldwide

[562]. Relevant to this thesis is the prevalence of TMD, which is the most prevalent of these

functional pain syndromes (discussed in more detail in Section 2.2.3).

2.1.4.1. Central abnormalities in functional pain syndromes

Functional pain disorders are more prevalent in women than in men [581,582]. Furthermore,

some functional pain disorders are more likely to have increased co-morbidity in women than in

men [582]. Functional pain syndromes, such as TMD, are diagnosed as such because of a lack of

a clear physiological cause for the pain. This means that there are no consistent findings for the

source of the pain, and there are no clear peripheral abnormalities. Functional pain syndromes

do, however, share some common pathophysiological features, including changes to

neurobiological, physiological and anatomical aspects of the CNS. A number of behavioural

studies have also shown that patients with functional pains are hypersensitive to experimental

pain [581]. These patients also have dysfunctional pain-modulatory systems (e.g., [491,693]),

and, in some cases, have cognitive deficits (e.g., [365,412,455,489,666]). There is also a greater

degree of comorbidity with mood and anxiety disorders in patients suffering from functional pain

disorders [637]. With the lack of peripheral aetiological factors, these behavioural abnormalities

have led researchers to suspect that central mechanisms contribute to and mediate functional pain

syndromes. Findings specific to idiopathic TMD are described in further detail in section 2.2.3,

and support the hypothesis that central sensitization may be the underlying pathophysiological

processes underlying functional pain syndromes. These hypotheses have been corroborated by a

number of functional and structural neuroimaging studies, and are described in further detail in

Section 2.5.2.

2.1.5. Pain-cognition interactions

There is a two-way interaction between cognitive processes and pain in that cognitive processes

can interfere with and modulate pain, and pain can interfere with and modulate cognitive

processes (e.g., [123,193,290,529,776,778,781,893,935]). The interaction is best illustrated by a

common example: that of a football player who injures his ankle, but doesn’t feel the pain until

after the game. This suggests that pain can be attenuated when a person is distracted or engaged

in a cognitively-demanding task [851]. Conversely, pain can hamper cognitive performance (e.g.

20

reduced task accuracy or slower task reaction times) [713]. The cognitive interference of pain

may be due to the possibility that pain has a cognitive load, and that there could be limited

cognitive resources in the brain. Therefore, the brain must prioritize between competing tasks,

such as pain and a cognitive task. More recent work has demonstrated that cognition-pain

interactions are the product of complex interactions between two modes of attention selections:

bottom-up and top-down [529]. The bottom-up attention selection is related to the salience of

pain in that pain can be attention demanding because it is behaviourally relevant. Interestingly,

the salience of pain is partly relevant to its novelty [530,632,886]. This process of habituation is

likely related to top-down attention selection processes. These processes allow the redirection of

attention to other behaviourally relevant stimuli, and thus reduce the attention-grabbing (or

salience) effect of pain [529,892].

Several seminal studies in the 1990s demonstrated the interaction of pain and cognitive

performance in healthy subjects, showing that their task performance decreased (i.e., longer

reaction times) when concurrently presented with a noxious stimulus [192-194]. These studies

indicate that pain can interrupt or detract from the ability to perform a cognitive task, and thus

suggests that it is difficult to disengage from the painful stimulus. Later studies showed that

difficult tasks, which required presumably more cognitive resources, could briefly reduce

subjective pain ratings. Therefore, pain can affect cognitive performance, and cognitive

performance can modulate pain ratings (for a comprehensive review see [778]). These processes

have been summarized in the neurocognitive model of attention to pain [529]. This model

describes a “bottom-up capture of attention by pain”, whereby pain grabs one’s attention to

signal a relevant and threatening stimulus. However, this bottom-up process can be modulated by

top-down factors, such as cognitive load, emotional state and motivation. Inherent to this form of

modulation is the context of the noxious stimulus. If a subject is in an emotional state he/she may

report heightened pain, as he/she is more focused on stimuli with a negative valence.

Alternatively, a subject may report a heat stimulus as less painful if the stimulus is perceived as

beneficial, such as a heating pad, or a hot tub in a therapeutic setting.

In the context of chronic pain, a number of studies have shown that chronic pain negatively

impacts cognitive abilities, including concentration, learning, memory, and attentional focus

[26,250,275,365,369,455,489,666,928]. Eccleston and colleagues [289] showed that chronic pain

patients could briefly be disengaged from pain during attentional switches. They further showed

21

that pain has a cognitive load, and that only when there is heavy use of cognitive resources, do

pain ratings decrease. Taken together, these findings suggest that the brain has a limited

cognitive resource, where a more salient task can overcome pain – however, only for a brief

period of time. Over time, patients report getting tired, and not being able to focus on the task,

and reducing the subject’s ability to block out pain from consciousness. In a later study,

Eccleston and colleagues [292] showed that the cognitive interference effect of chronic pain was

related to levels of negative affect and somatic awareness. In fact, it has been shown that

negative affect can enhance perceived pain [936].

A study by Harman and Ruyak [409] reported that people with persistent low-level pain have

cognitive deficits, indicative of the interference of pain. There is evidence of functional and

structural plasticity that occurs in chronic pain (see Section 2.5.1 and 2.5.2, respectively and

2.6.2): over time, pain can change the connectivity of brain regions, which can lead to cognitive

impairment. Another possible mechanism is the aforementioned limited resource model – the

cognitive load of pain reduces resources to perform a task. These mechanisms are not mutually

exclusive, as prolonged nociceptive input can alter the connectivity of cognitive regions, and

further reduce resources available for cognitive processing.

In line with the limited resource model, is the concept of cognitive branching – the ability to put

a pending task “on hold” to execute a more salient (or rewarding) task [497]. Therefore, the

cognitive branching model provides a functional basis for the limited resource model, and

provides a model that explains reductions in pain ratings when engaged in a cognitive task, or

reduced cognitive performance whilst receiving prolonged nociceptive input. Specifically, if the

cognitive branching system is limited, as is suggested by Koechlin and Hyafil [497], then the

system can be overloaded, and affect task performance. This explains findings that tasks that

require little cognitive engagement do not modulate pain ratings, and that low noxious stimulus

intensities do not have a significant effect on task performance [776]. Conversely, as cognitive

processes increase in difficulty, pain ratings can be modulated; or if nociceptive input increases,

cognitive performance can be negatively affected. For example, Weissman-Fogel and colleagues

[928] found no group differences in task performance between controls and patients with TMD

in a neutral Stroop task (no interference). However, when presented with a more difficult

interference counting Stroop task, patients showed sluggish reaction times – their cognitive

branching system was either overwhelmed, or prioritized pain over the Stroop task. Finally, this

22

effect was modulated by an emotional counting Stroop task. Subjects were asked to report the

number of TMD-pain related words visually presented. Specifically, these words were: ache,

tight, annoying, pain, stress, hurt, clench, anxious, throb, unpleasant, pressure, and tiring. The

patient group reported that the TMD pain words affected their pain intensity and unpleasantness

during the Stroop task, whereas controls did not. The increase of the patients’ pain rating affected

the TMD group’s reaction times for the Stroop task.

In sum, pain and cognition have a complex interaction that can be modulated by affective state,

context and motivation. Pain is inherently salient, but its behavioural relevance can be up- or

downregulated based on these factors. Furthermore, chronic pain can alter brain structure,

connectivity and function, which may further impact cognitive abilities.

2.1.6. Salience and pain

Pain plays an essential role in survival – it signals a potential or actual injury. Because of the

behavioural relevance of pain, noxious stimuli are inherently salient. Furthermore, studies have

shown that consciousness is required to perceive pain. Specifically, these studies have

investigated the neural correlates of nociception in subjects in a persistent vegetative state using

neuroimaging techniques (these techniques are discussed in detail in Section 2.5). We must

clearly indicate that these studies were not investigating pain (recall that the definition of pain

requires that a person report pain – it is a subjective perceptual experience), and so, these studies

were comparing neural correlates of normally noxious stimuli in persistent vegetative state

patients to pain perception in the control group. One study found that the regions along the

ascending nociceptive pathways, involved in the sensory-discriminative dimension (midbrain,

thalamus and primary somatosensory cortex (S1)), showed activation, but the other regions that

showed activation in the control group (including the secondary somatosensory cortex (S2),

insula, cingulate, prefrontal cortex (PFC) were not activated [516]. Conversely, Kassubek and

colleagues [479] found that the patients did have activation in brain regions involved in

nociception and pain modulation. However, it is worth noting that recent studies have shown that

some, but not all, patients in a vegetative state show brain activity in response to stimuli in the

environment [517,659]. Therefore, it is possible that in a subset of these patients, where

consciousness may persist, pain perception and modulation occurs. However, in patients that

23

have lost consciousness, nociception occurs [768], but there is no evidence that they feel pain, as

regions underlying different dimensions of pain do not show activation.

2.1.7. Pain-motor interactions

Pain-motor interactions are a hallmark of pain-related behaviours, the most obvious example

being nocifensive behaviours. Descartes portrayed nocifensive behaviour in the Treatise of Man

[242] in his famous depiction of the boy burned by the fire, and Charles Darwin describes

nocifensive behaviours in his text Expression of Emotion in Man and Animal [203]. There is no

doubt that the sensory and motor system interact, a concept supported by the many sensorimotor

reflexes observed in animals and humans. However, the interactions of pain and motor functions

are not as clear. Furthermore, the effects of prolonged (or chronic) pain on motor output are

equally ambiguous. A number of experimental pain paradigms have attempted to shed light onto

these topics. For instance, it has been shown that noxious input inhibits the motor cortex

excitability [3,157,301,302,638,856]. Similarly, motor cortex stimulation has been shown to

modulate pain [103,117,118,301,331,342,344,345,535,548,657,689]. Furthermore, a study by

Boudreau and colleagues [109] reported that motor cortex plasticity is inhibited by noxious

stimulation of the intraoral cavity. Therefore, it seems that motor learning is modulated, in this

case inhibited, by a tonic noxious stimulus.

A number of theories have been postulated to explain the relationship between pain and motor

output. The two dominant theories are the vicious cycle theory, first postulated by Travell [866],

and the pain adaptation model, postulated by Lund and colleagues [556]. The vicious cycle

theory suggests that the presence of a reciprocal relationship between pain and dysfunction: pain

causes muscular dysfunction, which increases pain, creating a vicious cycle. This theory states

that, over time, the muscle hyperactivity contributes to and/or maintains pain. Specifically,

tonically hyperactive muscles become ischaemic, which in turn leads to an increase in algesic

chemicals within the musculature. This has been supported by findings of increased motor

activity in experimental models of muscle pain [431]. However, some studies of orofacial pain

have failed to demonstrate that pain leads to muscular hyperactivity [634,675]. Therefore, Lund

and colleagues [556] postulated the pain adaptation model, which suggests that there is both

muscle hyper- and hypo-activity. Specifically, the model states that there is a decrease in activity

of muscles that are painful or that hurt with movement, and that antagonist muscles have

24

increased activity. However, more recent experimental findings have not supported the concept

of uniform inhibition of a painful muscle. Instead, they have shown variable patterns of change

in activity both spatially and temporally (for a review, see: [431]). To address these

inconsistencies, Hodges and colleagues propose a “new theory of motor adaptation to pain”

[431], which maintains the basic premise that pain modifies muscular activity as a nocifensive

mechanism. This theory proposes that the body adapts to muscular pain is several ways. When

there is pain or threat of pain, there is a redistribution of activity within a muscle, and between

muscles, so that force is maintained. This redistribution occurs upon the recruitment of new

motor units, and changes the mechanical behaviour of the muscle including increased stiffness

and altered load distribution. In the short term, this reduces movement of the muscle and

decreases the load on the muscle, which is beneficial and promotes healing. In the long term,

however, these changes can lead to increased load, decreased movement and decreased

variability, which can increase the risk of injury.

2.1.8. Neuroticism and pain

Psychological factors have been associated with sensitivity to pain, and the development and

maintenance of chronic pain. Inherent personality-related factors reduce the brain’s capacity to

effectively modulate nociceptive input and pain perception, likely affecting the cognitive

modulation system (discussed in Section 2.1.2 and Section 2.3.6). Of particular interest to this

thesis is the personality dimension of neuroticism [178]. Neuroticism has been defined as a stable

personality trait associated with heightened sensitivity and/or processing of negative affective

stimuli [178,179,907]. Therefore, it is not surprising that patients with high neuroticism scores

have a lower pain threshold and pain tolerance, and are more distressed by pain in experimental

settings and chronic pain [408]. Specifically, neuroticism has been associated with pain-related

suffering [35,408], pain sensitivity [177,373,907], nerve injury outcomes and the development of

neuropathic pain [848], and TMD [315]. However, not all chronic pain patients have high

neuroticism scores, and not all persons with neuroticism have chronic pain [179]. Therefore,

neuroticism alone is not sufficient to develop chronic pain; however it can contribute to the

severity of the pain reported by chronic pain patients.

25

2.2. Temporomandibular disorders

2.2.1. Historical perspectives

Reports of TMD-like symptoms date back 1500 years when Hippocrates described a method for

reducing dislocated mandibles [423]. These treatments often led to TMD-like disorders,

including arthritis and infection of the TMJ. The first documented surgeries for disc

displacements in the TMJ occurred in the nineteenth century, and shortly thereafter, articular disc

removal was used to manage TMD symptoms [587]. In the late 1800s and early 1900s, dentists

primarily attributed TMD symptoms to problems with the muscles of mastication and their

contribution to malocclusion. By 1934, malocclusion was formally considered to be the cause of

TMD symptomology [181]. Costen’s range of symptoms, that far exceeds the recognized

pathophysiological understanding of TMD today, led to the development of numerous treatments

and devices to treat TMD and other symptoms believed to be related to malocclusion (for a

review of these, see [587]). However, this range of symptoms was criticized for being too broad

[587]. More recent, randomized control research studies have refuted the use of occlusal devices

in the treatment of TMD, suggesting that the positive effects of these sorts of treatments are

related to the patients’ expectation of the treatment [174,882]. Therefore, it has been suggested

that less invasive and less expensive methods of treatment may be as efficacious in pain

management as occlusal splints [872]. Despite an exceptional historical presence and extensive

research, little is known of TMD pathophysiology, and few treatments are effective. Today,

TMD continues to be a prevalent pain syndrome that poses a significant financial burden on

healthcare and society.

2.2.2. Prevalence and social costs

TMD are estimated to affect between 3-20 percent of the United States’ adult population

[269,539,540,562]. Of these, only a small proportion of persons suffering from TMD pain seek

treatment and end up at tertiary and quaternary medical facilities such as pain clinics [350].

TMD are 1.5-9 times more prevalent in women [113,201,284,285,312,721]; a study found that

84% of persons with TMD were women in a nationwide epidemiological study in the United

States [284]. It has been estimated that 5.3 million people in the United States seek medical

attention for a TMD-related problem per year [349]. Furthermore, TMD physical symptoms

(e.g., pain) and psychological symptoms (e.g., depression) have been reported as more severe in

26

women than in men. The predominance of TMD symptoms in women has been attributed to

differences in sex hormones, such as oestrogen, as well as the use of oral contraceptives. Older

(55-74 years old) and younger (< 20 years old) subjects report fewer TMD-like symptoms than

subjects in reproductive years (20-55 years old). The annual incidence of TMD is 2-4%, with

0.1% going on to develop chronic pain. There are different disease outcomes for different

patients, with 31% of cases persisting over 5 years, 33% whose pain resolves, and 36% who have

recurring TMD. The direct healthcare costs for treatment of TMD in the United States have been

estimated at $2 billion. However, TMD is often co-morbid with other disorders, such as FMS,

IBS, anxiety, and depression, and may have greater economic impact based on indirect costs

(such as missing work, unemployment, disability, limitations, etc.). The estimated total of both

direct and indirect costs in the United States exceeds $4 billion dollars per annum [350]. Taken

together, then, it is clear that TMD are a prevalent disorder with a significant economic burden

on society.

2.2.3. Clinical aspects

TMD comprise of a cluster of clinical problems involving the structures of, and around, the TMJ,

the masticatory musculature, or both [1,418]. TMD are primarily characterized by spontaneous

pain and/or pain associated with jaw function, which can affect mandibular range of motion and

produce TMJ sounds (e.g., crepitus, clicking) during jaw function. TMD pain can also spread to

the muscles of the neck. The most common diagnostic criteria (and classification scheme) for

TMD is the TMD-Research Diagnostic Criteria (RDC) established by Dworkin and LeResche in

1992 [285]. The American Association of Orofacial Pain (AAOP) has adopted this classification

scheme. Based on the TMD-RDC, there are two primary types of painful TMD: myogenous

(pain generated by the masticatory muscles) and arthrogenous (pain generated by the TMJ)

[285]. Myogenous TMD, often called masticatory muscle pain, is diagnosed on the basis of

patient history and clinical examination. Arthrogenous TMD conditions are usually caused by

displacement or derangement of the TMJ disc, but also by inflammation within or around the

joint. Hence, the pain in arthrogenic TMD can be attributed to inflammatory processes that occur

as a result of intra-articular events or anatomical factors including disk displacement – although

extra-articular events can also cause TMJ inflammation [646]. Additionally, TMD can be

classified as idiopathic or as post-traumatic. Idiopathic TMD have no clear aetiology for pain

onset and/or maintenance, although, as discussed above, there are a large number of factors that

27

have been considered in the past to play an important aetiological role. However, over time, and

with increasing research, these factors have either been put into question or even eliminated. In

fact, this is why it has been suggested that TMD is generally an idiopathic condition. In contrast,

post-traumatic TMD are caused by traumatic events in the craniofacial region. These traumatic

events can be as serious as head injury, hyperextension-flexion injury (e.g., following a motor

vehicle accident), or as minor as a dental treatment (e.g., a long dental procedure) [81].

2.2.3.1. Sensory abnormalities

Sensory function in TMD has been assessed with a variety of psychophysical measures and

quantitative sensory testing (QST) approaches to determine detection and pain thresholds to

experimental stimuli across thermal, mechanical, electrical and chemical modalities. The details

of these studies are provided in Table 2-1, and a summary of the findings is provided in Table 2-

2. In general, these studies demonstrate that patients with TMD can have sensory abnormalities

such as hyperalgesia, reduced pain tolerance and reduced DNIC.

QST studies have demonstrated that patients with TMD have hyperalgesic responses2 to acute

experimental pain in the trigeminal region as well as in extra-trigeminal regions, as measured by

pain thresholds. Specifically, studies investigating pressure pain thresholds in TMD have

reported that patients with TMD have reduced pain thresholds to noxious pressure stimuli in the

trigeminal region, and outside the trigeminal region [42,139,305,565,841]. Interestingly, one of

these studies did not find significant group differences in pressure pain thresholds between

controls and patients with TMD, but did so when they exposed study participants to a

psychosocial stressor [139]. Most studies investigating heat thresholds in the trigeminal region

and outside the trigeminal region reported that patients with TMD have lower heat pain

thresholds than controls [313,491,563-565,841]. However, two studies failed to find differences

in heat pain thresholds in the arm between the TMD group and controls [113,313]. Also, five

studies have investigated pain thresholds to experimental ischaemic pain in the arm of patients

with TMD versus controls. Three of these studies reported reduced pain thresholds in patients

with TMD versus controls [313,563,564], whereas two studies [113,180] did not identify any

differences between patients with TMD and controls. Patients with TMD also have lower

2 Whether decreased pain thresholds are related to hyperalgesia or allodynia, or both remains a contentious issue.

28

pinprick pain thresholds in the orofacial region [42]. Four studies have investigated electrical

pain thresholds in the orofacial region, and one study examined these thresholds outside the

orofacial region [42,211,391,792]. Of these, only one identified lower pain thresholds to electric

stimuli in the orofacial region [42]. Together, these studies demonstrate that, in most cases,

patients with TMD have reduced sensory thresholds to noxious stimuli across different

modalities. Furthermore, patients with TMD show prolonged pain and increased peak pain in

response to hypertonic saline injected into muscles in the trigeminal and extra-trigeminal regions

[841]. These hyperalgesic responses may be related to either peripheral sensitization or central

sensitization, or both (see Section 2.1.2.1).

In addition to pain thresholds, some studies have investigated pain tolerance in TMD. It is

believed that pain tolerance reflects a more integral view of the pain system, and its various

dimensions. Because pain tolerance reflects “how much” pain a subject can withstand, this

measure not only captures the sensory-discriminative dimension of pain, but also cognitive-

evaluative and affective-motivational dimensions. Pain tolerance data are variable so it is not

clear what mechanistic factors underlie this attribute seen in patients with pain. Specifically, two

studies [42,841] have reported reduced pressure pain tolerance in trigeminal and extra-trigeminal

regions of TMD patients compared to controls. Four studies have investigated ischemic pain

tolerance, and two studies [563,564] reported lower pain tolerance in patients with TMD

compared to controls, and two studies failed to find any group differences [90,313,563,564]. Five

studies investigated heat pain tolerance in TMD, and four of these identified lower pain tolerance

in patients with TMD compared to controls, compared to controls, whereas one of these studies

failed to identify any group differences. Three studies tested pain tolerance to experimental

electrical stimulation in trigeminal and extra-trigeminal regions, and found no differences

between patients with TMD and controls [211,391,792]. In sum, there are variable findings with

regard to pain tolerance in patients with TMD.

Taken together, there seems to be variability in the findings of quantitative sensory testing

studies. This may be due to the heterogeneity of TMD, which was the motivation for the

development of the TMD-RDC. Therefore, it is difficult to compare studies conducted prior to

the establishment of the RDC, or that do not use the RDC to select patients to studies that have

abided by the RDC.

29

Temporal summation is the production of a gradual and progressive increase in pain from the

repeated painful stimuli. A typical protocol to test for temporal summation involved delivering

painful stimuli at 3s inter-stimulus interval [711] in a short timeframe [31]. It is believed that

temporal summation is related to “wind-up” – a process of increased nociceptive firing of dorsal

horn neurones to repeated noxious stimulation, related to central sensitization [915]. Studies have

demonstrated that the degree of temporal summation is related to central hyperexcitability: if the

subjects have central sensitization, they have increased temporal summation. Two studies have

reported augmented temporal summation in patients with TMD, compared to pain-free controls

[42,565]. One study by Sarlani and colleagues reported that women with TMD had greater

temporal summation than men with TMD, and healthy controls [755]. Therefore, these studies

support the hypothesis that patients with TMD may have central sensitization.

In support of central sensitization in TMD, King and colleagues [491] reported that patients with

TMD had dysfunctional endogenous pain modulation systems. The authors used an experimental

method that modulates pain via descending inhibition: a counter-stimulation method. Under

normal circumstances, when a conditioning noxious stimulus is administered to a remote body

part, it activates the descending modulation system. This activation of the modulation pathway

attenuates pain responses to a second focal stimulus, a process called DNIC [519,521]. In King et

al’s [491] study, patients with TMD did not report less pain to a second focal noxious stimulus,

but rather reported higher pain ratings than when the second stimulus was administered without

the conditioning stimulus. Similarly, Maixner and colleagues [563] reported that patients with

TMD could not effectively engage anti-nociceptive mechanisms when given a remote noxious

stimulus, which would be expected to dampen their spontaneous TMD pain. In this regard, it has

been shown that in concert with findings showing that patients with TMD have increased pain

sensitivity outside the orofacial region, they are also unable to modulate pain perception when a

second noxious stimulus is presented.

In sum, there is evidence that patients with TMD have both peripheral and/or central

sensitization, as well as dysfunctional anti-nociceptive systems.

30

Table 2-1: Summary of quantitative sensory testing studies in TMD

Stimulus TMD subtype*

Stimulus laterality Body site n patients

(# females) n controls (# females)

Pain threshold

Pain tolerance

Temporal summation Other Ref

Pressure myofascial Bilateral V1a, V3

b and wrist 20 (20) 20 (20) !

! ! !![305]

Ischaemia combined Leftc

arm 36 (N/A) 80 (N/A)

! ! ! ![313]

Heat forearm "! ! ! !

Heatd myofascial Left palm 14 (14) 28 (28) ! ! ! !!!DNIC!

[491]

Heat Combined

Left forearm

23 (23) 24 (24)

! ! ! !

[565] Left palm ! ! " !Bilateral face ! ! ! !

Pressure Bilateral face and forearm ! ! ! !

Heat Combined Left forearm 64 (64) 23 (23)

! ! ! !![564]

Ischaemia ! ! ! !!Heat combined Left forearm 52 (44) 23 (19) ! ! ! !

[563] Ischaemia ! ! ! !Heat

myofascial Bilateral

masseter and anterior tibialis

22 (16) 21 (15)

! ! ! !!

[841] Pressure ! ! ! !!

Hypertonic Saline

! !

"!peak!pain!"!pain!duration!

Heat combined Left forearm 20 (20) 42 (20) "! "! ! ! [113] Ischaemia Right arm "! "! ! !

31

Stimulus TMD subtype*

Stimulus laterality Body site n patients

(# females) n controls (# females)

Pain threshold

Pain tolerance

Temporal summation Other Ref

Ischaemia combined Non-dominant arm

28 (10) 20 (20) "! "! ! !!

[180]

Pressure myofascial Bilateral

masseter and temporalis muscles and finger

35 (33) 35 (33)

!e ! ! !

[139]

Pressure

arthralgia

Bilateral

TMJ, non-dominant finger

43 (19) 20 (17)

! ! ! !!

[42] Tactilef TMJ and

face ! ! ! !!

Electric

Ipsilateral to pain in TMD, random in controls

TMJ

! ! " !!

Electric myofascial N/A incisor tooth pulp 12 (11) 12 (10) "! "! ! !

[792]

Electric combined Bilateral masseter 30 (19) 31 (18) "! "! ! ! [391]

Electric myofascial Bilateral masseter and forearm

12 (10) 12 (10) "! "! ! !!

[211]

a - Frontalis muscle, which receives afferent input from the ophthalmic branch of the trigeminal nerve (V1) b - Masseter muscle, which receives afferent input from the ophthalmic branch of the trigeminal nerve (V1) c - No information was provided about pain laterality or handedness.

32

d - The study investigated thermal heat pain thresholds with and without a cold water immersion conditioning stimulus e - Group differences only found when a psychosocial stressor was presented f - Tactile stimuli were two von Frey filaments; 5.16g and 84.96g for tactile and pin-prick stimuli, respectively * "combined" refers to patients with both myofascial and articular (joint) pain or a mixed population of myofascial TMD or joint TMD.

Abbreviations: DNIC – diffuse noxious inhibitory controls; TMJ – temporomandibular joint; V1 – ophthalmic branch of the trigeminal

nerve; V3 – mandibular branch of the trigeminal nerve; - no difference; ! – decrease

33

Table 2-2: Number of quantitative sensory testing studies testing sensory thresholds and tolerance in TMD

Data are based on studies presented in Table 2-1. ! indicates a decrease in TMD compared to controls. Blank cells indicate that the measure

was not tested or reported.

Stimulus (n studies)

Trigeminal Region Extra-trigeminal Regions

Threshold Tolerance Threshold Tolerance

!! no change ! !! no

change not

tested ! !! no change ! !! no

change not

tested Electric (4) 1 3 0 3 1 0 1 0 1 3 Ischaemia (5) 5 3 2 2 2 1 Pressure (5) 5 2 3 5 0 2 3 Thermal heat (7) 2 1 1 5 2 4 1 2 Pin-prick (1) 1 0 1 1

34

2.2.3.2. Motor abnormalities

Some studies have shown that patients with TMD have motor abnormalities in the orofacial

region, although there is much discordance across studies. For instance, three studies reported

that patients with TMD have hyperactivity in their muscles of mastication, whereas other studies

did not find any differences when comparing patients to controls (for a review, see [197,832]).

Furthermore, some studies have shown that there are individual differences in muscle reactivity

to pain [634,675]. One study that tested the excitability of central motor controls of masticatory

pathways in patients with TMD, and found that central motor pathways were not hyperexcitable

[195]. These variable findings between the interaction of TMD-related pain and orofacial motor

function may be related to a unique experience of pain, reflective of the heterogeneity of TMD.

In addition, some patients with TMD have limited unassisted jaw opening (with or without pain)

[539], whether they were symptomatic or asymptomatic [286]. Therefore, in the context of TMD,

it seems that the revised pain adaptation model (see section 2.1.6) best explains the observed

variability in motor abnormalities. This model accounts for individual differences in changes in

motor output to maintain homeostasis – the muscle, or region maintains contraction force, and

promotes nocifensive behaviour. Based on these findings, Svensson and Graven-Nielsen [839]

proposed that TMD pain interferes with sensorimotor integration and inhibits motor output to the

craniofacial region. They also suggested that differential patterns of excitation of masticatory

muscles occur in a phase-dependent manner to dampen jaw movements. Specifically, there can

be increased excitation of antagonistic muscles during contraction of the agonistic (painful)

muscle. Additionally, there is increased activity in the antagonistic phases of muscle movement

and decreased excitation during agonist muscle contraction, presumably to reduce agonistic

muscle activity, and thus pain [374]. In contrast to this proposed model, Cairns, Sessle and Hu

[133] reported that the injection of noxious chemicals into the TMJ increased both agonist and

antagonist muscle activity.

2.2.3.3. Cognitive abnormalities

There is extensive literature about the cognitive dimension of pain, and interactions between

cognition and pain in general (see section 2.1.4). A few studies have examined cognitive

performance in TMD and reported various cognitive deficits. For instance, Goldberg and

colleagues [370] showed that post-traumatic TMD patients performed poorly on simple and

35

complex reaction time tasks, as well as neuropsychological tests testing interference effects,

compared to patients with idiopathic TMD. In two follow-up studies, Grossi et al. [379,380]

determined that a patient’s performance on neuropsychological test, including he simple and

multiple-choice reaction-time tests, California Verbal Learning Tests, the Brown-Peterson

Consonant Trigram Auditory Memory Test, Sleep Assessment Questionnaire, and Beck

Depression Inventory, as well as fatigue and energy level assessments (100-mm visual analog

scale), predicted their treatment outcome. Weissman-Fogel and colleagues [928] used a cognitive

interference task, the counting Stroop task, and found that patients with idiopathic (non-

traumatic) TMD had sluggish reaction times compared to controls. Together, these studies

suggest that patients with TMD may have cognitive deficits. However, although there is evidence

that patients with TMD have a variety of cognitive deficits, it is not clear whether these deficits

existed prior to the onset of TMD pain or are caused by TMD pain.

2.3. Anatomy and physiology of orofacial pain and sensation

In order to understand concepts related to the possibly neurological and pathophysiological

aspects of orofacial pain and TMD in particular, it is important to describe the ascending sensory

and nociceptive pathways from sensory transduction in peripheral receptors, to transmission to

the brainstem, and supraspinal levels. The emphasis will be on the trigeminal system but some

differences with the spinal (i.e., in the trunk, neck and limbs) system are also noted.

Sensory information, such as touch, temperature, proprioception, kinaesthesia, itch and pain are

processed by the somatosensory system. In general, peripheral receptors transduce stimulus

energy and encode them as neural signals. These neural signals are propagated from the

periphery to the CNS via primary afferent neurones. The cell bodies of these afferents are

generally located in the periphery – with the notable exception of the mesencephalic nucleus of

the CNV (see below). These peripheral neurones project to central neurones. There are two main

routes for orofacial sensory information to travel from the brainstem to the brain: the trigeminal

lemniscus, and the trigeminothalamic tract (TTT) for craniofacial information.

2.3.1. Peripheral receptors and sensory afferents

The nervous system monitors the state of the internal and external milieus. There are many types

of specialized peripheral receptors, each with specific properties to encode different stimuli

36

[428]. Receptors in muscles, deep tissues, joints and tendons provide sensory information about

movement and limb position [573]. These include Golgi tendon organs and muscle spindles.

Other aspects of the somatosensation, such as touch, flutter, vibration and skin indentation are

encoded by encapsulated nerve endings that are either fast adapting (Meissner corpuscles for

flutter, Pacinian corpuscles for vibration) or slowly adapting (Merkel discs for touch, Ruffini

endings for stretch – both are related to skin indentation) [573]. Sensations of pain and

temperature in the skin and muscles are encoded by free-nerve endings [941]. There are also

receptors specific to hair follicles, which are mechanically-sensitive (mechanoreceptors) [346].

These nerve fibres are inside the hair follicle and wrap around the sheath. These rapidly adapting

fibres encode movements of the hair. Information regarding position of a limb is provided by

proprioceptors. In relation to this investigation, the TMJ is innervated by proprioceptors that

provide information about jaw position and the force being applied by masticatory muscles

[843].

Each receptor is the peripheral ending of a primary afferent fibre. The peripheral somatosensory

primary afferents can be classified based on their conduction velocity, a classification system

developed by Erlanger and Gasser [297]. The free nerve endings have unmyelinated small-

diameter C-fibres with slow conduction velocities (~< 2 m/s) [612], and Aδ fibres, which have a

conduction velocity of ~2-30 m/s [612]. The Merkel discs and encapsulated endings are

generally associated with large-diameter myelinated fibres (Aβ) with conduction velocities of

~30-70 m/s [449]. These axons transmit information pertaining to touch and proprioception from

the periphery to the CNS. Muscle spindles and Golgi spindle afferents that carry proprioceptive

and movement information are large myelinated fibres (Aα), which have a conduction velocity

of 70-120 m/s. Receptors that respond preferentially to noxious stimuli are called nociceptors

[796] and lack a receptive structure or corpuscle, but rather are free nerve endings [826]. Primary

afferents with free nerve endings respond to mechanical, thermal and/or chemical stimuli [941].

In cutaneous tissues, LTM and thermoreceptors (receptors that detect changes in temperature)

also have free nerve endings [941]. Similarly, LTM in skeletal muscles have free nerve endings

[598]. Free nerve endings have small-myelinated Aδ fibre or unmyelinated C-fibre axons fibres.

For instance, a subpopulation of unmyelinated fibres (C) detects stimuli that warm the skin [462]

and a subpopulation of small-myelinated fibres (Aδ) detects stimuli that cool the skin [202].

37

There are also some reports that a small number of Aß afferents may have nociceptive functions

[255].

Nociceptors can respond to a single modality (mechanical, thermal, chemical) or to multiple

modalities (e.g., mechano-heat, polymodal). Mechanical nociceptors respond to suprathreshold

mechanical stimuli [681], and thermal nociceptors respond to noxious temperatures [202].

Polymodal nociceptors respond to various modalities of noxious stimuli, including mechanical,

thermal and chemical stimuli [612]. Furthermore, nociceptors can be classified based on the type

of afferent fibre to which they are associated, Aδ or C fibres. There are also C fibre mechano-

heat sensitive afferents (CMH) and Aδ-mechano-heat sensitive afferents (AMH). It is believed

that these different types of nociceptors encode the different qualia of pain, such as burning,

aching, pricking, prickle [612]. The CMH and AMH are considered polymodal nociceptors, and

some respond to noxious chemical stimuli [224].

There are other classification schemes for primary afferents, including nociceptors, but they are

outside the scope of this thesis.

2.3.2. Orofacial nervous system: the trigeminal nerve

The CNV is a mixed nerve with both sensory and motor components. However, the sensory

segment is much larger than the motor portion [884]. The CNV has three main branches: the

ophthalmic branch (V1), the maxillary branch (V2) and the mandibular branch (V3).

V1 and V2 are entirely sensory nerves, while V3 is comprised of both sensory and motor fibres

[843]. The V1 innervates the scalp, forehead, upper eyelid, the cornea and the eye’s conjunctive

tissue, the nose, the nasal mucosa, the frontal sinuses and the dura and blood vessels of the

meninges. The V2 innervates the lower eyelid, the cheek, the nostrils and nasal mucosa, the upper

lip, the upper teeth and gingiva, the palate, the roof of the pharynx, the maxillary, ethmoid and

sphenoid sinuses and some of the meninges. The third branch of CNV, V3, innervates the lower

lip, the lower teeth and gingiva, the chin and parts of the jaw, parts of the external ear, and some

of the meninges. Somatosensory information from the floor of the oral cavity is also encoded by

V3. CNV is not the only nerve that encodes signals from the orofacial region. In relation to this,

somatosensory information from certain parts of the oral cavity, ear and the meninges is encoded

by the facial nerve (cranial nerve VII; CNVII), the glossopharyngeal nerve (cranial nerve IX;

38

CNIX) and the vagus nerve (cranial nerve X; CNX). Sensory fibres from CNVII, CNIX and

CNX enter the brainstem and synapse onto the same nuclei as CNV (see below).

The sensory aspect of the nerve receives input from receptors that encode touch, nociceptive,

proprioceptive and temperature stimuli from the face, facial and masticatory muscles and the oral

cavity [784]. Aδ and C fibres in the orofacial region, especially in the TMJ and muscles of

mastication are very similar to those present in the spinal nociceptive system [784]. There are

some notable tissues that are innervated by the trigeminal nerve, such as the tooth pulp, the

cornea and the dura regions that are only (or predominantly) innervated by nociceptors, but do

not have other somatosensory receptors. For instance, Davis and Dostrovsky [216] have

demonstrated that stimulation of the middle meningeal artery and the sagittal sinus activates

central nociceptive neurones in the trigeminal system.

The sensory fibres of the three branches of the CNV converge at the trigeminal ganglion, where

the cell bodies of afferent fibres are located. The trigeminal ganglion is similar to the dorsal root

ganglion in the spinal somatosensory system in terms of markers associated with nociception,

and nociception-related receptors [9,33,82,487,569]. From the ganglion, a single sensory root

emerges and enters the CNS at the level of the pons. Primary afferent fibres terminate in the

principal sensory nucleus (or main sensory nucleus; MSN) and the spinal trigeminal nucleus

(STN). There is another specialized trigeminal nucleus in the brainstem: the mesencephalic

nucleus of the CNV (MeT) (see below). Together, the MSN, the STN and the MeT form the

trigeminal brainstem sensory nuclear complex (VBSNC). Unlike the dorsal horn of the spinal

cord, the VBSNC is comprised of several nuclei with unique cytoarchitectonic and

organizational features [82].

The MSN is similar to the dorsal column (gracile and cuneate) nuclei where large-diameter

myelinated primary afferents that encode touch and position information from the body synapse

onto second-order fibres. Like these nuclei, the MSN maintains a somatotopic map of the regions

from which it receives touch and position information, namely the orofacial region [411]. The

second-order fibres from the MSN bifurcate, and a larger branch of these second-order fibres

forms the trigeminal lemniscus, which decussates and runs parallel to the medial lemniscus. The

medial lemniscus, which carries touch and position from the body, terminates at the

39

ventroposterior lateral (VPL) nucleus of the thalamus. The trigeminal lemniscus terminates in the

ventroposterior medial (VPM) nucleus of the thalamus.

The STN is a long nucleus that is located between the pons and upper cervical spinal cord

(around C2) and is comprised of three subnuclei, organized rostro-caudally: oralis, interpolaris

and caudalis (Vo, Vi and Vc, respectively) [656]. Vc is cytoarchitectonically similar to the spinal

dorsal horn [656], and has been termed the medullary dorsal horn [366]. A-β fibres bifurcate and

terminate in the MSN and along the STN, and Aδ and unmyelinated C fibres terminate

predominantly in the STN at the level of Vc and to the upper cervical spinal cord dorsal horn.

Interestingly, some Aδ primary afferents carrying pain and temperature information synapse onto

second-order fibres in Vo and Vi, but this does not occur with C-fibres [947]. There are several

somatotopic maps along the VBSNC. Specifically, somatotopy is maintained in the dorsoventral

plane of each nucleus: the mandibular region is encoded in the dorsal segment, then the

maxillary region, and, ventrally, the ophthalmic region. In Vc, the somatotopic organization has

been shifted, and has an “onion-skin” arrangement, where oral regions are represented rostrally,

and lateral regions of the face are represented more caudally. There is a second somatotopic

representation along the mediolateral axis of Vc where the head is inverted [453,801]. Second-

order fibres from the STN form the TTT and, as implied by the name, project to the thalamus,

while maintaining a somatotopic organization. These pathways are described in greater detail in

Section 2.3.3.1.

A unique feature of the trigeminal system is that the cell bodies of primary afferents that encode

proprioceptive information from the masticatory muscles and periodontal ligaments are located

within the central nervous system in the MeT [167,176,549]: this is the only site where soma of

afferent neurones occur within the CNS. The MeT fibres bifurcate in the brainstem. One branch

of these fibres forms a reflex loop: it synapses on to the trigeminal motor nucleus, which has

efferent fibres that projects back to the periphery. These loops form the basis of orofacial reflex

arcs (such as the jaw jerk reflex) [399-401,411,557]. The other branch of these fibres projects to

other brainstem targets via the tract of Probst [138]. For instance, some of these axons project to

the MSN where they synapse onto second-order neurones [918].

40

2.3.3. White matter pathways

This section will provide a brief overview of the major white matter tracts in the brain, especially

those related to brain regions implicated to nociception and pain processing and modulation will

be described, particularly in terms of the functional neuroanatomy of these tracts (although there

has actually been little work on the functional anatomy of white matter in humans).

White matter tracts are classically divided into three groupings: (1) projection fibres, that connect

the cortex with subcortical, brainstem and other regions; (2) commissural fibres, that connect

cortical regions from one hemisphere to the other hemisphere; and (3) association fibres that

connect cortical regions within a hemisphere [147] (See Figure 2-3).

Today, it is recognized that there are several types of white matter tracts in the brain, including

long association fibre pathways, striatal pathways, commissural fibres, and projection fibres

[237,760] (see Figure 2-3). There are also short “U” association fibres that connect neighbouring

gyri [237,311].

The corona radiata contain ascending and descending projection fibres that include motor tracts

that emerge from the primary motor cortex (M1), supplementary motor area (SMA) and

premotor cortex (PMC) and project to the basal ganglia, the brainstem motor nuclei

(corticobulbar tracts) and the spinal cord (corticospinal tract; CST) [147,644]. There are also

corticobulbar tracts that emerge from various PFC regions that are involved in various

homeostatic functions. Of particular interest are tracts from the PFC that project to the

periaqueductal gray (PAG), which is implicated in pain modulation. The corona radiata also

includes ascending tracts that project from the thalamus to the cortex, forming the

thalamocortical tracts (TCT) [644]. Furthermore, there are corticothalamic tracts that run parallel

to the TCT.

Somatosensory (including nociceptive) input from the orofacial region is mainly transported to

the CNS via the CNV. The primary afferents project to the VBSNC. Second-order neurones form

the TTT and the trigemino-bulbar tracts, which project to the thalamus and the various brainstem

nuclei (see above). Third-order neurones from the various thalamic nuclei project to cortical

regions, forming the TCT. For instance, tracts from VPM project to S1 and S2. These tracts

ascend to these cortical targets through the corona radiata, via the posterior limb of the internal

41

capsule (plIC) [237]. The MD thalamus projects to the cingulate cortex and the insular cortex via

the anterior corona radiata (ACR), which course through the anterior limb of the internal capsule

(ICAL) [237]. Similarly, there are corticothalamic tracts, which run parallel to the TCT.

Association tracts are intrahemispherical and project between distant cortical regions. For

example, the cingulum connects the entire cingulate cortex, and is connected to every lobe of the

brain: the medial frontal, insular, parietal, occipital, temporal lobes [147,644]. The superior

longitudinal fasciculus is comprised of three major tracts (I, II, III). It runs antero-posteriorly,

and connects frontal regions along the sylvian fissure, and the insula to the parietal and temporal

lobes [147], and form most of the external and extreme capsules (EC/ExC). Another association

fibre is the uncinate fasciculus, which connects the OFC to the temporal lobe, including the

hippocampus and amygdala [146].

Animal studies have demonstrated that the trigemino-bulbar (as well as the spinobulbar) tracts

project to supraspinal regions via the hypothalamus [258], thereby bypassing the canonical

thalamocortical projections. Therefore, it is likely that these fibres provide input to the circuit of

Papèz [664], the limbic circuit of the brain. This circuit includes the hippocampus, the

mammillary bodies, the anterior thalamic nucleus, and the cingulate gyrus [644]. These gray

matter regions are connected through several white matter tracts: the fornix, the

mammillothalamic tract, the thalamocingulate tract, and the cingulum. The hippocampus projects

to the hypothalamus via collaterals of the fornix [644]. The mammillary bodies receive input

from braintstem nuclei, including the parabrachial nuclei, and further project to the anterior

nucleus of the thalamus via the mammillothalamic tract. The thalamus projects to the cingulate

gyrus via the thalamocingulate tract. The cingulum bundle courses the entire length of the

cingulate gyrus, and projects posteriorly to the hippocampus [644,664]. Anteriorly, there are

reciprocal projections with the PFC [644]. Therefore, the circuit of Papèz may be implicated in

processing the affective-motivational dimension of pain. Furthermore, key parts of the circuit of

Papèz, especially the cingulate cortex, are connected to other brain regions implicated in pain

[21,23].

Commissural tracts comprise of interhemispherical tracts. The corpus callosum contains fibres

that project from one hemisphere to the other, including the frontal, parietal, occipital and

temporal lobes, and so is thought to transmit information between the hemispheres. Little is

42

known about the role of the corpus callosum in the context of pain, but one study has

demonstrated abnormalities in the anterior body of the corpus callosum of patients with CRPS

[356]. The bilateral amygdalae and the ventromedial temporo-occipital cortices are connected by

the anterior commissure.

Figure 2-3: Diffusion tensor imaging reconstruction of Meynert’s classification of white matter

tracts in the brain. The tracts are superimposed on medial and lateral images of the brain.

Reproduced with permission from: [147].

43

2.3.4. Nociceptive pathways and mechanisms

The ascending nociceptive pathways can broadly be divided into three components: primary

sensory afferents that transmit nociceptive information from the periphery to the CNS; second-

order neurones that transmit nociceptive information to the thalamus; and third-order neurones

that transmit nociceptive information from the thalamus to cortical brain regions. There are also

collateral projections from trigeminal brainstem nuclei to other brainstem nuclei, including the

reticular formation (RF) and other brainstem nuclei. Here, the peripheral nociceptive

mechanisms and both the spinal nociceptive pathways and the trigeminal nociceptive pathways

will be described.

2.3.4.1. Peripheral nociceptive mechanisms

Free nerve endings that respond preferentially to noxious stimuli are called nociceptors

[89,124,796]. The transductive properties of nociceptors are based on the receptor molecules

and/or ion channels on the free nerve endings.

Some nociceptors are specialized and have specific response properties (unimodal), whereas

other types of nociceptors (polymodal) respond to several noxious stimuli. Unimodal nociceptors

are the basis of different pain qualia, such as burning, stinging, and aching. Polymodal

nociceptors, for example, can either be C or Aδ (and some Aß) afferents that respond to intense

mechanical and heat stimuli (CMH, AMH), and can also respond to chemical stimuli [224,944].

Furthermore, AMHs can be classified as type I and type II [870]. Type I AMH cells (Aß or Aδ

fibres) have low mechanical thresholds and a high heat threshold. Type II AMH cells (Aδ fibres)

have high mechanical thresholds and moderate heat thresholds. Another class of nociceptive

afferents is C and Aδ fibres that do not respond to mechanical stimulation, or have very high

mechanical thresholds, and are thus called mechanically insensitive afferents or silent

nociceptors [611]. There are several types of C fibre nociceptors: CMH cells, mechanically-

insensitive C fibres nociceptors and polymodal nociceptors.

Another feature of the different fibres is the quality of pain they encode. Type I AMH fibres are

present in both hairy and glabrous skin, and respond to sustained high-intensity heat stimuli (i.e.,

they are slowly adapting to the noxious thermal stimulus). Type II AMH fibres are rapidly

adapting fibres and are also present in hairy and glabrous skin [444,863]. CMH cells respond to

44

noxious heat stimuli near the pain threshold, and the neural response increases with the stimulus

temperature. Because of the different conduction velocities of the fibres, the percepts from the

different fibres are felt at different times: myelinated Aδ neural responses reach the CNS before

neural responses from unmyelinated C fibres. For instance, when a noxious heat stimulus is

presented to glabrous skin, Aδ (type II AMH) fibres give rise to first-pain, which is a sensation

of pricking, sharp and aching pain. This sensation is followed by second pain, which is a

sensation of dull or burning pain elicited by C fibres [136,543,710,711,867,870]. It is also

noteworthy that stimulation of glabrous skin does not elicit a dual percept of pain; the absence of

type II AMH fibres eliminated the possibility of a first-pain sensation [612].

Muscles and joints also have nociceptors. It is believed that the primary cause of muscular pain is

ischaemia (reduced blood flow and oxygenation), which activates nociceptors [572].

Furthermore, the gradual accumulation of metabolites formed by muscular exercise can also

sensitize nociceptors and induce allodynia [375,376,432]. The percept of muscular pain is

usually mediated by Aδ and C fibres, which are HTM [81], which can be sensitized by algesic

chemicals [941]. There are also mechanically insensitive nociceptors that respond to algesic

chemicals (chemonociceptors) [480]. The TMJ is also innervated by nociceptors. Specifically,

the TMJ and its periarticular connective tissues are innervated by nociceptors. Furthermore,

evidence suggests that the articular disk contains low-threshold mechanoreceptors [34].

Another distinction in pain qualities is based on the spread of the pain. Specifically, cutaneous

pain is often localized, whereas muscle pain is can be diffuse [482,805,946]. This is caused by

the size of the receptive field of a single nociceptor and less dense innervation of muscles. The

receptive field of cutaneous nociceptors is in the mm2 range [89,483], whereas that of muscle

nociceptors is in the cm2 range [504,599]. Further, muscular and cutaneous pains have several

different characteristics. For instance, deep muscular pain is reported as more unpleasant [838]

and it has a different quality than cutaneous pain. These differences can be attributed to different

neurophysiological phenomena in the peripheral innervation of muscles and skin, as well as

differential central pathways encoding nociceptive information from these tissues.

2.3.4.2. Spinal nociceptive pathways and mechanisms

Primary afferents encoding innocuous and nociceptive information from the skin, the viscera,

muscles and joints enter the central nervous system at the dorsal horn of the spinal cord. These

45

tracts join Lissauer’s tract, posterolateral to the spinal cord, and send collaterals into the dorsal

horn at various levels of the spinal cord [941]. Thermoreceptive and nociceptive primary

afferents synapse onto second order neurones in the dorsal horn of the spinal cord. Rexed

[727,728] identified several layers in the spinal cord based on cytoarchitectonic features. He

identified six layers in the dorsal horn, three layers in the intermediate region and ventral horn

and one layer surrounding the central canal. The dorsal horn layers can further be classified as

the superficial layers (laminae I, or the marginal zone, and II, or the substantia gelatinosa) and

the deep layers (laminae III-V, or the nucleus proprius, and lamina VI) [410,720,941]. Cutaneous

Aδ fibres terminate in laminae I, II, and V, and cutaneous C-fibres synapse onto laminae I and II.

Deep tissue, including muscles and joints, and visceral afferents synapse onto various layers.

There are two paths for HTM afferents: one set terminates exclusively onto lamina I, and the

other set terminates onto laminae I, IV, and V [857]. Unmyelinated nociceptive afferents from

the deep tissues project to lamina I and deeper laminae, but not lamina II. There are two types of

nociresponsive second-order neurones: nociceptive-specific (NS) neurones and wide-dynamic

range (WDR) neurones, which predominate in laminae I/II and V/VI. A small proportion of

nociceptors also synapses onto second-order cells in the other laminae (III, IV and VI). Fibres

that encode tactile information other than pain and temperature bifurcate upon their entry to the

spinal dorsal horn, and form an ascending branch which forms the dorsal column and a

descending branch that synapses to fibres in laminae IV and V [857].

There are two types of second-order neurones that receive input from primary nociceptive

afferent fibres: NS and WDR cells. NS neurones respond exclusively to noxious stimuli [160].

WDR neurones respond to innocuous stimuli but have greater responses noxious stimuli. The NS

cells include the polymodal heat-pinch-cold (HPC) cells [397]. Second-order neurones form the

STT. The second-order fibres originating in the marginal zone are mostly modality-specific, and

include thermoreceptive (which respond to cooling), HPC, and NS cells [20,397,964,967]. These

fibres form 50% of the fibres in STT [185]. The substantia gelatinosa (lamina II) receives

nociceptive input and is mostly comprised of interneurones which project to laminae I and V and

to adjacent spinal segments, and thus these cells do not contribute to the STT. STT neurones

originating from deeper laminae (the border of IV/V, the lateral aspect of VI, and a smaller

proportion from VII and VIII) [24] of the spinal cord are mostly WDR cells [943]. There are two

STT pathways: the lateral and the medial STT [545]. The lateral STT is a monosynaptic tract that

46

encodes sensory-discriminative aspects of pain, which project directly to the ventrolateral

thalamus in a somatotopically organized fashion. The lateral STT receives input mostly from the

superficial layers of the dorsal horn [545]. The medial STT is comprised of monosynaptic and

polysynaptic tracts that encode affective-motivational aspects of pain, and receives its input from

deeper laminae of the dorsal horn. The medial STT projects to the medial thalamus as well as

several brainstem nuclei (see below). The receptive fields of the medial STT neurones are

typically large, unspecific and not related to stimulus intensity [363], and there is no somatotopic

organization [102]. Unlike the neurones in lamina I, lamina V cells do not maintain somatotopic

organization [913,941,942,944]. Furthermore, lamina I cells have small receptive fields

compared to lamina V cells, which have more extensive receptive fields. Therefore, it is likely

that lamina I cells are more involved in the stimulus localization, whereas lamina V cells, which

have a graded response to innocuous and noxious stimuli, are likely more involved in intensity

coding.

As mentioned above, lamina II neurones receive input from primary nociceptive afferents, but

rather than ascend along the STT, these cells play a largely modulatory role in the spinal cord.

Specifically, lamina II cells comprise mostly of interneurones that modulate second-order

neurones originating in lamina I and V. These cells may, in fact, represent the gate discussed in

the Gate Control theory of pain (see Section 2.1.1.4). In addition to the STT, there are other

spinal nociceptive pathways, but these are outside the scope of this thesis.

About 85% of the STT decussates and 15% of the fibres remain ipsilateral to the input [258,545].

Both pathways ascend the spinal cord to project to several thalamic nuclei. For instance, there are

STT projections to ventroposterior region of the thalamus, including VPL and the inferior

ventroposterior thalamus (VPI) [258]. STT fibres originating from lamina I also project to the

ventromedial aspect of the posterior (vmPO) nucleus and the ventrocaudal aspect of the medial

dorsal (MDvc) nucleus of the thalamus [184]. Lamina V and deep laminae neurones also project

to the centromedian (CM) and centrolateral (CL) nucleus of the thalamus [258]. Canonically,

these formed the lateral and medial pain systems. The lateral system included projections to VPI

and VPL and vmPO and their cortical projections. The medial pain system includes MDvc and

CL thalamus [869].

47

As mentioned above, the medial STT has collaterals that project to several brainstem nuclei,

which form the spinobulbar tract [184]. There are four main regions where this tract projects to:

(1) the catecholamine cell groups (A1-A7), the parabrachial nuclei, the PAG and the brainstem

RF. The role of these regions is pain is discussed in further detail in Section 2.3.6. There are

other spinal nociceptive pathways, including the spinohypothalamic pathway, the postsynaptic

dorsal column system, and the spinocervicothalamic systems [258], but specific details about

these pathways are outside the scope of this thesis.

2.3.4.3. Trigeminal nociceptive pathways and mechanisms

The trigeminal nociceptive system includes primary afferents from peripheral nociceptors.

Nociceptors are specific receptors in the skin and muscles and joints that are free nerve endings.

Different nociceptors encode different types of noxious stimuli, such as mechanical, thermal

(heat and cold) and chemical stimuli, and some encode more than one of these stimuli

(polymodal nociceptors) (see Section 2.3.3.1).

Small myelinated Aδ and unmyelinated C primary nociceptive afferent fibres have their cell

bodies in the trigeminal ganglion, and project to the VBSNC [784]. Nociceptive afferents

synapse all along the VBSNC, including the three subnuclei of the STN and MSN [214,216,784].

Specifically, these project mostly to Vc, although small-myelinated primary nociceptive afferents

(Aδ) fibres have been shown to synapse (in lesser density) along the entire VBSNC, including Vo

and Vi, and MSN [453]. The laminar organization of the Vc is similar to that of the spinal dorsal

horn. Cutaneous nociceptors from the orofacial region project to the outermost lamina (lamina I).

Dense connections within and between the STN nuclei, termed deep fibre bundles of the VBSNC

[367,452,639], link different nuclei within the complex that have a common somatotopic

representation [82]. Although the role of these connections has yet to be established, evidence

suggests that the connections from Vc have a facilitatory effect on the other nuclei. Evidence for

the facilitatory role of these interneurones comes from studies reporting that inhibiting or

lesioning Vc reduced or abolished the excitability of rostral portions of the STN to noxious

stimuli in the trigeminal region, and the rostral nuclei showed increased responses to innocuous

stimuli in the periphery [82,83,452]. There is also evidence of ascending tracts in the deep fibre

bundles [82,215,450,452,501] that have a modulatory effect of rostral regions of the STN onto

Vc [425].

48

Similar to the spinal dorsal horn, primary afferents have specific targets within the Vc (i.e., Vc)

[784]. For instance, LTM project to the deeper layers of Vc (laminae III – VI) [453]. Small-

diameter nociceptive afferents synapse onto laminae I, III, V and VI of Vc. Many of the

aforementioned local circuit neurones, which modulate rostral regions of the STN emerge from

lamina II [784]. This lamina, akin to the spinal substantia gelatinosa modulates other neurones in

the Vc. Second-order neurones in the STN can be categorized like those in the spinal dorsal horn:

NS neurones, which receive input from nociceptive Aδ- and C-fibres, and WDR neurones, which

receive convergent input from neurones which encode both innocuous and nociceptive

somatosensory information [784]. Both cell types show graded responses to noxious stimulus

intensity, or as more of the receptive field is stimulated [784]. NS and WDR neurones

predominate in laminae I/II and V/VI [783-785]. NS cells have ipsilateral, localized receptive

fields in the skin, whereas WDR neurones have larger, more-complex receptive fields, with

tactile and, often, pinch-sensitive areas [784]. There are also NS and WDR in more rostral

portions of VBSNC, including Vo, Vi and nuclei outside the VBSNC, such as interstitial and

paratrigeminal islands [784]. However, as these neurones show a graded response to stimuli in

the oral cavity and perioral regions, they are outside of the scope of this thesis.

Additionally, most NS and WDR neurones in Vc receive converging input from both cutaneous

afferents and from afferents supplying other structures in the craniofacial region, including dura,

cerebrovasculature, the cornea, the tooth pulp and, relevant to this thesis, the muscles of

mastication and TMJ [17,141,214,586,597,645,708,787,788]. Furthermore, some of these

neurones also receive afferent input from other cranial nerves and upper cervical cord neurones

[784]. Together, these converging afferent inputs provide a neural substrate for the poor

localization, spread and referral of pain in painful conditions of the TMJ and deep craniofacial

tissues [784].

As in the STT, the TTT projects to the thalamus, as implied by its name. Sensory fibres from the

VBSNC synapse onto third-order neurones in the VPM, the posterior nucleus (PO), the

parafascicular nuclei (Pf), and the submedial nucleus (SM) of the thalamus [257] (for further

discussion, see 2.3.5). Other tracts emerging from second-order neurones in the Vc project to

brainstem and brain regions involved in nociception and pain modulation (see Section 2.3.5 and

2.3.6). These include projections to the brainstem RF, the PAG, the hypothalamus, brainstem

49

motor nuclei and brainstem autonomic nuclei [784]. The brainstem targets likely form the neural

basis for sensorimotor and autonomic reflex loops.

2.3.4.4. Trigeminal motor pathways

Abnormalities in motor output in the trigeminal system have been associated with parafunctional

habits, including bruxism (grinding) and clenching of the jaw [628]. These habits have been

associated with pain in the orofacial region, and with TMD [840]. Furthermore, patients with

TMD show abnormal motor function in the orofacial region (see Section 2.1.6), and so an

understanding of the interaction of sensory and motor pathways is essential to this thesis.

The sensory and motor systems interact and show connections throughout the CNS [361]. These

interactions allow for rhythmic motor activity (such as walking and chewing), reflexes (such as

withdrawal from a noxious stimulus) and smooth execution of a voluntary task. These

interactions are the neural basis of feedback loops. Feed-forward loops allow for the use of

sensory information and cues in the external environment to prepare for motor output. These

loops are used for rapid actions, such as catching a ball [361]. The sensory visual information

about the trajectory of a ball provides necessary information for the motor system to properly

anticipate where the ball can be caught. Conversely, feedback loops allow for sensory

information to modulate or correct motor output to attain a desired output. The sensory signal is

compared to a reference, and if these do not match, error signals are produced to correct the

proper trajectory. These loops are used for slow movements, e.g., such as reaching for an object

[361].

The motor aspect of the CNV supplies the muscles of mastication, including the masseter

muscles, the temporalis muscles, the anterior digastrics, the medial mylohyoids, the medial

pterygoids, and the lateral pterygoids [912]. The motor fibres originate in the trigeminal motor

nucleus in the brainstem, and receive input from supraspinal motor centres and sensory

information from afferents in the brainstem [555]. The trigeminal motor nucleus is located in the

lateral pons, medial to the MSN. Motor efferents emerge from the trigeminal motor nucleus and

leave the pons at the ventral surface, forming the motor root of the CNV. The motor root is much

smaller than and medial to the sensory root [256]. Other cranial nerves also innervate the

masticatory muscles. For instance, CNVII innervates the posterior digastric muscle (as well as

50

other facial muscles), CNX innervates the palatoglossus part of the tongue, and the rest of the

tongue is innervated by CNXII [644].

There are three levels of parallel and hierarchical motor control: the cortex, the brainstem and the

spinal cord (or in the case of the trigeminal system, the sensory and motor nuclei of the CNV)

[361]. At the most basic level, the trigeminal sensory and motor nuclei mediate simple reflexes,

such as the jaw-jerk reflex. This monosynaptic reflex occurs when there is pressure on the

mandibular teeth. Proprioceptive information of the mandible is relayed to the trigeminal MeT

nucleus in the midbrain via primary afferents in the third branch of the sensory root of the CNV

[176,644]. These neurones’ cell bodies are located in the MeT nucleus, and some collaterals

synapse directly onto the trigeminal motor nucleus. The trigeminal motor nucleus projects

efferent fibres to the masseter muscle, the internal pterygoid, and the temporalis muscle, which

contract and depress the mandible.

The brainstem motor nuclei are also responsible for repetitive (or rhythmic) motor behaviour,

such as mastication (or chewing) [644]. In the case of mastication, it is believed that a central

pattern generator maintains the rhythmic behaviour. The controlled and patterned contraction of

muscles to produce rhythmic mandibular motion can be broken down into three different stages

[555]: the preparatory, reduction and pre-swallowing stages. The first step is breaking food into

chewable pieces. In some cases, such as a steak, this is done before the food reaches the mouth.

In other cases, such as biting into an apple or a sandwich, this step occurs at the mouth. In this

stage, the digastric muscles (jaw-opening muscles) are very active, and the masseter and the

temporalis muscle show little activity, but are involved as they are responsible for fast-closing.

The chewing (or reduction) cycle is characterized by three phases: opening and fast-closing and

slow-closing. The slow-closing phase occurs when teeth come in contact with food, and muscles

move the jaw laterally to grind food between the molars. The final stage of chewing – pre-

swallowing – consists of fast-closing, slow-closing, and three sequential opening phases. These

coordinated movements rely heavily on sensory feedback – location of food in the oral cavity,

state of the food, position of the jaw, lips, and tongue, etcetera. These sensory cues modulate the

rhythmicity of masticatory muscles while chewing. Taken together, there is evidence that a

central pattern generator at the level of the trigeminal motor and MeT nuclei exists, and

coordinates muscular activity during chewing [555,644]. The central pattern generator in the

brainstem can be modulated by corticobulbar tract (motor efferent tracts from motor cortical

51

regions) [555]. Further modulation can come from RF of the brainstem, which receives

convergent input from the cortex, and the VBSNC. It is important to distinguish that reflexes and

pattern generators do not require input from upper motor neurones (corticobulbar tracts), but

cannot function without this input [555].

The next level of motor control comes from the brainstem – namely, the RF [361]. As described

above, the RF modulates the central pattern generator based on sensory cues and efferents from

cortical regions and other brain regions such as the amygdala and the hypothalamus. Other

brainstem motor regions that project to the trigeminal motor nucleus include the red nucleus

[368], and the vestibular nucleus [256]. These findings suggest that motor output to the orofacial

muscles can be modulated by a number of different sources. The cortical regions represent the

highest level of motor control: voluntary motor movement [361,644]. These movements are

initiated in the brain. Specifically, M1 (Brodmann’s Area 4 (BA4)) is where voluntary motor

output is initiated. M1 receives dense input from the S1 (BA 1, 2, 3a, and 3b) [674,706,747,748],

and two regions involved in motor planning: the SMA, and the PMC (both regions are

designated as BA6) [644]. It has been suggested that somatosensory input into M1 may be

related to motor learning. In line with these findings, it has been shown that acute nociceptive

input inhibits motor learning, and that patients with TMD have abnormal motor output (see

Section 2.2.1.2).

M1, SMA and PMC have motor efferents that project to the trigeminal motor nucleus either

directly, via corticobulbar tracts, or indirectly, via the RF [644]. In primates, M1 is

somatotopically organized: the map is organized ventrolateral-dorsomedially, with the toes along

the medial wall, and the hand region in the middle of gyrus, and the orofacial region is

represented on the ventrolateral surface of the gyrus [644]. The corticobulbar tracts originate at

the lateral aspect of the M1 (or SMA, or PMC), descend along the genu of the internal capsule

(IC), and course through the cerebral peduncles in the brainstem [644]. There is both ipsilateral

and contralateral contribution (contralateral being dominant) to the corticobulbar tracts, which

decussate at the level of the cerebral peduncles in the midbrain, and synapse onto lower motor

neurones in the trigeminal motor nucleus. Therefore, the sensory and motor systems are highly

interconnected, and sensory information modulates motor output at several levels within the CNS

[556,644].

52

2.3.4.5. Supraspinal nociceptive and pain regions

Peripheral and brainstem nociceptive pathways project to the thalamus and other brain regions

implicated in pain. Herein, the subcortical and cortical brain targets of second-order neurones of

the TTT and its collaterals that bypass the thalamus will be described. Additionally, this thesis

will describe experimental findings from animal studies and electrophysiological studies

implicating these regions in nociception and the putative role of these regions in the experience

of pain. Neuroimaging studies describing neural correlates of acute and/or chronic pain will be

discussed in Section 2.5.

It is important to recognize that it has not always been clear that the cerebral cortex is involved in

pain perception. It has been proposed that the brain is where perceptions occur (c.f., sensations) –

even Aristotle recognized that perception was different from sensation. Sensation means to

encode stimuli, and perception is the interpretation of this sensation. Aristotle posited that

sensations were a property of the physical being, whereas he relegated percepts to the soul (or the

mind) [810]. Nonetheless, until the mid-twentieth century, there was little evidence for the cortex

to be involved in pain perception. For instance, Hitzig [426] stated that pain is a subcortical

phenomenon. Head and Holmes [417] found little evidence that the cortex is involved in pain – a

patient with extensive cortical damage still felt pain, and pain unpleasantness. In fact, they

suggested that, in the case of central pain – a condition where a stroke affects the thalamus – pain

arises because the thalamus is released from cortical control [514]. Cushing could not produce a

painful response in patients with electrical stimulation of the cortex [196], and so thought that

pain was a percept of the thalamus. Penfield and Boldrey echoed these findings in their paper in

1934 [677], stating that electrical stimulation of the brain did not elicit painful responses.

However, in a later study, they found that stimulating the postcentral gyrus (S1) did elicit painful

responses, but that these responses were rare – 11 of 800 stimulations produced a “painful”

experience, none of which were intense enough for the patient to object [679]. Additionally, they

suggested that the removal of cortex does not relieve pain [678]. From these findings, they

concluded that pain has little cortical representation. In support of this hypothesis, Adrian [6]

could not identify any impulses reaching the cerebral cortex from noxious or thermal stimulation

in the periphery in several animals.

53

Conversely, to treat a patient with thalamic syndrome (or central pain), procaine was injected

into the somatosensory cortex of a patient, who reported pain relief for a period of 2 months

[845]. Further, Russel and Horsley [746] excised a portion of S1 (in the area of the arm), and the

subject reported a loss of sensation and analgesia in the affected limb, however not all sensory

qualities were abolished. Similarly, a number of studies reported that patients who suffered

cortical lesions, especially in the region of S1, had altered pain sensations – in the absence of

thalamic damage (reviewed by [571]). Electrical stimulation of the cortex provided variable

results, as some groups found that stimulating S1 caused the patient to experience pain, whereas

several other studies S1 stimulation did not elicit pain. In one of these cases, stimulation of S1

reproduced the patients’ pain, and resection of S1 alleviated their pain [542,571]. Furthermore,

Dusser de Barenne and McCulloch showed that the function of VPL and S1 were interdependent,

and when deeper layers of S1 or VPL were ablated, the spontaneous electrical activity in the

other was diminished [281].

Although other theories have postulated that the percept of pain occurs in the brain, none had

incorporated a specific role for it in their model. However, no study has shown that pain is solely

due to a single area in the brain. As Marshall [571] suggested, S1 (and likely other cortical

regions) are involved in pain, but may not be the “endpoint” of pain – that S1 receives and

processes nociceptive input (likely sensory-discriminative information); however, this is not the

region where the nociceptive input is perceived as pain. Rather, this information is processed and

“sent” to other brain regions where the percept of pain occurs. To address this point, Melzack

suggested that a network of brain regions working together is required for the percept of pain

[592-594]. This pain “neuromatrix”, as he called it, is supported by the multidimensional

experience of pain. This theory required the concerted activation of several brain areas to

produce the experience of pain, each set underlying a dimension of pain. Therefore, it was

proposed that to have a conscious experience of pain, many brain regions, each related to a

dimension of pain (cognitive, modulatory, sensory-discriminative, affective, motor) are required.

This theory provides an explanation for the finding that stimulating a single brain region does not

consistently produce pain. However, as will be described, to some extent in this section, and in

further detail in Section 2.5, few (if any) studies have been able to show that these regions of the

brain are, effectively, functionally connected – a fundamental tenet of a “pain matrix”. In fact,

54

the concept of the so-called “pain matrix” idea has now fallen out of vogue, and other theories

about cortical brain regions have been postulated, and will be discussed herein.

Thalamus

The STT (and the TTT) project to the thalamus. Interestingly, the TTT ascends bilaterally, albeit

the contralateral thalamus receives a greater density of projections [784]. These tracts project to

the ventroposterior nuclear group (ventrobasal complex in subprimates) in a somatotopic fashion.

In fact, the lateral STT projects to the lateral portion (called the VPL) [351] and the TTT projects

to the medial portion (called the VPM) [340]. These projections maintain the somatotopic

organization of the VBSNC. The ventral STT projects to other thalamic nuclei, including MD, Pf

and CL [869]. Gaze and Gordon [351] identified NS neurones in the thalamus of cats and

monkeys for the first time. Similarly, Perl and Whitlock [684] identified NS neurones in the

ventrobasal thalamus of cats and monkeys. These neurones showed responses only to noxious

stimuli, and had a specific (c.f. large or diffuse) receptive field in the contralateral body.

Similarly, Applebaum and colleagues [29] confirmed these findings with antidromic mapping of

VPL neurones in the upper lumbar region of the spinal cord. There are also WDR neurones in the

thalamus of humans and animals [142,527]. Electrophysiological recordings and stimulation

studies in humans undergoing neurosurgery have confirmed that the human thalamus receives

nociceptive input, and these inputs are topographically organized in the ventroposterior nucleus

[223,396,534]. The somatotopic organization of STT and the ventroposterior thalamus suggests

that these nuclei may be related to the sensory-discriminative aspect of nociception, i.e., the

location of the stimulus, and the graded response to nociceptive stimulus suggests that this region

is implicated in magnitude encoding, i.e., the intensity of a stimulus. These concepts were

supported by findings that stimulating VPL results in localized pain in humans [415].

Additionally, some STT neurones projecting to VPL have collateral projections that synapse onto

CL [30,362,589].

Furthermore, Perl and Whitlock [684] demonstrated that the STT also projects to CM and

intralaminar nuclei, which only receive nociceptive input and have bilateral receptive fields, and

the posterior nuclear group of the thalamus, which has diffuse receptive fields. These findings

were confirmed by several tracing studies [25,102,129,485,488,603,701]. Additionally,

connections from the STT to the VPI were also found [25,85,701].

55

The STT and TTT also have topographical input to MDvc: the STT projects to the anterior

portion, and the TTT projects more caudally [186,341]. There is also evidence of a pain- and

temperature-specific nucleus – the VMpo [188]. Neurones that responded to cold perception, in

the vicinity and had some of the same properties attributed to VMpo were located in the human

thalamus [223]. Blomqvist et al. [101] found a site in the human thalamus with STT projections

in the vicinity of VMpo.

Projections of the TTT are similar to those of the STT. As mentioned above, these ascending

tracts project to different nuclei within the ventroposterior thalamus: the VPM and VPL,

respectively. The VPM receives input from all levels of the VBSNC, however, only dorsal MSN

neurones project to the ipsilateral VPM, whereas ventral MSN and STN project to the

contralateral VPM [126,127,140,574,618,719,811,812,831,861,948,957]. Furthermore, there are

also projections to the PO and MDvc [127,341], and projections to CL and CM have also been

identified [948].

Finally, the ventrolateral nucleus of the thalamus (VL) receives input from deep dorsal horn, via

the STT [85]. These projections overlap with cerebellar projections to VL. Third-order neurones

project from VL to the motor cortex, and so it has been suggested that these inputs are related to

sensorimotor integration, or the motor dimension of pain [794].

In summary, the thalamus receives thermal and nociceptive information from the STT and TTT.

The tracts project to several nuclei within the thalamus, which have various cortical targets.

These various projections suggest that the different targets contribute to various dimensions of

pain. These concepts are discussed in further detail herein.

Primary somatosensory cortex

In primates, S1 is located along the postcentral gyrus, and is comprised of four cytoarchitectonic

rostrocaudally organized regions: BA 3a, 3b, 1, and 2. These regions have functional differences:

areas 3b and 1 receive cutaneous tactile input, and areas 3a and 2 respond to proprioceptive

input. Furthermore, each of the four regions in the somatosensory cortex has a complete body

(homuncular) map: the foot is located medially, and the face is located on the ventrolateral bank

[196,652,679].

56

S1 receives input from the somatosensory thalamus. Specifically, in primates, there is evidence

of projections from the VP thalamus to each of the cytoarchitectonic regions of S1 [644]. As

information is processed from BA 3b to BA 1, the size of the neuronal receptive field increases,

and the neurones have more complex response properties, which is an indication of integration

and modulation of sensory input within S1 [644]. Similarly, BA 3a projects to M1, BA 2 and S2.

The receptive fields in BA 2 and S2 are larger and the response properties are more complex than

in BA 3a. Furthermore, there is evidence of projections from S1 to other cortical regions,

including reciprocal connections between 3b, 1, and 2 and the posterior parietal cortex (PPC, BA

5 and 7), and contralateral somatosensory cortex. Area 3a projects to M1. There are also

projection to distant cortical regions, including the insula, the cingulate cortex and the parietal

lobule and operculum [644]. There is also evidence for descending projections from S1 to

subcortical regions [644].

As discussed above, the role of the S1 in pain processing is contentious. Some of the earliest

documented evidence comes from a study by Dusser de Barenne in 1916 where he reported that

cats showed an allodynic response (which he called hyperalgesic) to noxious stimuli in the

periphery by applying strychnine, an alkaloid that blocks inhibitory signaling [825], to S1 cortex

[280]. In humans, electrical stimulation of S1 only sometimes elicited a feeling of pain in the

patient (e.g., [679]). Similarly, in some but not all cases, phantom limb pain was resolved after

S1 ablation [293,383,542,833] (reviewed in: [296,571,845]). Furthermore, resection of S1 was

ineffective in relieving pain in other pain disorders. Conversely, patients with Rolandic seizures

(epilepsy in the region of S1 and M1) reported pain [961], and lesions to parts of S1 caused a

disorder similar to thalamic syndrome [941].

Studies in non-human primates support the aforementioned findings in humans undergoing

neurosurgery – that S1 has a role in nociception. For instance, Peele showed that ablating

specific BA in S1 or PPC in monkeys led to a loss of nociception for nine months, and a

permanent loss of the ability to locate a noxious stimulus [676]. Furthermore,

electrophysiological studies in S1 in non-human primates have demonstrated that nociceptive

regions of the thalamus project to BA 1 and 2 [485,631]. There is also evidence of somatotopic

organization [270,484]. Kenshalo and Isensee [486] identified neurones in areas 3b and 1 that

responded maximally to pinching in the receptive field, and a proportion of these showed a

57

graded response to noxious heat stimuli in the receptive field. Furthermore, there is evidence that

area 3a neurones contribute to allodynic and hyperalgesic phenomena [929].

Specific to the trigeminal system – and of particular interest to this thesis – Biedenbach and

colleagues electrically stimulated the tooth pulp in non-human primates, and reported graded

responses in the face area of S1, mainly in area 3a [91]. Similarly, Chudler and colleagues found

cells that in the somatotopic region of the face in BA 1 and 2, which also showed a graded

response to nociceptive input [161]. Interestingly, Biedenbach and colleagues suggested that S1

cortex is likely essential for localizing noxious stimuli on the body, but also observed responses

in neurones outside S1, and so suggested that S1 contributes to pain perception, but this role is

not exclusive to S1.

The parasylvian cortex: S2 and insula

The parasylvian cortex includes several regions: S2, insular cortex, and the ventrolateral

prefrontal cortex (vlPFC, which will be discussed later). S2 is comprised of two adjacent

cytoarchitectonic regions: BA 40 and 43, which are in the ventrolateral bank of the postcentral

gyrus, and within the parietal operculum [644]. The insula is a very large, heterogeneous region.

It is comprised of two long sulci in the posterior portion, and three short sulci in the anterior

portion [39,40]. The insula can be divided into three subregions: the anterior, middle and

posterior insula, based on functional and anatomical features.

S2 receives dense connections from S1, and some input from the VP thalamus. S2 also has a

somatotopic organization, but much less detailed than that of S1 [735,737,930,941,952]. S2

maintains this somatotopic organization in its projections to M1 [859]. Evidence from non-

human primate studies has shown that, in addition to performing integrative aspects of

somatosensory processing [294,644], neurones within S2 respond to noxious stimulation

[65,736,930], and that these neurones show somatotopic organization and graded response to

nociceptive laser-evoked potentials [65]. Electrophysiological studies in humans have shown that

S2 responds to noxious stimulation [336,338,583,734,871,890,953]. Frot and colleagues have

demonstrated that S2 shows graded responses to increasing intensity of non-noxious and noxious

somatosensory stimuli [335]. Taken together, it would seem that S2 is implicated in processing

the sensory-discriminative dimension of pain. However, the receptive fields in S2 are larger and

more complex than those of S1, and so it is possible that S2 plays a more integrative role and

58

outputs this information to motivational and motor regions to respond to a noxious stimulus

accordingly.

The insula is implicated in a variety of functions. Evidence for this comes from anatomical

studies that have revealed that the insula receives input from primary sensory, limbic, motor, and

other brain areas, and projects to a number of brain regions [40]. The insula is considered to be a

visceral sensory and visceromotor area [40]. Further evidence for the role of the insula in several

systems comes from neuroimaging studies – the insula is one of the most commonly activated

brain regions in various experimental paradigms [956]. Craig has suggested that the insula is the

homeostatic/interoceptive centre of the brain [187 10572,190]. It has also been suggested that the

insula is a node in a general salience network [263,265,616]. In terms of pain perception, the

insula is considered as a part of the medial pain system [869]– albeit it has also been suggested

that it may be part of the lateral pain system (i.e., its role in sensory-discriminative dimensions of

pain are under debate). The dorsal posterior subregion of the insula receives direct input from

VMpo, a nucleus in the thalamus that exclusively receives thermal and nociceptive input [188].

Evidence for nociceptive input to the insula comes from studies of electrical stimulation and

recordings of the insular cortex [7,337,583,658,688]. Further evidence comes from lesion

studies, where patients suffered sensory loss after a lesion of the insula (Obrador, 1957, cited in:

[343,377,532]) and surrounding areas [92]. Another lesion study found that the patient showed

hypalgesic responses to noxious stimuli after an ischaemic stroke affected the parasylvian cortex

(insula, parietal operculum and S1) [233]. Greenspan and Winfield [378] reported that a patient

with a tumour that was compressing the posterior insula and the parietal operculum had

increased pain thresholds to heat, cold, and mechanical noxious stimuli. Conversely, Starr and

colleagues [829] found that two patients with insular lesions showed increased heat pain intensity

ratings. Greenspan and colleagues [377] showed that patients with lesions to the parietal

operculum, where the insula was spared, had decreased discrimination of noxious stimuli,

whereas in patients where the insula was damaged, there was decreased motivation to escape the

noxious stimuli (it was less aversive).

Because of the diverse input into the insula, a number theories about its role in pain perception

have been postulated. For instance, it has been postulated that the insula is a multidimensional

integration site for pain [116]. This suggests that the many dimensions of pain are integrated

and/or interoceptive region [183]. Alternatively, Coghill [169] suggested that because of the

59

behavioural relevance of pain, and its essential role is survival, there are several “backup” pain

intensity coders, including S1 and the insula. In line with these findings, it has recently been

suggested that the insula is a central magnitude estimator for the brain [48], and that there is a

also pain-specific magnitude estimator in the insula. These estimators have differential

connections and are hubs of cortical and subcortical networks, respectively. An alternative

perspective is that the insula may be a salience detector in the brain (for a more comprehensive

discussion of these ideas, see [616], and Section 2.5.1.1)

The cingulate cortex

The cingulate cortex surrounds the corpus callosum, anterior to the genu, dorsal to the body and

posterior to the splenium. In 1937, Papèz [664] included the cingulate cortex as part of the limbic

circuit he identified, the brain’s emotional circuit. Although this remains true today, many other

functions have been ascribed to the cingulate cortex, including cognitive, motor, limbic

nociceptive, pain-modulatory, visuospatial and memory [899]. However, there are several

subregions within the cingulate that are specialized for these functions, although, like many other

associative brain regions, there is much integration and overlap of function between adjacent

subregions [899].

There are many classification systems used to differentiate subregions within the cingulate

cortex, which add to the complexity of comparing studies that use different species or different

nomenclatures. In this thesis the nomenclature developed by Vogt and colleagues [905] will be

used, whenever possible. It is noteworthy that the region previously referred to as the anterior

cingulate cortex (ACC) has now been divided into two regions: the mid-cingulate cortex (MCC)

and the ACC. This nomenclature was developed based on a multimodal approach, including

eletrophysiological properties of cells, cytoarchitectonic studies, lesion studies and neuroimaging

studies [900,905]. The subdivisions of the cingulate include the ACC, MCC and the posterior

cingulate cortex, and the following description of these regions comes from: [899]. The ACC can

be further subdivided into the subgenual ACC (sgACC; BA 25) and the perigenual ACC

(pgACC; BA 32, 24 a, b, c). The MCC can be further subdivided into the anterior MCC (aMCC;

BA 33’, 32' and 24 a', b', c'), which includes the rostral cingulate motor area (rCMA) and the

posterior MCC (pMCC; BA 32' and 24 a', b', c'), which includes the caudal cingulate motor area

(cCMA). The PCC is also divided into dorsal (dPCC; BA 31, 23) and ventral (vPCC) areas

60

[906]. There is additionally the rostral caudal zone (RCZ; BA 29, 30), which is ventral and

anterior to the dPCC [900]. This thesis will focus on the ACC and MCC, whereas the PCC and

RCZ are outside the scope of this thesis.

It is worth prefacing the specific descriptions of the roles of the ACC and the MCC with the

observation that these regions are among the most often activated in brain imaging studies,

across various paradigms [956]. Therefore, descriptions of specific findings may be related to

broader functions, such as salience, rather than specific functions.

Tracing studies in non-human primates have elucidated thalamocingulate connections. The ACC

receives the ventral anterior, CL, Pf, and other midline and intralaminar nuclei of the thalamus

[798]. There are some, albeit a small proportion, projections from the MD nucleus. Specifically,

the anterior thalamus projects to the BA 32 region of pgACC and sgACC. MD thalamus projects

to both subregions of the ACC (all three BAs). The ACC has been implicated in modulating

emotional, mood state and visceromotor functions in humans. In monkeys, the ACC projects to

subcortical and brainstem structures involved in these functions, including the hippocampus and

bed nucleus of the stria terminalis, hypothalamus and the PAG and parabrachial nucleus. There

are also extensive projections from area 24 to the MD thalamus [803]. The sgACC is

interconnected to several frontal brain regions, including the pgACC, the OFC, vlPFC, and

mPFC. The pgACC receives input from the sgACC, MCC, PCC, the insula, the frontal pole, the

OFC, mPFC, the frontal operculum, and the dorsolateral prefrontal cortex (dlPFC).

The MCC includes the CMA, which has direct projections to motor neurones in the ventral horn

of the spinal cord [623,691]. This subregion is involved in the planning and execution of

volitional movement. The MCC receives input from VL, VPL, CL, CM-Pf, and MD [798].

Corticothalamic projections from the MCC synapse terminate onto ventroanterior nucleus (VA)

and MD. Tracing studies in non-human primates have also elucidated anatomical connections

between the MCC and other cortical regions, including the insula, the temporal, parietal, frontal

and limbic association areas. The MCC connects to the sgACC, pgACC, the OFC, pre-

supplementary motor areas, mPFC, PCC, PMC, frontal operculum, the vlPFC and the dlPFC

[686]. Additionally, the CMA receives input from M1 and more ventral regions of the MCC

[416].

61

Of relevance to this thesis are the intrinsic connections (via the cingulum bundle) within the

cingulate cortex, which modulate and coordinate attention, cognitive processes and response

behaviour [622,901]. Additionally, these intrinsic connections are modulated by input from and

output to the dlPFC – a region involved in selecting information about the sensory environment

(attention, self-monitoring, personality, working memory) and decision-making. Therefore, the

connection between the dlPFC and the cingulate cortex is of particular interest in the context of

pain processing (heightened awareness of pain) and pain modulation. Another important

connection is between the OFC and the cingulate [150,686]. The OFC is involved in autonomic

and emotional processes, and goal-directed behaviour [742]. It is believed that the OFC and

dlPFC modulate sgACC output (potentially via the MCC’s intrinsic connections) to subcortical

and brainstem nuclei that are involved in pain modulation (see Section 2.3.6).

Lesion and electrophysiological studies in humans have implicated the MCC in pain processing.

Specifically, patients who had large regions of their cingulate cortex lesioned to relieve

intractable pain (a procedure called “cingulotomy” [50]) showed attenuated emotional responses

to pain – they still perceived pain, but it was no longer distressing [327,326]. These findings

suggest that the cingulate cortex is implicated in the affective dimension of pain processing.

Later studies reported equivocal post-cingulotomy outcomes for pain relief

[50,51,414,696,938,939]: from 23% [439] to 75% [327]. This variability is likely due to the

location and size of the lesion within the MCC [733]. In addition to changes in the affective

dimension of pain post-cingulotomy, Davis and colleagues [217] report that a patient with a

relatively small MCC lesion showed hyperpathic intensity and unpleasantness responses to

noxious thermal stimuli, suggesting that the cingulate may also have a role in the sensory-

discriminative (or intensity coding) dimension of pain. Furthermore, Dierssen and colleagues

elicited a painful response in the contralateral body and face of a patient with phantom limb pain

when they stimulated the MCC (case V.G.D., point no. 8 in [254]).

In line with these findings, Hutchison and colleagues [440] identified neurones in the human

MCC that respond to painful cutaneous stimuli, and Lenz and colleagues [533] reported laser-

evoked potentials in the MCC as measured by microelectrode recordings to noxious laser stimuli

in the periphery. Iwata and colleagues [451] identified neurones in the CMA that are

nociresponsive, which are involved in attention to pain, and nocifensive behaviour. Similarly, in

a reward and pain-avoidance cued task, Koyama and colleagues [499] found that cells in the

62

MCC (likely the CMA) were implicated in stimulus prediction and anticipation, and response

selection. Taken together, these findings suggest that the cingulate cortex is implicated in the

affective-motivational dimension of pain: as pain becomes more bothersome or aversive, there is

greater motivation to escape it, and these processes are encoded within the MCC.

However, it is essential to note that the MCC is involved in a number of complex tasks, and

therefore, may serve a more general role in the brain: salience detection and response selection.

Neurones in the MCC are activated during tasks that have a cognitive and/or emotional load and

that are attention demanding, such as mental arithmetic, word generation, and Stroop tasks

[218,229]. Furthermore, cingulotomy increased errors in motor output in a monetary reward task

[940]. Therefore, the MCC may, in fact, be a generalized system for orienting the self to a more

salient stimulus, and choosing an appropriate response, whether it is to avoid a noxious stimulus,

or to increase rewarding behaviour.

Prefrontal cortex

The PFC is a large, heterogeneous region of the brain comprised of a number of subregions

involved in various higher cognitive and limbic functions. Broadly defined, it is believed that the

PFC is central to the cognitive-evaluative dimension of pain. Specifically, various regions of the

PFC have been implicated in cognitive modulation of pain; however, most of these data come

from neuroimaging studies, and so will be discussed in Section 2.5.1 and 2.5.2. Non-imaging

evidence to support a role for the PFC in pain perception and/or modulation is paltry. One line of

evidence comes from an extreme surgical procedure: frontal lobotomies, which were used to

relieve intractable pain, amongst other conditions [154,288,330,880,923]. However, the extent

and specificity of these surgeries is unclear, and regions such as the cingulate cortex – which is

known to be involved in pain perception and modulation – could have been destroyed.

Furthermore, a tracing study that injected murine VBSNC neurones with a neuroanatomical

tracer, horseradish peroxidase (a retrograde tracer), did not identify any cortical targets in the

PFC [570]. Additionally, a later study injected a viral tracer (an anterograde, trans-synaptic

neural tracer) into the murine tooth pulp, and failed to identify any infected cells in the PFC.

Additionally, Dum and colleagues [277] injected the STT with a viral tracer, and also failed to

identify any infected cells in the PFC. However, tracing studies identified corticofugal tracts

from the PFC to brainstem regions involved in pain modulation in the rats [70,406,536], cats

63

[823] and monkeys [405,520], and stimulation of the PFC in rats modulates nociceptive input

[403,404]. Further, anaesthetization of the PFC of rats reduced pain modulation, as measured by

the jump reflex threshold [175]. In a recent histological study of a rat model of spared-nerve

injury, Metz and colleagues [608] reported that there are morphological and functional changes

in pyramidal neurones in the mPFC. Specifically, they found increased dendritic branching,

increased spine density in the nerve injury group, compared to the sham group. Additionally,

these structural changes were associated to functional changes in synaptic currents, which

correlated with tactile thresholds in the injured limb. More recently, Ji and Neugebauer [460]

have attempted to elucidate the neural underpinnings of observed abnormal mPFC activity in

neuroimaging studies of pain [47,356] (see Section 2.5.1). They found that cells in the mPFC

were effectively inhibited in a murine chronic pain model, and may be related to cognitive

deficits in chronic pain [461]. Additionally, there is evidence that transcranial stimulation of the

dlPFC modulates pain in experimental pain paradigms [310] and in chronic pain

[114,649,754,802]. Taken together, it seems that the PFC does not play a role in pain perception,

but rather may be involved in cognitive appraisal and evaluation of pain, and its subsequent

modulation.

Motor regions

Pain and motor interactions have been described in Section 2.1.6, and the anatomical pathways

pertinent to this thesis are described in 2.3.2. This section will describe tracing and

electrophysiological studies in animals and humans that have implicated suprabulbar motor

regions in pain. In general, nociceptive input into motor regions is believed to serve two

purposes: (1) orient the body toward the threat (noxious stimulus) and (2) initiate nocifensive

behaviour (e.g., avoid the stimulus). Several lines of evidence have implicated suprabulbar motor

regions in pain. The first line of evidence comes from a study that investigated dorsal column

stimulation in monkeys [53]. This study found that the stimulation of the dorsal column

modulated nociceptive input into the Pf thalamic nucleus, which indicated that modulation of

nociceptive input occured at the level of the thalamus, rather than in (or in addition to) the spinal

cord.

Further evidence for supraspinal motor-pain interaction comes from unexpected findings that M1

stimulation (motor cortex stimulation; MCS) reduced spontaneous hyperactive firing in a model

64

of central pain in the cat [874]. Interestingly, stimulation of S1 did not have an effect on

hyperactivity in the thalamus. Furthermore, studies have demonstrated that MCS in cats

suppresses nociceptive activity in the spinal and medullary dorsal horns [182,789]. MCS has

become a prevalent therapy for intractable neuropathic pain; however, the mechanism of action

remains unclear. It is believed that MCS recruits distant antinociceptive brain regions through

corticocortical connections [643]. Similarly, there is ample evidence that transcranial magnetic

stimulation (TMS) – a less invasive method to stimulate the cortex – of the M1 is effective tool

to modulate neuropathic pain (e.g., [8,331,463,643,821]). Additionally, there is likely activation

of CST that can modulate nociceptive input (as postulated in the Gate Control theory of pain, see

Section 2.1.1.4).

Several studies have investigated the interaction of pain and motor learning in the brain. For

instance, one study found that a novel tongue-protrusion task is associated with acute and

reversible neuroplastic changes in the face region of M1, as measured by TMS [41,842].

However, noxious stimulation of the oral cavity disrupts learning the motor task, and related

neuroplasticity in M1 [109].

Another line of evidence for motor-pain interactions in the brain comes from viral tracing studies

of the STT and that have identified infected neurones in subcortical and cortical motor regions.

For instance, Dum and colleagues [277] found that, in addition to S1, S2, and the posterior

insula, M1 (BA 4), the ventral premotor cortex (PMv; BA 6) and the CMA (BA 32) receive input

from thalamic targets of the STT.

The basal ganglia are a group of subcortical nuclei that have classically been associated to motor

behaviour [13,239,640,945]. However, anatomical, functional and behavioural studies have

elucidated several roles of the basal ganglia. The basal ganglia are organized in a number of

parallel and segregated anatomical circuits forming functional loops [665] and the role of these

loops is integrated to mediating goal-directed behaviours [388]. In the context of nociception and

pain, there are a number of findings suggesting that the basal ganglia have a complex, and yet

undetermined role. For instance, Dum and colleagues [277] reproduced (but did not present)

findings that subcortical structures, including the basal ganglia receive direct input from the

anterolateral tract of the spinal cord. Previous studies demonstrated that nocireponsive neurones

project directly to the globus pallidus and the putamen [642], and that neurones in the globus

65

pallidus respond to nociceptive input [88,164]. Furthermore, the caudate nucleus of the rat and

the cat has neurones that respond to noxious mechanical stimuli in the periphery and that are

somatotopically organized [164,544,732,769]. Also, there is evidence that striatum is densely

populated with opiate receptors in rats [38] and humans [97]. In line with these findings,

stimulation of the caudate nucleus has been shown to produce analgesic effects in the monkey

[550].

Further support for the role of the basal ganglia in pain comes from evidence that patients with

movement disorders related to the dysfunction of the basal ganglia, such as Parkinson’s disorder,

often report pain [162,385,398,624,635,707,759,965]. Furthermore, depletion of dopamine in the

substantia nigra (the region affected in Parkinson’s disorder) in mice leads to hyperalgesic

responses to noxious stimuli in the periphery [163]. Stimulation of the striatum also modulates

orofacial pain: stimulating dopaminergic neurones has been shown to modulate the nociceptive

jaw-opening reflex in rat [55,78].

In a recent study of patients with lesions in the putamen, Starr and colleagues [828] found that

patients showed lower cold pain detection thresholds (i.e., less sensitive), lower pain intensity

ratings to suprathreshold heat pain and decreased pain intensity ratings for prolonged noxious

heat stimuli on the affected side (contralateral to the lesion), compared to the affected side and to

controls. However, the patients’ tactile thresholds remained unchanged. Therefore, it seems that

damage to the putamen causes decreased pain sensitivity, but not tactile thresholds, suggesting a

role in the putamen in the sensory-discriminative dimension of pain.

The nucleus accumbens, integral to the reward circuit in the mammalian brain [388], is another

region within the basal ganglia that has recently been implicated in pain processing. The concept

of the reward pathway encoding aversive stimuli is reminiscent of the age-old dichotomy of

pleasure and pain, and can be dated back to Platonic and Aristotelian ideas [199]. It has been

reported that, in rats, the nucleus accumbens is pronociceptive, and that inhibiting the nucleus’

function with opioids induces antinociception [352].

It is tempting to conclude that the basal ganglia are involved in determining a behavioural

response to pain - the motivational and nocifensive dimensions of pain. However, the substantial

input of nociceptive information from different regions of the brain and the extensive

connectivity of the basal ganglia suggest that these subcortical nuclei may have a more

66

prominent role in pain perception. For instance, it has been suggested that the basal ganglia may

be the site of convergence of the different dimensions of pain (i.e., pain integration) [828];

however, more research is required to substantiate this theory. It is noteworthy that sensory

convergence has been demonstrated in several nuclei within the basal ganglia [162,165].

Finally, there is evidence that changes to sensory input to the brain, including prolonged

nociceptive input, can modulate M1 excitability and output. Specifically, when a peripheral

nerve is transected in mice, neuroplasticity occurs in M1 [2,4,328,329]. However, it is possible

that these changes in M1 reflect changes in S1 cortex, which has dense projections to M1 [41].

Furthermore, evidence in humans suggests that acute pain in the periphery can decrease M1

excitability [3,157,301,302,744] and acute intraoral pain can prevent learning a novel tongue-

protrusion task [109].

In sum, motor regions in the brain receive nociceptive input, have nociresponsive neurones and,

in some cases, are modulated by nociceptive input. The most obvious role for motor regions in

the context of pain is to initiate a motor response to pain (nocifensive behaviour), however,

evidence suggests that they may have a more prominent role in pain integration, and pain

modulation.

2.3.5. Descending modulation

The perception of pain is modulated by several psychological factors, including attention, arousal

and expectation, as well as the context in which the stimulus is presented [309]. This can be

illustrated by two examples: athletes injured during a sporting event and a soldier wounded in

combat. In both cases, the injured person often reports less pain than persons sustaining the same

injury in another context [153]. This variability in the perception of pain has been associated to

endogenous pain-modulatory systems, which are the neural basis by which attention, arousal,

expectation and cognitive processes alter the perception of pain. It has been suggested that these

networks modulate the transmission of nociceptive information to suppress or enhance

nocifensive responses to increase the chance of survival [309]. In the example of a soldier in

combat, suppressing pain reduces pain-related distraction, and facilitates escape from combat.

Conversely, the pain-modulatory systems can enhance sensitivity to pain, and enhance

nocifensive behaviour, which promote healing [309].

67

It has long been recognized that pain-modulatory mechanisms exist. In 1911, Head and Holmes

[417] proposed that the thalamus was the site of pain perception, and corticothalamic

connections modulated the nociceptive processing in the thalamus. Further, they suggested that

such inhibitory mechanisms were present at lower levels in the sensory pathway. Evidence

suggests that both sensory input and motor output can be modulated by supraspinal regions. For

instance, in 1915, Sherrington and Sowton demonstrated that, in a decerebrated animal, spinal

transection reduced stretch reflexes and enhanced flexor reflexes. In 1954, Hagbarth and Kerr

[389] demonstrated that supraspinal mechanisms do influence afferent (sensory) input from the

periphery to the spinal cord in the cat. Specifically, stimulation of M1, S1, S2, ACC, RF and

ventral anterior part of the cerebellar vermis depressed the dorsal root reflex, suggesting

modulatory effects of brain regions onto the spinal cord. It was also shown that cortical

stimulation could inhibit cutaneous sensory input to the VBSNC [e.g., 421]. In 1964, Kuypers

demonstrated that there are several tracts emerging from the brainstem that terminate onto the

dorsal horn [507]. However, the Gate Control theory of pain [595] (See Section 2.1.1.4) was the

first to propose that there were supraspinal regions that modulated nociceptive input from the

periphery at the level of the spinal cord.

Animal studies have demonstrated that there are, indeed, specific nuclei in the brainstem that

have a modulatory effect on dorsal horn nociresponsive neurones (for a comprehensive review,

see: [420,521,580,726]. The first study to demonstrate this was by Reynolds [730], in 1969. In

this study, electrical stimulation of the mesencephalic central gray matter in rats provided an

analgesic effect. Specifically, during an invasive surgical procedure, a laparotomy, the rats did

not show aversive behaviour, without paralysis. The potent analgesic effect of stimulating the

brainstem, specifically the periaqueductal gray (PAG), has been confirmed in the rat, cat, non-

human primate and in humans [580] [437]. This stimulation-produced analgesia is selective, in

that animals remain alert and respond normally to environmental stimuli, with the exception of

nociceptive stimuli – these responses are abolished [309]. Furthermore, if a noxious stimulus is

presented outside the region of analgesia, the animal responds with typical pain-related responses

[580]. Studies have demonstrated that stimulation-produced analgesia of the PAG is opiate-

based. Specifically, studies have found that injecting morphine, and other opiates induces

modulation [252,307,309,613]. Furthermore, naloxone, an opioid antagonist, abolishes the

analgesic effect when injected into the PAG [5,873,959] or the third ventricle [958]. The

68

analgesic effect of the stimulation of the PAG is mediated through a descending pathway from

the brainstem (via the rostral ventromedial medualla (RVM)) to the dorsal horn, and has been

demonstrated to inhibit nociceptive reflexes (e.g., the tail flick in rats), and nociceptive neurones

in the spinal and medullary dorsal horns [260,354,786,789].

It is noteworthy that descending pathways not only inhibit somatosensory input and pain

transmission, but can also play a facilitatory role at the spinal level [353,613,725]. It is believed

that this facilitatory role may be related to the chronification of pain [724].

2.3.5.1. Descending modulation pathways

There are several descending pathways that can modulate nociceptive input and pain

transmission. In general, these pathways are modulated by brain regions through corticobulbar

tracts and other tracts originating from subcortical nuclei that terminate in the brainstem

[95,309,420]. The best characterized of these pathways involves a circuit connecting the PAG to

nuclei within the RVM, including the nucleus raphe magnus (NRM), and the spinal dorsal horn,

via the dorsolateral funiculus [58]. Another modulatory region is the dorsolateral

pontomesencephalic tegmentum (DLPT), which connects to the cuneiform nucleus in the

midbrain, and the locus coeruleus/subcoeruleus (LC/SC), and other noradrenergic nuclei.

The PAG receives input from the ascending fibres from spinal lamina I, via the

spinomesencephalic tract [43,441,481,596] and the VBSNC [77,567,931], from the cerebral

cortex, the hypothalamus and other brainstem nuclei, such as the RVM, including the NRM, and

the nucleus cuneiformis [77]. There are direct projections from the PFC (with the exception of

the medial OFC), ACC and the insula to the PAG [406,405,413]. There are also some

corticobulbar connections to the PAG from M1 and the PMC (BA 4 and 6, respectively)

[405,619]. Furthermore, the amygdala has dense projections to the PAG. The amygdala, in turn,

is modulated by several cortical and subcortical regions. Furthermore, the PAG receives dense

input from the hypothalamus, and stimulation and opioid injection to the hypothalamus have

analgesic effects [433,566,731]. There are also cortical projections to the PAG via the anterior

pretectum [158,960]. In addition to its descending projections, and the afferent input from brain

regions to the PAG, the PAG has ascending efferents to MD thalamus and the OFC [135,168].

69

The RVM is another brainstem region essential for descending pain modulation. The RVM is a

group of nuclei that are the primary relay site for projections from the PAG to the spinal (or

medullary) dorsal horn [420]. The most important nucleus in the RVM in the context of pain

modulation is the NRM. For instance, inhibiting NRM neurones abolishes the analgesic effect of

PAG stimulation [10,71,355], whereas stimulation of the NRM produces analgesic effects by

inhibiting nociceptive neurones in the spinal dorsal horn [253,308,469]. The RVM projects to the

spinal dorsal horn via the DLF [58]. Therefore, the RVM is essential for the analgesic effect of

the PAG to occur. Therefore, the PAG is an important antinociceptive region, which receives

convergent input from ascending tracts, cortical regions and other brainstem regions. The PAG

projects to the RVM, which directly modulates nociceptive neurones in the STN and the spinal

dorsal horn.

Recent studies have also found that descending projections from the PAG to the spinal dorsal

horn and Vc, via the RVM, can also be facilitatory, i.e., these projections have a pronociceptive

effect, rather than an antinociceptive effect [654,835]. It has been suggested that these

mechanisms may contribute to the chronification of pain [704].

In addition to the PAG-RVM descending pathway, a parallel pathway exists in the brainstem: the

DLPT pathway [272,309,354]. The components of the DLPT include the nucleus cuneiformis,

which is adjacent to the PAG and is similar in function to the PAG in terms of pain modulation,

and pontine noradrenergic nuclear groups, including the LC/SC. It is believed that this pathway

modulates deeper laminae in dorsal horn, as it has moderate projections to these laminae,

compared to few projections to the superficial dorsal horn [166]. The LC/SC projects directly to

the dorsal horn via the ventrolateral funiculus, and electrical stimulation of these nuclei has an

inhibitory effect on nociceptive cells [726]. This pathway is separate from that of the PAG-RVM

pathway because inhibition of the RVM does not disrupt the analgesic effects of the LC/SC

[468].

2.3.5.2. Diffuse noxious inhibitory controls

It has long been known that a stimulus perceived as painful can diminish or mask a pain arising

from elsewhere in the body. For instance, Hippocrates stated in his Aphorisms: “of two pains

occurring together, not in the same part of the body, the stronger weakens the other” [424]. This

effect has been described as counter-irritation analgesia, and only occurs when “distant” body

70

parts are affected, i.e., extrasegmental body parts. In 1979, Le Bars and colleagues [520]

proposed the DNIC hypothesis. This hypothesis states that nociceptive signals can be inhibited

by nociceptive input from another body part, that projects to a different segment of the spinal

cord. To test this hypothesis, the sural nerve RIII reflex was recorded in subjects who were

presented with a painful “conditioning” stimulus (a hot water bath) [521]. Normally, when the

sural nerve is electrically stimulated, a painful pinprick sensation is felt in the dermatome of the

nerve and a knee flexor muscle nociceptive (RIII) reflex is evoked. It has been proposed that

DNIC is a spinal cord/STN mediated process [251]. However, when a noxious stimulus was

applied to the hand, the reflex and the perceived pain intensity of the sural nerve stimulation was

decreased. DNIC has also been demonstrated in animal models. For example, in the

anaesthetized rat, the C fibre reflex which is strongly inhibited by heterotopic noxious

stimulation [521]. It is believed that DNIC is a spinally-mediated process [521], and WDR

neurones, mainly in lamina V of the spinal cord, appear to be key for the effect of DNIC [519].

However, DNIC does not occur in animal models where the spinal cord is sectioned, or in

tetraplegic human subjects [739]. This suggests that supraspinal regions contribute to DNIC.

Specifically, there is evidence that the dorsal reticular nucleus in the caudal medulla contribute to

DNIC [519,896]. This nucleus receives afferent nociceptive input from the spinal cord, as well as

convergent input from PAG, RVM, RF, the hypothalamus, the amygdala, M1, S1, OFC, and the

insula [309]. The dorsal reticular nucleus projects to the thalamus, suggesting a role in both

nociceptive-signal transmission as well as antinociception [621]. Similarly to the descending

modulation pathways, it has been shown that DNIC may also have a facilitatory role for

nociceptive-signal transmission, but this is outside the scope of this thesis. For a review of these

concepts, see [546,895]. DNIC has also been demonstrated in the orofacial region [917].

Furthermore, a remote noxious stimulus can affect jaw muscle activity [131,132,894].

In sum, there are both negative and positive feedback loops that are consistently modulating

nociceptive-signal transmission to the brain. These loops are comprised of a network of

brainstem nuclei that are modulated by cortical regions. This delicate balance of inhibition and

facilitation can be very useful, based on the context of the nociceptive stimulus. Cortical regions

can upregulate the inhibitory effects in order to block out excess or unwanted stimuli.

Conversely, in a dangerous or threatening environment, when a subject is hypervigilant, the

facilitatory mechanisms can be upregulated as a protective mechanism.

71

2.4. Structural brain imaging

With the advent of functional neuroimaging, there has been a large increase of studies

investigating the function of the brain in health and disease. These studies have vastly improved

our understanding of brain function, and its relationship to perception, cognition and behaviour.

However, brain function is intimately related to brain structure, and so, studying neural structure

can provide a fundamental framework for understanding brain function. Aristotle, often

considered the father of comparative anatomy [145], described the basic tenet of physiology as

the structure-function relationship [100]. For Aristotle, function is what is of importance,

structure is secondary, and thus form and function are not separate disciplines, but are

complementary [100]. In the context of the brain, functional MRI (fMRI) has provided a wealth

of functional information, but the study of brain structure supporting brain function is only just

burgeoning [145]. Scientists have used the structure-function relationship to understand normal

function and abnormalities in disease. In a recent editorial, Catani [145] criticizes this approach

as reverse logic. He argues that form begets function, or, in evolutionary terms, the form comes

from the necessity for a function. Therefore, modification of form causes dysfunction. Only by

studying “form” (or anatomy) can we understand function, and dysfunction.

MRI technologies can be used to examine both gray and white matter structure in humans. These

techniques will be discussed below.

2.4.1. MRI

2.4.1.1. What is magnetic resonance?

A brief overview of the concepts related to MRI signal production, analysis and applications to

structural brain imaging will now be presented. For a comprehensive review of the basic

concepts of magnetic resonance (MR) and MRI, see [119,438] – the information reviewed below

was derived from these sources.

The concept of MR is based on the interaction of an atomic nucleus that possesses a spin and an

applied magnetic field (B). Nuclear spin angular momentum is an intrinsic property of an atom.

Most atoms have a nuclear spin, meaning that they are spinning at a constant velocity around an

axis. Nuclear spin angular momentum is a vector quantity, and it is conserved for different

elements. The human body is primarily comprised of water and fat, which contain hydrogen

72

atoms and so the 1H nucleus is an ideal choice for MRI. The 1H nucleus is a single proton

(termed the protium, and will be referred to as a proton herein), and has a spin of ½, and is the

most abundant hydrogen isotope (c.f. deuterium) for hydrogen. Also, its response to an applied

magnetic field is one of the most robust observed in nature.

A proton has a magnetic moment, which is parallel to the axis of rotation. The orientation of the

spin and the changes induced to it by a magnetic field provide the basis for the MR signal.

However, in MR, it is the population, ensemble or collection of spins (e.g., in a tissue) that is

key. The net magnetic moment of a population of spins (i.e., the vector sum of the proton

moments) is zero, and so no net magnetization is observed in a tissue. However, when the tissue

is placed into a magnetic field (B0), the protons begin to precess about the magnetic field. The

proton is slightly tilted away from the axis of B0, but the axis of rotation is parallel to B0. This

precession is caused by the interaction of B0 and the spinning proton. The rate of precession (ω0)

is proportional to the strength of the magnetic field and can be summarized by the Larmor

equation:

ω0 = γB0

where ω0 has units of Rad/s, B0 has units in Tesla (T; 1T is roughly five orders of magnitude

greater than the magnetic field of the Earth, and three orders of magnitude greater than that of a

refrigerator magnet), and γ is the gyromagnetic ratio, which is a constant for each nucleus. In the

case of protons, γ = 267.52 x 106 Rad/s/T (≈ 42.577 MHz/T).

Another feature of spins is that they have two possible states when they are in a magnetic field:

“spin-up” and “spin-down”. These are the two quantum-allowed states, although they can make

transitions between these two states, there are no intermediate states. These two states have

different energies associated with them (potential energy). The spin-up configuration is

associated with a greater potential energy than the spin-down configuration. Transitions between

the spin states require a specific energy, produced by molecular motion and collision. The energy

required for the transition to occur must be quantized – they must have the correct amplitude to

induce the “flip”. In an ideal system and at rest, there are an equal number of spin-up and spin-

down positions, and the net magnetization is zero. However, when the ensemble is placed in an

external magnetic field, the spins return to their lower energy level configuration due to

interactions with surroundings. The spin-down configuration is of lower energy, and is parallel to

73

B0, whereas the spin-up configuration is of higher energy, and is anti-parallel to B0. This effect is

called the Zeeman effect. Therefore, a small excess of the spins will try to flip to the lower

energy state. This process requires time because collisions are required to produce the quanta of

energy for the flip. Over time, the ensemble will reach an equilibrium state with the applied

magnetic field, and the transition probability of the spins becomes balanced because a small

excess of the spins have flipped to the lower energy state. This number of protons at each energy

level is governed by the Boltzmann distribution:

Nup/Ndown= e-∆E/kT

where Nup is the number of protons in the spin-up configuration, Ndown is the number of protons

in the spin-down configuration, and k is the Boltzmann constant, which is 1.381 x 10-23 J/K.

When a vector sum of spins in the transverse plane (perpendicular to B0) is performed, the net

magnetization is zero because of the random distribution of the position of the perpendicular

component of the protons’ precession. However, along the longitudinal axis (parallel to B0), there

is a net magnetization parallel to the field. This means that the tissue is polarized and magnetized

when placed in a field B0, and the level of net magnetization is M, which is parallel to B0. This

induced magnetization, M, is the basis of the MR signal.

A rotating frame is defined as a Cartesian plane rotating about an axis, which remains stationary.

If we assume that B0 is parallel to the z-axis, and that the xy-plane is rotating at the Larmor

frequency with the proton, then we observe what looks like a stationary spin. If we sum these

vectors, the resultant vector resembles a vertical magnetization. A time-varying magnetic field

creates an electric field. The created electrical signal is what we measure as the MR signal.

However, in the summated vector of M is not time-varying – recall that the transverse vectors

cancel each other, and the longitudinal vectors are stationary. To induce a time-varying magnetic

field, a new magnetic field (B1) is introduced at an axis perpendicular to B0 and M. B1 rotates at

the Larmor rate. M will rotate about the axis of B1 (and therefore perpendicular to B1,) at a given

frequency (ω1 = γB1). We can control the angle of rotation by setting the duration and amplitude

of B1. This rotation of the spins produces a time-varying magnetic field. For instance, if we rotate

the spins 90˚, then the spins are now precessing about B0 in the transverse plane. Given that a

rotating magnetic field induces a voltage in a loop of wire, if a loop of wire is placed around the

ensemble of the rotating spins, and attached to an amplifier, signal within the frequency range of

74

radio signals (3 kHz to 300 GHz) can be detected. The recorded signal is the summated

magnetization of the difference of the up and down spins of the entire population.

2.4.1.2. MR signals

The second magnetic field that is introduced to rotate M is a short pulse of radiofrequency (rf),

also called an excitation pulse. The excitation pulse is comprised of several frequencies within a

narrow range, or bandwidth. During the pulse, the protons in the sample absorb the energy of a

given frequency (ω0), and after a given time, depending on the T1, reemit this energy at the same

frequency. When the pulse is turned off, the protons begin to realign themselves to their stable

state, and they reemit the energy they absorbed in order to realign. As more protons undergo

relaxation and reemit their absorbed energy, the transverse magnetization of the protons creating

the signal is lost, and the signal decays.

Relaxation is a fundamental feature of MR, and relaxation time is measured for an ensemble,

rather than individual protons. For example, relaxation times for cerebrospinal fluid (CSF), gray

matter and white matter within the brain are different, and we can measure the relaxation time for

the entire tissue, rather than individual water, protein or fat cells. There are two relaxation times

that we can measure: T1 and T2.

Magnetization time (T1)

After the spins have been rotated, symbolized by M, the protons have a natural tendency to

return to their stable state – relaxation. The relaxation time “T1” represents the time required for

M to return to 63% of its original value (M0) before the rf pulse, in the z-axis (along B0). This

relaxation follows an exponential recovery function:

M(t) = M0(1-e-TR/T1)

where TR is the time between the rf pulse (repetition time), and T1 is a constant and is specific to

the tissue. T1 represents the time it takes for tissues to relax after magnetization. This relaxation

is named a “spin lattice” relaxation. The motion of the surrounding spins provides a fluctuating

magnetic field that enables spins to be flipped back to their stable state.

The mobility of atoms varies based on the size of the molecule it is part of, and its interactions

with the immediate environment. The consequence of this is that different tissues magnetize at

75

different rates. For instance, smaller molecules, such as water (the main component of CSF),

have a very long T1 (up to several seconds), whereas larger molecules, such as fat (which is the

main component of myelin, and thus white matter), have a very short T1 (less than a second).

Grey matter has an intermediate T1, between that of CSF and white matter. These differences in

T1 provide a useful contrast mechanism in MRI.

Signal decay time (T2)

The relaxation time “T2” refers to the time required for the transverse component of M to return

to 37% of its initial value. When M is rotated into the xy-plane, the spins of the protons are in

phase. However, over time, the signal decays because the spins begin to dephase. This loss of

coherence occurs because of molecular interactions with the protons, and because of the transfer

of energy from one spin to another. The signal decay can be summarized by the equation:

! ! = !"#(!)

!!!!!!!!!!!!!!!!= !!!!!"/!!

Where “TE” is time from excitation (or echo time), and is an adjustable parameter that represents

the time after the rf pulse during which the signal data are collected. T2 is also tissue specific,

and is slower in small molecules (the signal from CSF, which is mostly water, decays over

several seconds), and faster in larger molecules (e.g., gray matter and white matter, where the

signal decays in less than a hundred milliseconds). Again, differences in T2 allow for a contrast

useful to MRI. Further, by increasing TE, the contrast between tissues can be increased but less

signal is actually received as it decays.

Spatial Encoding

The creation of an image requires a process that encodes signals in tissues in 3D space. The

Larmor relation states that the signal’s frequency is related to the magnetic field affecting the

protons. For instance, if we maintain a homogenous magnetic field, B0, in the bore of an MRI,

there would be a single frequency present in the MR signal. However, if we vary the field with a

gradient coil, then frequency encodes location. To do so, we need to have different frequencies

representing different points in space. Therefore, we introduce a gradient pulse that distorts the

magnetic field, and produces a gradient in the magnetic field. The effect is that at one end of the

76

gradient there is slightly less magnetic field strength (B < B0), in the middle there is still a net

magnetic field strength, which is equal to B0 (B = B0), and at the other end there is a slightly

higher magnetic field strength (B > B0). In a simplified model, this will affect the frequency of

the nuclear spins – they are precessing at a slower rate at the lower end, and precessing faster at

the higher end, compared to the middle, where B = B0. This is done in three orthogonal planes to

localize signals in all three dimensions. The relationship between frequency and B, allows us to

encode spatial position as a function of precession frequency.

Contrast

Given that the relaxation times, T1 and T2, are specific to tissues, and that there are parameters

that are specific to each relaxation time’s contrast (TR and TE, respectively), we can vary these

parameters to adjust the contrast in MR images. The signal we obtain be summarized by:

where ρ is proton density, k is an instrument specific calibration factor.

We can experimentally adjust TR, which affects the T1 contrast. Alternatively, we can adjust TE,

which affects the T2 contrast. If we set low values for the TR and the TE, then the contrast of our

image will be T1-weighted. However, if we have a long TR and TE, the image will be T2-

weighted. Water molecules that interact with small molecules have relatively low signal in a T1-

weighted image, and relatively high signal in a T2-weighted image. Conversely, water molecules

that interact with larger molecules have relatively high signal in a T1-weighted scan, and a

relatively low signal in a T2-weighted scan. Therefore, in a T1-weighted scan white matter is

bright, and CSF is dark. The opposite occurs in a T2-weighted scan: white matter is dark and

CSF is bright.

In summary, rf pulses convert nuclear magnetization to signal. Rf pulses add energy by

displacing longitudinal equilibrium. The contribution of intrinsic tissue properties, T1 and T2,

can be manipulated by adjusting the timing of TR and TE, respectively.

77

2.4.2. Gray matter imaging

Structural studies of the human brain used to be limited to post-mortem investigation and

sometimes biopsies (e.g., [844]). However, for the past few decades, it has been possible to use

MRI to examine brain structure in living humans non-invasively. Structural MRI can be used to

measure many structural features including the thickness of the cortical mantle, the gyrification

index, and other measures of interest, at a very high spatial resolution: the millimeter scale. As

newer technologies emerge, even greater spatial resolution can be achieved, but these are outside

the scope of this thesis (for a review, see: [282]).

Structural MRI analysis of gray matter is commonly performed on T1-weighted images, as these

acquisitions can be optimized to increase the contrast between gray matter and white matter. In a

T1-weighted image, the CSF is dark (almost black, due to the long T1 relaxation time), the white

matter is bright (due to the myelin sheath surrounding axons) and the intensity of gray matter is

between that of CSF and white matter. Many automated and semi-automated techniques have

been developed to characterize gray matter differences within a population and between

populations. These measures include the shape of gray matter structures, gray matter density (the

relative amount of gray matter within a voxel) and cortical thickness. Brain imaging software

packages typically include toolboxes that use algorithms to segment the three tissue classes in the

brain (gray matter, white matter and CSF), and perform statistical analyses. Currently, the two

most common methods are voxel-based morphometry (VBM) and cortical thickness analysis

(CTA). Specific information about the methods and the relative advantages and limitations of

each are discussed below.

2.4.2.1. Voxel-based morphometry

One of the first automated toolboxes to analyze gray matter was Christian Gaser’s VBM toolbox,

implemented in the Statistical Parametric Mapping (SPM). A VBM analysis of brains from a

group of subjects includes the following stages: 1) segment high-resolution brain images into

gray matter, white matter and CSF; 2) spatially normalize the brains to a common standard

stereotactic space; 3) apply a spatial filter to “smooth” the gray matter segments (this effectively

blurs the gray matter to improve alignment) [36,37]; and 4) perform voxelwise parametric or

non-parametric statistics. Using either statistical method, the results are corrected for multiple

comparisons. More recently, Jacobian modulation has been added to the processing pipeline

78

prior to the smoothing step so that findings can be considered as volumes, rather than relative

gray matter density [36,371]. To do so, for each brain the software calculates the deformation for

normalization to the standard space (stereotactic) template – a Jacobian determinant, and then

modulates each voxel in standard space to account for the expansion that occurs. VBM is a

widely used tool in neuroimaging, and its popularity can be attributed to its relative ease of use,

and the little processing power required. However, this technique has been criticized because it is

sensitive to misalignment, especially in the cortex [104,210]. Therefore, results in cortical

regions should be interpreted with caution. Nonetheless, the alignment of subcortical gray matter

regions is robust, and thus more reliable.

2.4.2.2. Cortical thickness

It has been suggested that cortical thickness (rather than gray matter volume) may provide a

better measure to examine changes or variability of the cortical mantle and so a technique called

CTA was developed. CTA measures the thickness of the cortex in millimeters, which is a well-

defined physical measure. However, CTA is limited to the cortex, as implied by its name, and

thus we must use it in conjunction with VBM to get a full picture of gray matter in the brain.

Nonetheless, CTA is purportedly superior to VBM because of improved alignment of cortical

gray matter, especially gyral folding patterns. This method inflates the cortical gray matter sheet

into a sphere, and aligns the gyri and sulci based on landmarks. Several software packages offer

a CTA toolbox, and the most common one is FreeSurfer. The analysis pipeline relies on some of

the basic principles of VBM, but with a few differences. First, high-resolution T1-weighted brain

images are aligned to a different standard stereotactic space than the one used in VBM, which

uses the Montreal Neurological Institute’s standard template of 152 brains (MNI152). FreeSurfer

uses the Talairach and Tournoux [844] brain as a template for normalization. Then, the brains

undergo intensity normalization to remove coil-related inhomogeneities. The image is then skull-

stripped, and the subcortical structures are identified and labeled, which aids in the segmentation

of white matter from gray matter. The pial surface is then defined, and the hemispheres are

separated. The software then measures the distance at every point between the edge of the white

matter surface and the edge of the pial surface, which represents the thickness of the cortex. This

cortical sheet is then parcellated and the individual gyri and sulci are identified and labeled using

an atlas that is based on manual parcellations [243,318,321]. Parametric univariate statistics can

be used to perform group-level analyses on cortical thickness. To do so, the brains must first be

79

aligned to one another. This is done by inflating the cortical surface into a sphere, and iterative

alignment algorithms, which use landmarks and the individual parcellation to align specific sulci

and gyri [319,320]. These steps allow for better alignment of the cortex, and thus overcome the

limitations of VBM.

2.4.3. White matter imaging

DWI with MRI can be used to non-invasively evaluate human white matter. The study of white

matter paths is called brain hodology, and comes from the Greek hodos, which means path. Brain

hodology in humans has mainly been on post-mortem studies of white matter, which date back to

antiquity. Notably, in the seventeenth century, Steno stated:

To say that the white matter is but a uniform substance like wax in which there is no hidden contrivance, would be too low an opinion of nature’s finest masterpiece. We are assured that wherever in the body there are fibres, they everywhere adopt a certain arrangement among themselves, created more or less according to the function for which they are intended [translated by 760,830]

Neuroscientists in the nineteenth century then attempted to elucidate structure-function

relationship [145,306]. However, these scientists recognized that local structure-function

relationships could not fully explain dysfunction in the brain; they understood that distant brain

regions interacted to produce behaviour. This is exemplified by Wernicke’s statement that

sensory and motor regions must work together through connections (i.e., white matter tracts) in

the brain [145]. Déjérine was a French neurologist, medical historian, and neuroanatomist, whose

investigation of white matter in the brain is the most comprehensive of the time – and some

would argue until today [607]. He agreed with Wernicke, proposing that the key to

understanding brain function relied on understanding the connections of association fibres within

the brain [145,237]. More recently, the concept of studying the anatomic basis of behaviour, and

functional relevance in anatomic circuits has reemerged [147,760], notably by Geschwind

[359,360] and Kuypers (e.g., [387,663]). Geschwind proposed that “brain function is the result of

effective communication between structures geographically distributed around the nervous

system”[761]. Recently, Mesulam [607] suggested that the function of a neurone is defined by its

connections, and that understanding the connection pattern of the cortex is the only way to

understand how the brain functions. Therefore, in addition to functional and gray matter studies

of the brain, elucidating the connectional anatomy, and the integrity of these tracts in the brain

80

will provide essential insight into the connections (or disconnections) of patients with chronic

pain. The advents of DWI and diffusion tensor imaging (DTI) have provided a method to

essentially perform in vivo dissections of the white matter in humans [60], and to elucidate

abnormal white matter structure related to clinical correlates in disorders, such as chronic pain.

Furthermore, of particular relevance to this thesis, is that DTI (see below) can also be used to

examine peripheral nerves [429], such as the CNV in trigeminal neuralgia [525] (discussed

below and in Chapter 7).

2.4.3.1. Diffusion tensor imaging

DTI is a mathematical model that is fit to DWI data. DWI is a brain imaging technique that can

be used to visualize white matter. Specifically, DWI is an MR modality that is sensitive to the

diffusion of water molecules [63]. The following description of DWI and DTI are based on

information from: [52,60,62,63,67,522,524,626,660]. Diffusion is the process of molecules

moving, and is the result of Brownian motion. All molecules, when suspended in a liquid or gas,

show random translation movements (i.e., Brownian motion), resulting from thermal energy of

the molecules. The diffusion of water molecules can be summarized by Einstein’s equation:

X2= 2DTd

where X2 is the average mean-squared diffusion distance (in m2) along one direction, Td is the

diffusion time (in s), and D (in m2/s) is the diffusion coefficient. DWI’s ability to discern

microstructural features of a tissue is based on the notion that water molecules sample the

microscopic environment at a very high resolution (much higher than the average MRI

resolution), roughly 10 µm per 50 ms [524]. This sampling includes collisions, interactions and

crossing of water molecules with cell membranes, fibres and other macrostructural features, such

as macromolecules. A DWI provides a summary of the bulk motion of a diffusion molecule

within a voxel, which can reveal structural and geometric organization of a tissue [60,523]. The

molecule of choice in human DWI is the water molecule, because 90% of the protons in the

human body are located in water molecules. Diffusion occurs in all three dimensions of space,

and when unhindered, will diffuse equally in all directions (the expected vector-sum of

displacement is zero), a process known as isotropic diffusion. When diffusion is hindered, e.g.,

by a tissue, we observe anisotropic diffusion. Anisotropic diffusion of water has been observed

in muscle, spinal cord and cerebral white matter [524]. Einstein’s equation does not apply to

81

anisotropic diffusion, and we have to use a symmetric tensor, , to describe diffusion in this

scenario:

It is noteworthy that in the scenario of anisotropic diffusion, we assume that diffusion is

Gaussian in all directions.

Anisotropic diffusion in brain white matter has generated much interest in understanding the

connectional anatomy of the brain. Specifically, water molecules within or between white matter

tracts have anisotropic diffusion, as the bulk diffusion is parallel to the main axis of the fibre

tract. This anisotropy comes from the geometrical structure of the axon – it is tubular in shape,

and many axons form a fibre tract. Specifically the diffusion of water is restricted by cellular

barriers, such as myelin, cell membranes, and macromolecules (such as microtubules and

neurofilaments), and the molecules tend to diffuse along the primary axis of the tract, whereas

diffusion perpendicular to this axis is impeded. MRI acquisition parameters can be adjusted to

acquire images that provide information about the diffusion of water, and thus, the connectivity,

integrity and orientation of white matter in the brain (see below). This thesis will briefly review

the basic principles of DWI signal generation and image acquisition (for a comprehensive

review, see: [522,524,626]).

MRI can be used to acquire several different types of image contrasts (T1, T2; see section

2.4.1.2). As described above, the MR signal equation can be summarized as:

S = kρM0 (1-e-TR/T1) e-TE/T2

where S is the acquired signal, TR is the repetition time, TE is the echo time, ρ is proton density

and T1 and T2 represent relaxation times specific to tissues. This equation can be modified to

include a diffusion term, bD:

S = kρM0 (1-e-TR/T1) e-TE/T2 e-bD

D

D =

Dxx Dxy Dxz

Dxy Dyy Dyz

Dxz Dyz Dzz

"

#

$ $ $

%

&

' ' '

82

where the b, or b-factor, summarizes all of the gradient effects, and D is the diffusion tensor. To

measure the diffusion of water molecules, we must acquire two different b-values.

Therefore, for Sb we obtain:

!! = !"!!(1− !!!"!!)!!

!"!!!!!"

and for Sb0, we obtain:

We can then solve the equation for D:

where Sb is a signal acquired from the scanner with a given value for b and Sb0 is the signal

acquired with a b-factor equal to zero. The signal acquired from b = 0 is not sensitive to

diffusion, and is referred to as a non-diffusion-weighted image. All of the other terms in the

signal equation are kept constant, and so cancel out.

The signal (S) that we acquire is generated. To do so, we apply a phase encoding gradient pulse

along a particular direction, using different combinations of the x, y, z gradients. After the rf

pulse is applied (for a given duration), the protons which are precessing about the z-axis begin to

dephase. We then apply a second re-focusing pulse, which is the same as the first pulse, but in

the opposite gradient direction. This paired-pulse sequence sensitizes the signal to diffusion.

Specifically, the refocusing pulse can only rephase the protons if they have not moved. If the

water molecules diffuse (i.e., change location between the first and second rf pulse), then this

motion will disrupt the phase-gradient. This imperfect rephasing is measured as signal loss. As

previously mentioned, by using two different b-values we can acquire two different signals,

which allows us to solve for D. The b-value is related to γ, the strength of the gradient pulse

(Gi), the length of time we apply the gradient (δ), and the time between the application of

gradients ∆). These relationships can be summarized by the following equation:

D =− ln Sb Sb0( )

b− b0

83

which can be simplified to:

This equation demonstrates that the simplest way to modify the b-value is by manipulating the

time between the two gradient pulses.

In a DTI experiment, a symmetric b-factor tensor (b-matrices) is calculated for each DWI

acquired with a different b-value (b ≠ 0 and b = 0). Usually the b ≠ 0 is acquired at a b-value

equal to 1000 s/mm2. We then solve for D and the result is all 6 components of tensor model,

which can be summarized in an image of the apparent diffusion coefficient (ADC). Therefore,

the signal ratio equation becomes:

The intensity of a voxel in the ADC image represents the direction of the bulk diffusion of water

molecules with regard to the direction of the applied gradient. In the ADC image, the brighter the

voxel intensity, the more diffusion has occurred. However, complex features in microstructure,

such as crossing fibres, can produce erroneous measures of diffusion. Thus, to properly

determine the principal diffusion direction within a voxel, an enormous number of directions

should be sampled. However, this is not a feasible approach, as MRI scanning is costly, the

hardware is limited, and the subject cannot undergo such long scan durations. Therefore, to

overcome these limitations, a tensor model is used to calculate the bulk direction of water

diffusion within a voxel. In a tensor model, diffusion only needs to be acquired along six

directions; however, the more directions acquired, the better the signal-to-noise ratio of the

image, which will improve the precision of the DTI estimation, and hence the angular resolution.

Some imaging studies acquire up to 120 directions, although 60 directions has now become

standard practice. The tensor (D) is a positive, definite tensor, and describes a three-dimensional

ellipsoid in displacement space, called the diffusion-ellipsoid. The three diagonal terms of the

tensor model (TX, Dyy and Dzz) represent the three axes of the ellipsoid. These relative

b = γ 2 G t'( )dt '0

t∫[ ]

2

dt0

te∫

b = γ 2Gi2δ 2 (Δ−δ 3)

ln SbSb0

"

# $

%

& ' = −b(ADC)

84

diffusivities of water along these axes can be expressed by three eigenvalues (λ1, λ2, λ3), and

their orientation is represented by three corresponding eigenvectors (ν1, ν2, ν3) (see Figure 2-4).

For a comprehensive review of the tensor model, see: [60,61]. These values allow us to calculate

several useful measures (or contrasts), such as mean diffusivity (MD), fractional anisotropy (FA),

and radial diffusivity (RD) [62,494,524,626] (for the biological basis of these, see 2.6.4), and

their respective equations are presented in Figure 2.4.

Figure 2-4: Schematic of diffusion tensor imaging, and summary of eigenvectors and metrics.

Lambda (λ) represents each of the eigenvalues that make up the tensor model.

In DTI analysis, FA is the most commonly used measure, as it is scaled between 0 and 1, where

0 represents complete isotropic diffusion, and 1 represents complete anisotropic diffusion.

Assuming that these measures represent white matter integrity along a tract (see below), then we

can evaluate white matter integrity within a population, or compare between populations (see

next section). By combining FA as a threshold and the primary eigenvector, ν1, then we can

determine the orientation of white matter tracts, at each voxel.

Despite the great advances that DTI has afforded us, there are a number of inherent limitations

and challenges that must be overcome in data acquisition and analysis. For instance, as DTI

85

measures the relative displacement of water molecules, it is very sensitive to head motion and

physiological motion. To minimize head motion, the subject’s head is padded, and to minimize

physiological motion, some investigators measure respiration rate and cardiac rhythm during the

scan, and control for these during analysis. Furthermore, to correct for relatively small amounts

of motion during a scan, some scanners collect several non-diffusion-weighted images at several

points during the acquisition, interspersed between the diffusion-encoding runs (usually at a ratio

of 5:1 for diffusion-encoding directions: non-diffusion-weighted images). These non-diffusion

encoding directions are aligned to one another, and the transformation matrix is applied to the

DWIs between these images. Other motion correction algorithms also exist (for a review of

these, and other denoising methods, see: [694]). Another source of noise is the production of

eddy-currents artifacts, which occur due to residual gradient effects and the rapid switching of

gradient pulses. Most motion-correction toolboxes also correct for these artifacts.

There are some notable limitations with the diffusion tensor model. For instance, some voxels

may appear to have low FA values despite their location in white matter. We would expect

voxels in gray matter and CSF to have low FA, as there are few barriers and little geometric

organization that restrict the diffusion of water. Regions of complex white matter organization,

such as crossing fibres or fibre dispersion, cause these relative hypointensities of FA in white

matter [14]. Another limitation of DTI is that it cannot resolve the direction of the fibre, but only

its orientation in three-dimensional space [14]. Therefore, one cannot make any directional

interpretations about whether a tract is connecting to or coming from a gray matter region.

2.4.3.2. Between Group Comparisons of White Matter: Tract-based spatial statistics

One of the major challenges in applying DTI to study clinical populations has been the

development of an approach to compare white matter metrics between subjects, given the

intersubject variability, but now, several approaches have been developed. For instance, some

groups use an average value (or summary measure) of FA, MD and RD across the whole brain

[112,152]. Yet, summary measures of white matter integrity are not very informative, as there is

no information about the location of the white matter abnormality. This method is useful when

investigating a diffuse disease, such as multiple sclerosis [722]. Other groups have used a VBM-

style approach (see Section 2.4.1.2) to delineate regions of white matter abnormality between

subject cohorts, or to identify regions where white matter correlates with a behavioural measure

86

within a group of subjects [36,37]. However, VBM-style analyses have been heavily criticized

for their lack of reproducibility (due to the parameters that can be selected) and its lack of

accuracy (due to misalignment) [151,210,813]. In the former case, one example is smoothing.

Two studies have shown that results vary largely based on the selection of the size of the

smoothing kernel [466,667]. Another limitation of VBM-style white matter analysis arises when

aligning white matter tracts for comparison between subjects. To obtain a reliable measure, it is

essential to compare the same tracts between subjects, and to compare the same part of the tract

(as FA is not the same at the edge of a tract, compared to the centre of the tract), the so-called

correspondence issue [813,815]. These issues have been encountered in several studies (e.g.,

[80,804,822]. Another white matter analysis approach has been to confine the analysis to specific

white matter regions-of-interest (ROI). ROI analyses of white matter can be beneficial, in that

they are very sensitive to changes in measures of white matter integrity, so long as the ROI is

specific. Furthermore, this approach is not very computationally intensive. However, delineation

of an ROI requires expert knowledge of the anatomy, in order to identify appropriate landmarks

to draw reliable and reproducible ROI [151].

To address some of the issues of these white matter analysis methods, Smith and colleagues

developed tract-based spatial statistics (TBSS), a toolbox implemented in the Oxford Centre for

Functional MRI of the Brain’s (FMRIB) software library (FSL) [813,815,818]. TBSS is a fully-

automated program that has a similar pipeline to that of VBM, but it has been optimized for

white matter analysis. TBSS uses FA to align subjects to a common template using a non-linear

algorithm. A standard FA template (FMRIB58_FA) with 58 subjects in the MNI152 standard

space was developed for this purpose. Alternatively, FSL allows the creation of a study-specific

template for populations where a standard-space template is not suitable, such as a pediatric

population, or in the case of severe pathology. TBSS then calculates a group mean FA map, and

“thins” this map to create a skeletonized mean FA image (see Figure 2-5). This skeletonized FA

map represents the centre of the white matter tracts common to all subjects. The highest FA

value of each subject’s tract, deemed to be the centre of the tract, is projected onto the skeleton at

every voxel, in order to overcome any residual misalignments. Voxelwise univariate statistical

tests can then be performed on the skeletonized FA map. Furthermore, once alignment is

achieved, other indices of white matter integrity, or shape can be projected onto the skeleton for

analysis. Therefore, TBSS effectively addresses the limitations of the other methods described.

87

However, restricting the analysis to the peak FA value may reduce sensitivity to changes outside

the centre of the tract.

Figure 2-5: White matter skeleton from tract-based spatial statistics. The skeleton is shown in

green, and is overlaid onto the MNI152 standard brain in FSL.

2.4.3.3. Tractography

Another method to evaluate white matter in the brain with DTI is tractography. In general, this

method allows us to perform in vivo “dissections” of white matter tracts (i.e., visualize the tract),

and to delineate the anatomical connectivity of the brain. The concept of tractography is based on

the signal information collected in a DWI scan and the tensor model utilized in DTI. Briefly,

when we apply the tensor model to a DWI data set, we obtain eigenvalues and eigenvectors for

each voxel in the image, which represent the orientation and magnitude of water diffusion within

the voxel. When diffusion is restricted, by a biological tissue for instance, then the diffusion is

anisotropic. We can use the orientation information within each voxel to reconstruct white matter

pathways in the brain, called tractography. The fundamental assumption of white matter

tractography is that the primary eigenvector (υ1) is parallel to the primary axis of the white

matter tract. It is, however, noteworthy that this assumption is often violated in regions of

complex white matter (e.g., crossing fibres and fibre dispersions). More complex algorithms

have been developed to address these issues, and to increase the sensitivity and reliability of DTI.

Two of the most commonly used methods, deterministic and probabilistic tractography, are

described in further detail below. Although there are other, more complex methods being

developed, these are outside the scope of this thesis.

88

Streamline (deterministic) tractography

Deterministic tractography algorithms use the orientation information within voxels to track the

path of white matter tracts [14]. The resultant image is a tractogram, which is based on the

primary (or major) eigenvector within each voxel. The tractogram is initiated from a seed

location (a single voxel, or a group of voxels of interest). Because the algorithms use a threshold

for FA to distinguish between gray and white matter, the seed is usually located within the white

matter [14]. If the seed were in the gray matter, or in a voxel with FA below the threshold, then

tracts could not be generated. Several restraints can be placed on the algorithm to remove

spurious connections, and to ensure that the tracts are consistent with the predicted pattern of

connectivity [14].

There are several deterministic tractography algorithms that can produce tractograms. Most of

these methods utilize a streamline approach. This approach calculates the stepwise progression of

the tract (or the path), beginning from a seed. The progression of the path is based on the

direction of the primary eigenvector. The termination of a tract is based on an arbitrarily selected

parameter [465]. For instance, an FA value is selected where the primary eigenvalue is well

defined, e.g., FA ≥ 0.15. Another parameter used to terminate a tract is the curvature angle,

which is also arbitrarily chosen. The specific aspects of deterministic algorithms are outside the

scope of this thesis, but for a detailed review of these methods see: [14].

There are both advantages and disadvantages to deterministic tractography. The most significant

advantage is that fewer diffusion-encoding directions are required to produce tractograms,

compared to other algorithms, such as probabilistic tractography, or diffusion spectrum imaging

(see below). Furthermore, the analysis is not as computationally intensive as some of the other

methods (including probabilistic tractography). However, deterministic tractography can only be

used as a qualitative means of visualizing tracts in the brain, although the images can serve as

masks to perform quantitative analysis, e.g., evaluate measures of white matter microstructure,

such as FA. The images produced with deterministic tractography are visually appealing, but this

can instill a false sense that the visualized tracts are accurate. However, there are many potential

sources of errors in deterministic tractography. As previously discussed, DTI is very sensitive to

motion and other sources of noise. Even small perturbations in the image can lead to tracking

spurious connections. The largest source of error in tractography is tract dispersion, or the

89

variance of trajectories, which increases from the seed because there are an increasing number of

potential paths [14]. Another source of error comes from tract deviation, or errors in tract

position. These are related to the step size, especially in regions where the curvature of the tract

is high. In general, a larger step size will increase the chance of error [518]. Another source of

error in deterministic tractography is that of crossing fibres, as discussed above. Crossing fibres

can lead to false-negatives in the tractograms [145]. For instance, most deterministic algorithms

are not able to resolve tracts within the CST emerging from the lateral motor cortex and

projecting to the pyramids in the brainstem [14]. This is the result of the crossing fibres in the

centrum semiovale, which contains association fibres, callosal fibres and short U-shaped fibres,

which project in different directions. As a result of the multiple fibre directions in voxels within

this region, the primary eigenvalue is diminished, and the tractography algorithms fail to track

the rest of the tract. Therefore, experiments that seek to resolve tracts using deterministic

algorithms should have strong hypotheses and predictions based on known anatomical pathways,

from animal studies and human post-mortem studies.

Probabilistic tractography

Probabilistic (or stochastic) tractography is a method that attempts to address some of the

limitations of deterministic tractography. Specifically, probabilistic tractography acknowledges

that there are inherent errors that occur in tract-tracing algorithms, and accounts for this

variability. Therefore, the inherent variability in the data is included in probability estimates of a

tract [668]. To do so, algorithms use probability density functions (PDF), which considers the

expected distribution of the possible fibre orientations, at the level of the voxel [671]. The PDF

are used instead of discrete measures of fibre orientation. An advantage of probabilistic

tractography, compared to deterministic tractography, is that the PDF allow tractography to

continue in a region where deterministic tractography would normally stop. Another advantage

of probabilistic tractography is that we can quantitatively compare the connectivity of tracts

based on the probability of connections.

The PDF is comprised of three orientation density functions (ODFs): the diffusion ODF (dODF),

the fibre ODF (fODF) and the uncertainty ODF (uODF) [72]. The dODF and the fODF are

biophysical properties of the tissue that is being measured. The fODF describes the proportion of

the fibres in each direction. For instance, as in the example above, we have two orthogonal fibres

90

in a voxel, the resultant fODF will appear as a cross, in the direction of the fibres [862]. The

fODF is useful in that it contains a proportion, which can be used to estimate the connectivity of

a tract connecting A to B, providing a quantitative method that is both useful and biophysically

meaningful [72]. However, DWI measures the diffusion of water within each voxel in the brain,

and not the proportion of fibres, and so the measure we obtain is dODF, a measure of the

orientation of diffusion within a voxel [879]. If water molecules diffused exactly along the axis

of a tract, then the dODF would be exactly equal to the fODF. However, in reality, water

molecules diffuse primarily along the axis of the tract, but there is also diffusion in other

directions. Therefore, the dODF is necessarily broader than the fODF. Whereas the dODF and

fODF are properties of the tissue which can be measured, the uODF is a measure of the error that

can be predicted, or as Behrens puts it: “it [uODF] is a function that describes our belief” [72].

Specifically, the uODF represents the uncertainty that is included in the data because of the noise

inherent to it.

These variables have very important implications in tractography. It can no longer be assumed

that the primary eigenvector is the path of the tract, as there are an infinite number of paths

within each voxel, each associated to a probability of being the correct orientation. To calculate

the probability of connection of two regions, every possible path and its probability must be

considered, and then the probabilities associated with each path connecting the two regions is

summed [75]. Rather than attempt to solve this equation, more reasonable approaches have been

developed that allow us to sample the probability distributions [75,392,465,669-671]. Briefly,

these algorithms sample many paths from a seed at sub-voxel to sub-voxel steps (or for lower

resolution, or larger step size, at a voxel to voxel steps). Each step considers the distribution of

possibilities, resulting in many possible paths, each with a probability. These results are

summarized within the voxel as a proportion of the number of samples that pass through the

voxel [465]. Whereas deterministic algorithms rely on arbitrary FA values or curvature angles to

stop tractography, probabilistic algorithms do not rely on these parameters for termination. In

contrast to deterministic tractography, this allows the algorithm to propagate even when a seed is

in gray matter. The criterion for termination of a tract is usually the angular deviation between

successive steps (c.f. FA values in deterministic tractography), which prevents a tract from

looping back onto itself, or if a streamline enters the same voxel more than once [75,465]. The

91

specific differences between probabilistic tractography algorithms are outside the scope of this

thesis. For a comprehensive review, see: [72,671].

The results of probabilistic tractography should be interpreted with care. These results have a

very specific meaning: a calculated probability is the probability that a connection from a given

seed to a target through the diffusion data exists within the context of the model for the diffusion

signal and the model of connections. These tracts do not provide anatomical evidence for the

existence of a path within the brain. However, if a hypothesis is based on known anatomical

pathways from post-mortem studies in humans (and homologues from tracing studies in

monkeys), then we can infer that the tract we have identified is “real”. Otherwise, we must

interpret the presence or absence of a tract with caution.

Crossing fibres are still a consideration in probabilistic tractography; in fact, it is a shortcoming

of the tensor model. Some groups have developed alternate, non-parametric models to replace

the tensor model, that can better resolve crossing fibres, such as diffusion spectrum imaging

[878,925], q-ball imaging [877,879], spherical deconvolution [862], persistent angular structure

MRI [456], and diffusion orientation transform [661]. However, these are outside the scope of

this thesis (for a review, see: [15]). Some groups have tried to resolve crossing fibres within the

tensor model. The most common approach is to model more than one tensor per voxel: the multi-

tensor model [73]. The number of “fibres” that can be modeled per voxel depends on the signal-

to-noise ratio, and the number of diffusion-encoding directions.

In sum, there are several advantages and limitations to tractography. Nonetheless, when care is

taken to acquire, check and de-noise the data, and seeds are selected carefully, based on known

landmarks, and results correspond to known tracts within the brain, then tractography can be a

very useful and informative tool. Additionally, tractography can be qualitative (visualizing tracts)

and quantitative (calculating the probability of a connection). The latter allows group

comparisons of connectional anatomy.

2.5. Brain imaging of pain

2.5.1. Functional neuroimaging of pain

Functional neuroimaging has confirmed and advanced our understanding of brain regions

implicated in nociception and pain modulation – especially the role of the PFC. With the

92

exception of the PFC, the pain-related brain regions identified with functional neuroimaging are

similar to the brain regions previously identified in animal models and human

electrophysiological studies. These regions, however, are implicated in sensory, motor, attention,

salience, cognitive processes, decision-making, and other functions. Therefore, further research

is required to understand the role of these regions in pain perception and modulation, and further

to determine how pain is integrated.

Functional brain imaging encompasses a number of modalities, including positron emission

topography (PET), single-photon emission computed tomography (SPECT), fMRI, and

magnetoencephalography (MEG). Also, electroencephalogram (EEG) is sometimes considered

as a brain imaging technique when it is coupled with visualization of the brain areas activated.

Each of these modalities has its benefits and shortcomings because of spatial and temporal

resolution, fundamental principles (e.g., indirect hemodynamic measures, direct electrical

measures, etc.), invasiveness, availability, cost, and other scientific and practical issues. Because

this thesis focuses on structural abnormalities in the brains of patients with TMD, findings from

functional neuroimaging will only be briefly described.

2.5.1.1. Imaging acute pain

The first two brain imaging studies to investigate brain regions involved in nociception and pain

perception and modulation in healthy subjects were performed using PET [464,846]. One study

reported that heat pain, compared to an innocuous warm stimulus, evoked an increase in regional

cerebral blood flow (rCBF) in the contralateral thalamus, lentiform nucleus and ACC (MCC,

with the current nomenclature) [464]. The other study reported increased rCBF in S1, S2 and the

MCC [846]. The first fMRI study to investigate the neural correlates of pain perception found

that painful electrical stimulation increased the blood-oxygenation-level-dependent (BOLD)

signal in S1 and MCC [231]. Many studies have since investigated pain-related brain activation,

and some have specifically attempted to identify the neural correlates of the different dimensions

of pain. Several reviews and meta-analyses have summarized these findings

[23,95,143,212,274,304,690,709,714,869]. In general, brain imaging studies of pain have

reported the activation of S1, S2, MCC, insula, prefrontal and motor regions. Relevant to this

thesis, two studies have examined BOLD correlates of pain in the orofacial region. One study by

Ettlin and colleagues [300] reported that electrical stimulation of the tooth in healthy subjects

93

was related to BOLD activity in the cerebellum, superior temporal gyrus, anterior insula, MCC,

vlPFC and dlPFC. Furthermore, they reported that BOLD activity in the posterior parietal cortex,

anterior insula, MCC, cerebellum, and caudate nucleus were correlated to pain intensity ratings.

Another study of the neural correlates of acute orofacial pain reported that acute electrical

stimulation of the tooth activated nuclei believed to be related to trigeminal nociceptive

processing (within the VBSNC) and brainstem nuclei implicated in pain modulation, such as the

RVM [926]. Furthermore, they found significant clusters of activation in S1, S2, anterior insula,

MCC, PMC, M1, SMA and dlPFC. Therefore, it seems that acute pain in the orofacial region

activates the same set of brain regions as acute experimental pain outside the orofacial region.

The putative roles of these regions in pain-related processing as determined by neuroimaging

studies are described below.

S1

It is well known from classic brain stimulation studies in humans and animal electrophysiology

studies that S1 is somatotopically organized (see Section 2.3.5). This S1 homunculus has been

largely confirmed in human brain imaging studies (for a review, see: [23]), generally showing an

inverted organization with the face represented ventrolaterally, the upper limb more dorsolateral,

and the lower limb represented within the medial wall. Of relevance to this thesis, several studies

have demonstrated that the dermatomes of the three branches of the CNV are somatotopically

represented on the ventrolateral surface of S1: V1 is more superior and V3 is more inferior,

although there is some overlap [205,446,630]. Of these studies, one has explicitly tested the

somatotopy of the branches of the CNV with a noxious stimulus, compared to an innocuous

stimulus. Specifically, DaSilva and colleagues [205] found that the representation of noxious and

innocuous stimuli in the receptive fields of V1 and V2 were the same, whereas noxious stimuli in

the receptive field of V3 seemed to have a broader representation than innocuous stimuli in the

same receptive field. This broader cortical representation of V3 overlapped with the

representation of V1 and V2.

Similar to the electrophysiological data presented in Section 2.3.5, neuroimaging data has not

provided a clear role for S1 in pain processing [130]. It is unclear whether S1 is implicated in

pain perception. It has been proposed that the activation of S1 is due to the tactile component of

noxious stimuli, e.g., when a heat probe is used to stimulate the arm, the probe places some

94

pressure on the arm, and fibres that encode innocuous warm temperatures or touch may also be

firing. Furthermore, imaging studies have found vastly different findings in S1 activation in pain

paradigms. Specifically, Talbot et al. [846] reported significant contralateral S1 activation to the

stimulation of the arm, whereas Jones et al. [464] did not find S1 activation. In a later study,

Apkarian et al. [28] found that S1 actually showed deactivation in response to noxious thermal

stimuli. However, in a meta-analysis, Apkarian and colleagues [23] reported that 69% of PET,

76% of fMRI, 10% of EEG and 70% of MEG studies reported an S1 activation in acute pain

paradigms. More recently, MEG [699,855], EEG [885] and diffuse optical tomography [68]

studies have reported S1 responses evoked by painful stimuli. Also, a meta-analysis carried out

by Duerden and Albanese [274] identified S1 as a region that is activated across 140

neuroimaging studies of acute experimental heat and cold pain. Several imaging studies in

humans have found that S1 (in addition to other cortical areas) shows graded activation in

response to the increased intensity of experimental pain stimuli [169,240,629,705]. Also,

Rainville and colleagues [715] reported that when subjects’ perceived pain intensity and

unpleasantness were manipulated under hypnosis, S1 and possibly S2 responses correlated with

perceived pain intensity, whereas the ACC response correlated with the perceived pain

unpleasantness. In sum, neuroimaging studies have provided inconsistent findings with regards

to the role of S1 in nociception and pain processing. However, more recent and sensitive

methods have clearly demonstrated that S1 may, indeed, be implicated in pain perception.

Therefore, neuroimaging studies have replicated many of the previous findings from

electrophysiological and brain stimulation of the role of S1 in nociception and pain perception.

Specifically, neuroimaging studies have confirmed the somatotopic organization of S1, and have

shown that the haemodynamic response in S1 is graded to noxious stimulus intensity. Together,

these findings suggest that S1 is implicated in the sensory-discriminative dimension of pain,

encoding the location and intensity of a noxious stimulus.

S2/Parietal operculum

Although many imaging studies report that S2 is activated during painful stimuli, the findings

have not clarified the role of S2 in pain processing. Apkarian et al. [23] reported that 68% of

PET, 81% of fMRI, 60% of EEG and 95% of MEG studies reported S2 activation in studies

investigating the neural correlates of acute pain in healthy controls. Also, Duerden and

Albanese’s meta-analysis [274] identified S2 as one of the regions that is most commonly

95

activated in neuroimaging studies of experimental pain. However, it is not clear how S2

activations reflect various aspects of the pain experience. It is also unclear whether S2 receives

nociceptive information from the thalamus via a relay through S1 (i.e., serial processing)

[16,407], or directly in parallel to information being sent to S1 [699,702]. Furthermore, evidence

from EEG studies of laser-evoked potentials (LEP) have demonstrated that the early

electrophysiological nociceptive response (N2) is caused by activity in S2/posterior insular

cortex (e.g., [335]), which suggests that there is parallel processing in S1 and S2, although recent

findings have opposed a parallel processing model with regard to the sensory-discriminative

dimension of pain [885].

One study investigating painful LEP with subdural electrodes has demonstrated somatotopic

organization along the anterior-posterior axis, with the foot anterior to the face, within S2 [897].

This organization is different from that of tactile input, and it has been suggested that there may

be a pain somatotopic map independent of the tactile somatotopic map [23].

Insula

The insula also has several somatotopic maps for different noxious stimuli, and within different

subregions [64]. In this somatotopic map, the face is anterior to the foot. The insula was activated

in many neuroimaging studies [187], that investigated many paradigms, including gustation,

emotion, olfaction, empathy, interoception, motor output, attention, language, memory, speech,

and pain [506]. Apkarian et al. [23] reported that 88% of PET, 100% of fMRI, 30% of EEG and

20% of MEG studies reported insular activation in imaging studies of acute pain in healthy

volunteers. Duerden and Albanese’s [274] meta-analysis of 140 experimental pain neuroimaging

studies also identified the insula as one of the most likely activated regions. In a meta-analysis of

the function of the insula, Kurth and colleagues [506] investigated 79 studies that showed insular

activation in imaging studies of acute pain stimulation paradigms in healthy controls, and found

that all of the subregions of the insula were activated. It is, however, noteworthy that this same

study identified that different subregions of the insula showed responses to paradigms that tested

the neural correlates of processes related to pain, such as somatosensation, motor output,

attention, interoception, and emotion, and that these activations generally overlapped with that of

pain.

96

A number of groups have postulated a role for the insula in terms of pain perception. For

instance, evidence that the insula is involved in intensity encoding comes from a study by

Coghill and colleagues [169] who reported that the insula responds in a graded fashion to

increasing intensity of noxious stimuli. Brooks and Tracey [116] suggested that the insula is a

multidimensional integration site for pain. An alternative hypothesis about the role of the insula

proposed by Craig [183] suggested that the mid-posterior insula was a multimodal homeostatic

or interoceptive area. The former hypothesis maintained pain as a separate modality, whereas

Craig’s hypothesis established pain as a subset of homeostatic function. Baliki and colleagues

[48] have posited another hypothesis suggesting that, in addition to integrating the dimensions of

pain perception, the insula is a central, multimodal magnitude estimator and a nociceptive-

specific magnitude estimator. However, the insula has been implicated in a multimodal sensory

stimulus salience detection network, where salience stimuli are encoded and processed to

appropriately orient attention [263-265,445,530,632,633].

In sum, because the insula receives multimodal input, it is an ideal site for integrating

information from the various dimensions of pain – it is the site where nociception becomes pain.

However, it is activated in many other paradigms, and across different modalities [265,956].

Therefore, the insula may best be considered as a region that encodes behaviourally-salient

stimuli, including pain.

Cingulate cortex

The cingulate cortex is a large, heterogeneous brain region that can be subdivided into several

subregions (see Section 2.3.5) [169]. Apkarian et al. [23] reported that 94% of PET, 81% of

fMRI, 100% of EEG and 25% of MEG studies reported cingulate activation in imaging studies

of acute pain in healthy volunteers. Furthermore, Coghill and colleagues demonstrated that the

cingulate, specifically the MCC, showed graded responses to increasing intensity of noxious

stimuli [169]. The different subregions of the cingulate cortex have been implicated in different

dimensions of pain, including the affective, cognitive, modulatory and motor dimensions

[128,860,902,903]. The aMCC/pgACC is a complex, multimodal region that has been implicated

in a number of functions [69]. For instance, this region has been identified as a node in the

salience network, and has been implicated in multimodal salience detection

[213,229,262,263,774,849,927], including detection of noxious stimuli

97

[220,230,231,259,265,440,508,528,632,933]. The cingulate has also been implicated in the

cognitive and affective processing of pain [218,230,529,716,776,777,934,936], and the MCC,

which includes the CMA, is involved in action selection and modulation of motor output in

response to aversive stimuli [790,898,904]. The sgACC has been implicated in pain modulation

[95]. Furthermore, the pgACC and sgACC have been implicated in emotional processing and

depression [299,579,790] (See section 2.3.6).

Prefrontal cortex

The PFC is a large, heterogeneous region, in the frontal lobe anterior to the motor cortex. There

is no evidence suggesting that the PFC receives nociceptive information directly from the

thalamus (see Section 2.3.5). The PFC does, however, receive input from a number of

nociresponsive brain regions, such as the cingulate cortex, the insula, and the somatosensory

cortex [644]. Apkarian et al. [23] reported that 39% of PET, 70% of fMRI studies reported PFC

activation in imaging studies of acute pain in healthy volunteers. Interestingly, no EEG or MEG

studies reported PFC activation. The PFC has been implicated in the cognitive-motivational

dimension of pain. Evidence for this role comes from a number of studies that have investigated

the interaction of pain and cognition of the brain. For instance, the OFC is activated when

subjects are distracted from pain, which suggests that this region is involved in pain modulation

[888]. Furthermore, some studies have found that the dlPFC is implicated in pain modulation

[552], although other studies have not found this. Conversely, neuroimaging studies of persistent

pain, and experimental models of hyperalgesia and allodynia have shown that increased

sensitization is related to increased activity in the dlPFC [56,443,551,775]. Furthermore, one

study has demonstrated that regulation strategies can modulate pain via the vlPFC, which is

involved in top-down cognitive modulation of pain [932]. However, another regulation study did

not find that the vlPFC was implicated in pain modulation [753]. Alternatively, it has been

suggested that the vlPFC is implicated in pain anticipation, rather than modulation [752,908].

The medial PFC has been implicated in the integration of cognitive and motivational dimensions

of pain [935]. Therefore, although the PFC does not receive nociceptive input from the thalamus,

neuroimaging has revealed that the PFC plays an extensive role in the cognitive dimension of

pain, and, through it connections to antinociceptive brainstem regions (see Section 2.3.6), pain

modulation.

98

Cortical and subcortical motor regions

Motor regions are activated in acute pain paradigms in healthy controls, but less reliably than the

aforementioned brain regions [23]. As previously discussed (Section 2.3.5), in the context of

pain, it is believed that motor regions serve two purposes: to orient the body toward the source of

pain and to initiate nocifensive behaviour (e.g., avoid the stimulus). Similarly, another study

reported that hypertonic saline injection into the masseter muscle, an experimental acute pain

model, was related to deactivation in the face region of M1, as measured by fMRI [638]. It is,

however, noteworthy that some studies have not observed activation in M1 during noxious

stimulation [395,744].

The basal ganglia are a set of subcortical nuclei that have been associated with motor function.

However, it is noteworthy that the basal ganglia do receive input from many cortical and

subcortical regions, and form several functional loops [644]. It has been proposed that this

extensive wiring to the brain suggests that the basal ganglia may be implicated in more than just

motor functions [388]. For instance, the basal ganglia are often activated in neuroimaging studies

of experimental pain [107]. It has been suggested that the cortico-thalamo-basal ganglia-cortical

loops may provide a unique anatomical substrate for the integration of the various dimensions of

pain [107]. However, more work is required to further investigate this possibility.

Other motor regions are also commonly activated in brain imaging studies of experimental pain

in healthy volunteers, including the PMC and SMA [23,274]. The PMC has been implicated in

cognitive modulation of motor output, and the SMA has been implicated in motor planning, and

the temporal organization of movements [692]. Therefore, their activation during experimental

pain is likely related to the cognitive and motivational dimensions of pain, with regard to

nocifensive behaviours.

2.5.1.2. Imaging salience and pain

Salience and pain are closely related. Pain is inherently salient and therefore can engage brain

regions that serve as “salience-detectors”. Specifically, salient stimuli have been shown to

activate a right-lateralized brain network, including regions in the frontal, parietal, insular and

cingulate cortex [364,604-606,647]. In an fMRI study, Downar et al. [262-264] identified a

multimodal salience network in humans. In the context of pain, a network of regions, comprised

99

of the MCC, temporoparietal junction, PFC, and the putamen encode stimulus salience.

Similarly, an fMRI study investigating the neural basis of perceptual decision-making in the

context of pain as a perceived threat and increased salience stimulus, reported that the insula

integrated the perceived threat to form a decision about the stimulus [933]. It is thus conceivable

that these regions form a system that appraises, encodes and modulates a salient (and potentially

threatening) stimulus based on the context in which it is presented – or its relative salience.

Another study demonstrated that brain regions that are commonly activated by noxious stimuli

are also activated by equally salient, non-noxious stimuli [530]. Therefore, results from

experimental pain paradigms attempting to investigate the roles of brain regions in pain must be

interpreted with caution, as the aforementioned studies have demonstrated that they are non-

specific to pain, and may be implicated in encoding any behaviourally-relevant stimulus.

2.5.1.3. Chronic pain

Many studies have used functional neuroimaging to investigate the central abnormalities in

chronic pain patients. One of the differences between BOLD-fMRI and PET is that BOLD-fMRI

requires a stimulus, as it relies on differences in activation, whereas PET can assess baseline

function. In terms of imaging chronic pain disorders, fMRI cannot capture brain activity related

to spontaneous pain, or other clinical features. PET can capture these features, but is invasive as

it requires a radioactive tracer. Therefore, some groups use percept-related fMRI, where subjects

are continuously rating a quality of a painful stimulus throughout an fMRI scan [213,226,509]. In

a meta-analysis of all functional neuroimaging studies of pain up to 2005, Apkarian and

colleagues [23] compared noxious stimulus-evoked brain activation in healthy, pain-free subjects

and chronic pain patients. They reported that, in general, chronic pain patients showed less

noxious stimulus-activation of S1, S2, insula, and the ACC/MCC compared to pain-free subjects.

They reported no difference between the groups for the thalamus (although studies have reported

thalamic findings), and that the PFC showed significantly greater activation in chronic pain

patients, compared to normal subjects. However, not all chronic pain disorders have the same

symptomatology. Therefore, it is more likely that different types of chronic pains have different

“brain signatures” or “fingerprints”– the pattern of abnormally functioning brain regions is

specific to the symptom profile of the pain disorders. A challenge for neuroimaging of pain is to

disentangle which abnormally-functioning brain region is specific to a pain disorder, and which

is common to all pain disorders.

100

Another method of analyzing functional brain imaging data is to identify regions that may be

functionally connected [333]. Conceptually, this method is based on the idea that brain regions

do not process information in isolation, but rather process information in different schemes,

either hierarchical, or in parallel. Therefore, these regions are functionally connected during a

task, and in task-free paradigms (or at rest) [213]. Recent evidence has shown that some of these

functionally connected brain regions may be disrupted in chronic pain (e.g., [49,560]).

Few studies have investigated functional brain abnormalities in TMD. The first study to do so,

investigated neural responses to tactile-flutter stimuli on the index finger [641] and found that the

patients reported the stimuli as more intense, but not painful, than the control subjects. Further,

compared to the control subjects, the tactile stimuli evoked less activation in S1, S2 and insula,

and greater activation in the thalamus, MCC, the amygdala, and another region of S1 in the

patients with TMD. This study reported that patients with TMD have both increased and

decreased activation across several regions, including S1, S2 and the insula to a vibratory

stimulus applied outside the orofacial region, compared to controls. Patients with TMD show

increased activation in the MCC, in the contralateral insula and thalamus. Therefore, patients

with TMD show abnormal processing of an innocuous tactile stimulus outside the orofacial

region. It would have been more informative to have also applied a vibratory stimulus to the

orofacial region, to test whether tactile thresholds were different in this region, and the neural

correlates of these differences. However, a vibratory stimulus on the face in an MRI magnet

would likely be a source of noise and therefore difficult to test.

Our laboratory has conducted an fMRI study to investigate brain responses to cognitive and

emotional attention-demanding tasks in patients with TMD [928]. In this study, a group of

patients with idiopathic TMD and a group of age- and sex-matched pain-free controls underwent

fMRI scanning while performing three counting Stroop tasks (neutral, conflict numbers,

emotional). We found that behaviourally, the patients with TMD had sluggish reaction times,

compared to controls, in all three Stroop tasks. The fMRI data indicated that patients had greater

deactivation in the mPFC, OFC, medial temporal gyrus and dlPFC during the neutral Stroop task,

compared to controls. Furthermore, patients showed increased activation in the dlPFC, vlPFC,

insula, ACC, SMA and amygdala during the high conflict Stroop task, compared to controls.

During the emotional Stroop task, patients showed greater activation than controls in the sgACC,

MCC, SMA, PCC, insula, vlPFC, and the inferior parietal lobule. Finally, patients showed less

101

functional connectivity than controls between the dlPFC and the aMCC, and the pgACC and the

amygdala during the high conflict task. During the emotional Stroop task, patients showed

increased connectivity between the dlPFC and the MCC. In sum, these findings suggest that

patients with idiopathic TMD have cognitive deficits related to poorer task performance,

compared to controls. These deficits may be related to abnormal brain function in

cognitive/affective and attention-related brain regions.

Resting state networks reflect the intrinsic connectivity of brain regions when subjects are not

actively engaged in a task. It has been posited that these networks reflect the anatomical

connectivity, as well as the “history” of prior co-activation during task [235,236]. In the context

of TMD, a recent study investigated the connectivity of the insular and cingulate cortices in

myofascial patients with TMD during rest, and during experimental pressure pain in the orofacial

region [448]. These studies reported that during resting state, in patients with TMD compared to

controls, the left anterior insula was more functionally connected to the right pgACC and the left

parahippocampal gyrus, and the right anterior insula was more connected to the right thalamus.

During noxious stimulation, both groups showed greater functional connectivity between the left

anterior insula and the left S2, cerebellum, the left posterior insula showed greater connectivity

to the right dlPFC. Finally, the study reported a group-by-pressure interaction in the functional

connectivity between the left anterior insula and the left pgACC, bilateral frontal polar cortex,

and the right anterior insula with the right pgACC. Therefore, these findings suggest that in TMD

there is increased functional connectivity between sensory and pain-modulatory regions at rest,

and increased connectivity between cognitive and modulatory regions during experimental

pressure pain.

In sum, few studies have used functional neuroimaging in TMD. These studies have found

abnormal function in sensory, motor, cognitive/modulatory and salience-related brain regions.

2.5.2. Structural imaging of pain

2.5.2.1. Acute Pain

To date, only two studies have investigated structural gray matter correlates of acute pain.

Teutsch and colleagues [852] demonstrated that eight days of noxious heat pain stimulation

induced increased gray matter volume in MCC and S1. Furthermore, as subjects habituated to the

102

noxious stimulus over eight days , these gray matter increases resolved (i.e., the gray matter

volume returned to baseline). In a study by Erpelding et al. [298], cortical thickness in S1

correlated with individual subjects’ heat and cold pain sensitivity, and cortical thickness of the

MCC correlated with heat pain sensitivity. To date, the relationship between white matter

microstructure and acute pain thresholds or tolerance has not been studied in healthy subjects.

2.5.2.2. Chronic Pain

Gray matter abnormalities

There are several approaches to investigating gray matter structure in chronic pain patients. For

instance, studies can either investigate the whole brain, or focal areas (ROIs) that have been

hypothesized to show abnormal gray matter structure, or investigate a specific ROI. Furthermore,

the question posed can also vary: (1) do the patients have abnormal structure, compared to a

control group?; and (2) Are there structural neural correlates of behavioural measures, such as

pain characteristics and personality traits? Furthermore, as discussed in Section 2.4.2, there are

various MRI-based analysis methods to measure gray matter. It is noteworthy that studies that

only used CTA to evaluate differences in gray matter structure did not study subcortical

structures.

A summary of all studies examining gray matter structure in chronic pain populations is provided

in Table 2.3. It includes chronic back pain [27,121,764,782], FMS

[125,502,558,712,738,766,949], IBS [98,225,779], migraine [207,490,740,762,889], trigeminal

neuropathic pain [206,382], trigeminal neuralgia [382], TMD [358,963], phantom-limb pain

[268], provoked vestibulodynia [772], complex regional pain syndrome (CRPS) [356], cluster

headache [578], chronic tension-type headache [765], medication-overuse headache [765],

hypnic headache [434], chronic pelvic pain [303], chronic posttraumatic headache [650],

persistent idiopathic facial pain (previously atypical facial pain) [763], hip osteorarthritis [741],

rheumatoid arthritis [919] and pain disorder [887] (a diagnosis based on the DSM-IV, defined as

persistent and chronic pain at one or more sites that cannot be fully explained by a physiological

process or physical disorder). Overall, the most common brain regions showing gray matter

abnormalities were the PFC, insula, the ACC and the MCC. Other regions that were abnormal in

chronic pain included the thalamus, basal ganglia, S1, S2, and the brainstem. Additionally, some

studies demonstrated abnormalities in the temporal lobe and the PCC. In every region, with the

103

notable exception of the basal ganglia and the brainstem, most studies showed a decrease in gray

matter in chronic pain.

Of particular interest to this thesis, three studies used VBM to investigate gray matter

abnormalities in TMD. One study by Gustin and colleagues [382] reported no gray matter

abnormalities in the brain of patients with TMD. Gustin and colleagues did not provide the

diagnostic criteria used to select patients with TMD. Another study, by Younger and colleagues

[963], tested for gray matter abnormalities in myofascial patients with TMD and reported four

main findings: 1) compared to controls, patients with TMD had increased gray matter in the

thalamus, the basal ganglia, insula, the vlPFC, and the brainstem; 2) a negative correlation

between jaw pain and gray matter volume in the dorsal posterior cingulate, the MCC,

SMA/preSMA, and the superior temporal gyrus; 3) a negative correlation between experimental

pressure pain thresholds (as measured by pressure algometry) and brainstem gray matter volume;

and 4) a positive correlation between TMD duration and gray matter volume in the posterior

cingulate, the hippocampus, and the brainstem. Another study by Gerstner and colleagues [358]

investigated myofascial patients with TMD and found that patients with TMD had decreased

gray matter in the vlPFC, MCC, PCC, SMA, insula and the regions within temporal lobe. These

findings seem to conflict with the Younger et al. [962] study, but could be attributed to

methodological differences between the studies. For instance, the age range and the range of

durations of TMD in the Younger patient sample were larger than those of the Gerstner patient

sample. These factors could contribute to changes in gray matter, and they cannot be adequately

controlled with the statistical methods used to evaluate group difference in structural gray matter

analyses. For a complete discussion of these points see: [615].

Interestingly, recent evidence suggests that some gray matter abnormalities in some chronic pain

disorders are caused by the pain, rather than pre-existing abnormalities. For instance, Rodriguez-

Raecke and colleagues [741] showed gray matter abnormalities in several brain region of patients

with primary hip osteoarthritis. These patients underwent surgery, and for the most part, the pain

was resolved. The post-operative scan of a subgroup of these patients revealed that some, but not

all, of the gray matter abnormalities had resolved (i.e., there were no differences between

controls and patients). Four other studies have since demonstrated similar effects – partial

reversal of gray matter changes in brain once patients’ pain has been resolved, in the same [384]

and other chronic pain disorders, including chronic low back pain [782]. However, it remains

104

unclear whether these changes in gray matter are due to the resolution of pain, or are secondary

to the resolution of pain. For instance, it is possible that a person with chronic pain in the hip will

be more sedentary, compared to when they are no longer in pain. Furthermore, pain has been

associated with changes in mood, including increased incidence of depression and anxiety, which

could resolve once the pain is gone. Furthermore, it has recently been shown that medications

can alter the structure of the brain [916,962]. Therefore, when patients are no longer taking pain-

killers, the changes in the brain may resolve.

Furthermore, it is noteworthy that not all of the observed gray matter abnormalities were

abolished with the cessation of pain. This may be related to lasting effects of medications, such

as opiates, which have been shown to produce long-term changes in the pgACC, MCC,

amygdala, the vlPFC, hypothalamus, brainstem, and the hippocampus of patients with chronic

low-back pain [962]. These changes did not show reversal after the patients had ceased managing

their pain with morphine for an average period of 4.7 months. Therefore, there may still be pre-

existing abnormalities that contribute to the development of chronic pain.

Additionally, it is noteworthy that functional pain syndromes, such as IBS, FMS and TMD, are

defined as disorders that disturb normal function without any obvious structural or biochemical

abnormality [246]. Therefore, it is possible that mechanisms other than increased nociceptive

drive from the periphery may be driving gray matter abnormalities in the brain. An alternative

possibility is that gray matter abnormalities pre-date the onset of chronic pain. In this scenario,

gray matter abnormalities are pre-existing vulnerabilities that may contribute to the development

of chronic pain [213]. Evidence for this proposition comes from studies that have identified

regions of gray matter abnormalities related to stable personality traits, such as neuroticism or

pain catastrophizing [98,772], which are related to heightened pain sensitivity (see Section

2.1.7).

In sum, there are both pre-existing and pain-driven gray matter abnormalities in the brain of

chronic pain patients. In general, there appears to be increases in sensory/nociceptive brain

regions, indicative of increased nociceptive drive; decreases in pain modulation brain regions,

indicative of potentially dysfunctional pain-modulatory systems.

105

White matter abnormalities

There have only been eight studies investigating white matter abnormalities in the brain of

chronic pain patients. Of these, two have used VBM to evaluate white matter volume from T1-

weighted images, and four have used DTI to delineate abnormalities in white matter

microstructure, or gray matter microstructure. The two studies that have used DTI to investigate

gray matter [334,558] and the white matter VBM studies [358,553] are omitted from this thesis

because of flaws in the experimental design and methodologies. Of the four remaining studies

that used the approach of DTI, one did not find any group differences between patients with

chronic pelvic pain and controls [303]. The other three that did identify abnormalities are

described below.

The first study to test for abnormalities in white matter microstructure associated with chronic

pain, by DaSilva and colleagues [208], reported decreased FA along tracts between the brainstem

and the thalamus and the thalamus and S1 cortex of patients with migraine. The authors

concluded that there are abnormalities along the ascending nociceptive pathways in migraine

patients.

Geha and colleagues [356] reported white matter abnormalities in patients with CRPS that

consisted of lower FA in the cingulum bundle and the adjacent corpus callosum. This was the

first study to investigate the abnormalities in white matter tracts by elucidating the connectivity

of the region. They reported that abnormal white matter region had fewer connections per unit of

distance in CRPS patients compared to controls.

A study by Chen et al. [156], from our laboratory, investigated white matter abnormalities in

patients with IBS, and reported that patients had increased FA in the fornix and the

external/extreme capsule, adjacent to the insula. It was also found that FA in the anterior insula

and the VPL thalamus were significantly correlated to pain severity. Further, FA in the left insula

correlated with pain unpleasantness and duration. Finally, in patients with IBS, FA in the

cingulum bundle was negatively correlated to the PCS, and a positive correlation between FA in

the MD thalamus and neuroticism. These findings suggest that IBS patients have abnormal

connections in white matter tracts related to nociception and/or cognitive/limbic function.

106

In addition to the aforementioned studies that have investigated white matter abnormalities in the

brains of chronic pain patients, two DTI studies have investigated abnormalities along the CNV

in patients with trigeminal neuralgia. One study by Fujiwara et al. [339] did not find any

significant abnormalities in the FA, MD, or the cross-sectional area of the patients’ CNV. This

study reported that no significant group differences in the ratio of FA of the affected to the

unaffected nerve. However, the study reported significantly more variance in the FA values of

patients’ CNV, compared to controls, and this value was positively correlated to the

affected:unaffected ratio of the cross-sectional area of the CNV. The other study, by Leal et al.

[525], reported that the affected CNV had significantly increased MD, and decreased FA, nerve

volume, and cross-sectional area. Furthermore, the decrease in FA was positively correlated to

the decrease in volume and cross-sectional area of the CNV. Decreased FA and increased MD

indicate that there is increased diffusion, or less organization. Therefore, it is possible that the

CNV has a larger diameter, is inflamed, or damaged. However, the interpretation is limited as

other measures of white matter integrity (RD and λ1) were not provided. These studies indicate

that there are structural abnormalities along the CNV in patients with trigeminal neuralgia, and

suggest that DTI can be used to investigate the CNV for abnormalities in TMD.

107

Table 2-3. Summary of regions with gray matter abnormalities in chronic pain

Chronic pain disorder Method n patients

(# females)n controls

(# females) Thal BG S1 S2 IC ACC MCC M1 PFC Bstem ThresholdMultiple

comparisons correction

Contributing factors Ref

VBM 26 (16) 26 (16) ! ! dl p<0.05 Permutation D [27]VBM 18 (9) 18 (9) " " Put ! ! dl ! pTL,

TL p<0.001 Nonea D, P [764]VBM 8 (3) 8 (4) ! PPC p<0.001 None [121]VBM 18 (10) 16 (8) ! ! !

! vl, dl, OFC " MTL p<0.05 Cluster [782]

VBM 10 (10) 10 (10) ! ! ! m ! PCC, PHG p<0.05 RFT; Cluster t

>3 A, D [502]

VBM 20 (19) 22 (20) ! " " OFC ! STG, " Cb

p<0.05, 0.001 ROI; Unc P [766]

VBM 30 (30) 30 (30) ! HC p<0.05 Permutation, Bonferroni [558]

VBM 30 (30) 20 (20) !! PHG, PCC p<0.0005 None [949]

VBM 14 (14) 14 (14) ! ! vl ! Amyg p<0.001 None De, An, M [125]VBM 5(5) 5(5) ! SMA p<0.05 Permutation [712]VBM 14 (14) 11 (11) ! ! ! PCC p<0.02 FDR [738]CTA VBM 9 (6) 11 (7) ! ! ! ! p<0.001 None [225]CTA VBM 11 (11) 16 (16) ! " HT p<0.05 FDR,

Permutation PCS, D [98]

CTA VBM 56 (56) 49 (49) ! ! vStr, Put " " "

" OFC, dl; ! vl, FP, m

" HC p<0.05, p<0.01

Cluster, None A, An, De [779]

VBM 16 (15) 15 (13) ! ! dl, vl " PAG ! TL p<0.001 None A, D [740]VBM 35 (32) 31 (31) ! ! ! p<0.05 SVC [762]

27 (21) 27 (20) ! ! vl ! TL p<0.05, 0.001

FWE, SVC

11 (9)/16 (12)b ! ! ! ! vl, dl ! Amyg p<0.05 SVC P12 (9) [w/ aura] " p<0.05 ROI12 (7) [w/o aura] " " p<0.05 ROI

VBM 20 (17) 33 (29) ! ! !! dm, dl,

OFC

! PPC, PMC, V1

p<0.05, 0.001

SVC, None D, P [490]

[889]

[207]

Other?

CBP

FMS

IBS

Migraine

CTA 12 (9)

VBM

108

Chronic pain disorder Method n patients

(# females)n controls

(# females) Thal BG S1 S2 IC ACC MCC M1 PFC Bstem ThresholdMultiple

comparisons correction

Contributing factors Ref

TNPCTA 6 (5) 6 (5) !/" !/" " ! !/" " FP, vl,

dl "PCC p<0.05 None D [208]TN 21 (17) 30 (24) " " vStr, Put " !/" p<0.01 FDR

20 (16) 31 (26) p<0.001 NoneVBM 15 (15) 15 (15) ! ! GP, Put " ! ! vl ! p<0.05 FDR P, D [963]

VBM 9 (9) 9 (9) " " " vl " PCC, TL, PHG

p<0.05, 0.001

Permutation, None [358]

Phantom-limb VBM 28 (8) 28 (8) "VPL p<0.05 FWE D, P [268]PVD VBM 14 (14) 14 (14) ! GP, SN, NC " HC,

PHG p<0.05 Cluster P, PCS [772]CRPS VBM 22 (19) 22 (19) " NAcc " " " vm p<0.05 Permutation D, P [356]CH VBM 25 (2) 29 (0) ! HT p<0.001 Bootstrap [578]CTTH 20 (10) " " " " OFC "

" PCC, PHG, p<0.05 SVC D

MOH 20 (18) " OFC " p<0.0005 NoneHypH

VBM 14 (9) 14 (9) " " dl" HT, PCC, TL, Cb

p<0.001 None [434]

CPP VBM 19 (0) 16 (0) Not provided D, P [303]32 (20) 30 (17) p<0.05 FWE

12 (7)/20 (13)c " " dl p<0.05 FWE

32 (20)d ! ! p<0.05 FWEPIFP

VBM 11 (9) 11 (9) " " " " " dl, vm p<0.001 None [763]

32 (19) 32 (19) " " " "

" vm, OFC, vl,

!/" dle"

" Cb, Amyg p<0.001 None

10 (6)f ! ! ! ! dl ! ! Amyg p<0.05 None

[765]

[382]

[741]

[650]

VBM

Other?

VBM 40 (?)

Hip OA

VBM

CPTH

VBM

TMD

109

a. the p-value was reduced for some findings, but the lower threshold is not provided b. Between patient groups analysis: Chronic vs. Episodic c. Comparing temporary post-traumatic headache to chronic post-traumatic headache (temporary > chronic) d. Longitudinal study comparing time 1 scans with subjects after the pain has resolved e. Decrease GM on the right, increase GM on the left f. Longitudinal study comparing time 1 scans with post-operative scans, when pain is resolved g. Pain disorder is a diagnosis based on the DSM-IV, defined as persistent and chronic pain at one or more sites that cannot be fully explained by a physiological process or physical disorder."

Abbreviations: Amyg — amygdala; ACC — anterior cingulate cortex; BG — basal ganglia; Cb — cerebellum; CH — cluster

headache; CPP — chronic pelvic pain; CPTH — chronic posttraumatic headache; CRPS — complex regional pain syndrome; CTA —

cortical thickness analysis; DBM — deformation-based morphometry; dlPFC — dorsolateral prefrontal cortex; dmPFC —

dorsomedial prefrontal cortex; FDR — false discovery rate; FWE — family-wise error; GMV — gray matter volume; GP — globus

pallidus; HC — hippocampus; Hip OA — hip osteoarthritis; HT — hypothalamus; HypH — hypnic headache; IBS — irritable bowel

syndrome; IC — insular cortex; iTL- inferior temporal lobe; M1 — primary motor cortex; MCC — midcingulate cortex; MOH —

medication-overuse headache; mPFC — medial prefrontal cortex; MTL — medial temporal lobe; NC — caudate nucleus; OFC —

orbitofrontal cortex; PCC — posterior cingulate cortex; PFC — prefrontal cortex; PHG — parahippocampal gyrus; PIFP — persistent

idiopathic facial pain; PMC — premotor cortex; PPC — posterior parietal cortex; pTL — posterior temporal lobe; Put — putamen;

Chronic pain disorder Method n patients

(# females)n controls

(# females) Thal BG S1 S2 IC ACC MCC M1 PFC Bstem ThresholdMultiple

comparisons correction

Contributing factors Ref

Rheum Arth VBM CTA DBM

31 (22) 25 (17) ! Put, NAcc, NC p<0.05 FDR D [919]

Pain

disorderg VBM 14 (14) 25 (25) " "" vm, OFC

" iTL, PHG p<0.05 FDR D [887]

Other?

110

PVD — provoked vestibulodynia; RFT — random field theory; Rheum Arth — rheumatoid arthritis; ROI — region of interest; S1 —

primary somatosensory cortex; S2 — secondary somatosensory cortex; SMA — supplementary motor area; SN — substantia nigra;

STG — superior temporal gyrus; SVC — small-volume correction; Thal — thalamus; TMD — temporomandibular disorder; TN —

trigeminal neuralgia; TNP — trigeminal neuropathic pain; TL — temporal lobe; V1 — primary visual cortex; vlPFC — ventrolateral

prefrontal cortex; vmPFC — ventromedial prefrontal cortex; vSTR — ventral striatum

111

2.6. Neuroplasticity

It is clear that brain structure is defined by genetics and experience [577]. In fact, it has been

suggested that neuroplasticity is “ the normal ongoing state of the nervous system throughout the

life span” [672], in order to better adapt to the environment [969]. This concept of brain

neuroplasticity was recognized early in the twentieth century. For instance, Cajal described the

concept:

The labor of a pianist [...] is inaccessible for the un-educated man as the acquisition of new skill requires many years of mental and physical practice. In order to fully understand this complex phenomenon it becomes necessary to admit, in addition to the reinforcement of pre-established organic pathways, the formation of new pathways through ramification and progressive growth of the dendritic arborization and the nervous terminals. [134] (cited and translated in [672])

The term plasticity was first used by Ernesto Lugaro in 1906 [554] to describe changes in the

nervous system during maturation, compensatory changes after lesions, and changes related to

learning and memory. He suggested the specific chemicals, today known as neurotrophins, that

are responsible for the development and organization of the nervous system, continue to

reorganize the neuronal connections in the brain. This property of the nervous system, which he

called plastic, occurred through chemical signaling. Lugaro recognized two processes specific to

neurones: conduction of an impulse, and transduction of a neural signal between neurones [86].

In line with Hebbian doctrine [419], Lugaro postulated that transduction would lead to

modifications of the neural circuit: increased firing would enhance a connection by modifying

the synaptic distance [86]. Despite Lugaro’s postulation that the brain remains plastic throughout

life, scientific dogma maintained that only the developing nervous system could undergo

changes. The concept of plasticity was not established until pioneering studies were published in

the second half of the twentieth century. These studies used animal models to demonstrate that

plasticity and “learning” occur at several levels of the nervous established systems, from the

periphery, and the spinal cord, to the brain (see [477]). Today, it is established that plasticity

occurs in the mature nervous system. It is also now known that plasticity can occur rapidly at

several levels of the nervous system. These neuroplastic events can either be beneficial, as in

learning a novel task, or maladaptive, as is the case in chronic pain (discussed below).

112

One of the first studies [914] to report plasticity in the somatosensory system found that after

destruction of the dorsal column neurones in the VPL region of the thalamus that normally

responded to tactile stimulation in the foot changed their receptive fields. The representation of

the arm in VPL expanded to the region that previously represented the foot. Later studies have

confirmed that plasticity occurs throughout the adult nervous system, including in the

somatosensory system (for a complete review, see: [347,475]).

2.6.1. Use-dependent neuroplasticity

One type of beneficial plasticity occurs during learning. This concept was first described

anecdotally by Hebb [419]: rats that were exposed to enriched environments performed better on

tasks requiring complex behaviour, compared to rats raised in cages. In humans, a study by

Pascual-Leone and colleagues [673] reported that learning a novel motor task is associated with

motor cortex plasticity, as measured by TMS. Interestingly, these changes returned to baseline

once the task had been learned, and became part of implicit knowledge. An fMRI study [478]

investigating motor learning (the learning of sequential finger movements) found increased brain

activation in M1 in subjects that have learned a new task, but, in contrast to Pascual-Leone et

al.’s study, these changes persisted for several months. Another example of beneficial plasticity

is a study by Nudo and colleagues [648], which showed that after an experimentally-induced

ischaemic stroke in the hand region of M1 of non-human primates, retraining of the affected

hand preserved the cortical tissue adjacent to the lesion. This preserved tissue underwent

plasticity and contributed to the functional recovery of the hand. These studies demonstrate the

concept of use-dependent plasticity. Motor learning, or retraining, can change the structure and

function of the brain – in these cases, M1.

Longitudinal studies of use-dependent plasticity have found that behavioural changes are related

to both functional and structural gray and white matter changes in the brain

[111,266,267,271,771]. Animal studies in non-human primates have demonstrated that the

nervous system undergoes functional reorganization (changes in cortical representation fields) in

response to massive changes in the environment, such as sensory loss due to deafferentation or

brain lesions [454,458,703]. Furthermore, studies in non-human primates have demonstrated

massive white matter plasticity in response to learning a novel task [422] and brain lesions [200].

113

Other studies have shown that experience-dependent (or use-dependent) plasticity is related to

cellular changes, specifically in dendritic spine density [436,493,865].

Plasticity in the nervous system can include changes to a receptive field (e.g., with increased use

of a limb, its representation in the brain can be expanded), changes in neuronal responses or

firing thresholds, and structural changes in neuronal connections.

However, not all use-dependent neuroplasticity is beneficial [576]. For instance, Rosenkranz and

colleagues [745] compared healthy musicians to a group of patients with musician’s dystonia,

and a group of controls. The authors reported that musicians showed functional plasticity in the

motor cortex. Patients with dystonia, however, had even more changes, which disrupted normal

motor output. This study demonstrates that neuroplasticity can actually interfere with normal

function. Similar mechanisms may occur in patients that develop chronic pain: increased or

prolonged nociceptive input can induce “excess” plasticity, which can itself become the

aetiology for pain (see below).

2.6.2. Neuroplasticity and pain

Normally, pain is a behaviourally-relevant signal that can promote survival by signaling injury or

potential injury. However, in some cases, pain persists long after its usefulness, even after the

injury has healed. This occurs due to a process called the chronification of pain (see Section

2.1.2). In cases of chronic pain, pain is no longer a useful protective signal which is promoting

survival, but rather the pain becomes a disease in itself. This chronification occurs due to

functional and structural changes at several levels of the nervous system, or maladaptive

plasticity.

One level of plasticity is in the periphery at the nociceptor. When injury occurs, pain experience

is heightened by increasing the sensitivity of nociceptors to thermal and mechanical stimuli (i.e.,

peripheral sensitization) [470]. This occurs, in part, because of chemical mediators that are

released at the site of injury, which modify the properties of the nociceptors such as their

threshold. There are also changes in the neurotransmitters produced by these neurones. The

dorsal horn of the spinal cord and the VBSNC also undergo plasticity [159,784,950,951]. For

instance, in phantom limb pain, where the peripheral end of the nerve undergoes changes that

cause pain, blocking conduction along the peripheral nerve at the stump does not eliminate

114

ongoing pain in all patients [96,323]. Furthermore, when a nerve is severed, there is a cascade of

events that occur in the CNS. For instance, microglia are activated, and chemical mediators

(including inflammatory molecules) are released [568]. Interestingly, when these chemical

mediators are inhibited in the spinal cord and VBSNC [159], patients with neuropathic pain

report that their pain is resolved [568]. These factors can initiate or maintain pain after peripheral

nerve injury by sensitizing or disinhibiting second-order neurones. Importantly, immune cells are

not implicated in acute pain conditions, but rather in persistent and chronic pain states (see

[159,568]). Further neuroplastic events occur in suprabulbar regions, in the thalamus and the

cortex. For instance, after peripheral nerve injury (including amputation) in non-human primates,

spinal, brainstem, thalamic and cortical neurones can undergo large-scale reorganization

[324,325,467,472-474,703]. Further evidence for pain-related neuroplasticity comes from a study

that showed that thalamic stimulation in phantom limb patients could elicit their phantom limb

pain [219]. This suggests that neuronal representation of the phantom limb remains in the

thalamus, even after the loss of the limb.

Neuroimaging studies have contributed to our understanding of brain plasticity in response to

repeated noxious stimulation. For instance, healthy subjects who received 20 minutes of noxious

stimulation over eight days habituated behaviourally (i.e., they reported the same stimulus as less

painful) [93,94,852] and showed decreased pain-evoked activation in the thalamus, S2, insula

and the putamen, and increased activation in the sgACC [94]. Structural imaging in these

subjects also revealed increased gray matter volume in the PMC, MCC, S1, inferior parietal

lobule, and the medial temporal gyrus after the eight days of stimulation, compared to baseline

[852]. Although not explicitly tested, it is inferred that these structural and functional changes

observed were related to the observed behavioural changes. Although the functional and

structural changes were located in different regions, the findings nonetheless provide evidence

that repeated noxious stimulation induces functional and structural brain plasticity. Further

support for the concept that increased nociceptive input alters the structure of the brain comes

from studies that have demonstrated that gray matter abnormalities are reversible upon the

successful treatment of the chronic pain disorder [384,741,782]. However, not all gray matter

abnormalities in chronic pain conditions are reversible. It is possible that chronic pain patients

may not have effective pain-modulatory systems. In TMD, evidence suggests that some pain-

modulatory systems, such as DNIC [113,491,756], are dysfunctional. Also, most studies

115

investigating the structure of the brain in chronic pain have reported decreased gray matter in the

pain-modulatory regions [575]. Additionally, it is possible that, to some extent, hypersensitivity

to nociceptive stimuli is related to changes that have occurred in brain structure. In line with this

concept, studies have demonstrated that some patients with TMD show allodynic and

hyperalgesic responses to experimental stimuli, consistent with the concept of central

sensitization (See section 2.2.3). It is possible that the combination of increased barrage of

nociceptive processing, an increase in the activity of pain-facilitatory pathways and decreased

activity in pain-inhibitory pathways induce central sensitization. This central sensitization can

lead to further functional and structural changes that can maintain painful states.

Based on the concept of that pain may drive CNS plasticity, structural abnormalities in chronic

pain patients would correlate with pain characteristics pain, such as duration, intensity and

unpleasantness. For instance, if we expect increased nociceptive activity, then we would expect a

positive correlation between measures of brain structure (volume or thickness in gray matter; FA

in white matter) in nociceptive brain regions and pain duration, intensity and unpleasantness.

Conversely, we would expect a negative correlation between structural measures and pain

characteristics in antinociceptive regions. Alternatively, if there are pre-existing abnormalities in

the brain that predispose subjects to the development of pain, then we would not expect

correlations between structural measures in pain-related brain regions and pain characteristics.

However, in the latter scenario, we can expect that brain structure would be related to personality

factors that predispose persons to developing chronic pain, such as neuroticism.

2.6.3. Age-related neuroplasticity

It has long been recognized that, over time, the brain undergoes atrophy. These age-related

changes have been observed in post-mortem studies [115,191,232,261,427] as well as in brain

imaging studies [322,749-751,911,968]. These studies have found that, in general, the brain

underwent atrophy with age, although some regions were spared, and others showed an increase

with age. For instance, Ziegler and colleagues [968] showed that the MCC and sgACC were

thicker in older adults than in younger adults. Some disease states have been shown to modulate

this aging process. For instance, Alzheimer’s disease has been shown to increase the rate of gray

matter atrophy. Furthermore, two studies have shown that patients with chronic pain, specifically

FMS [502] and chronic back pain [27], also had accelerated gray matter atrophy. It is not known

116

whether this form of plasticity is related to duration of pain, or other factors that may contribute

to accelerated gray matter atrophy.

2.6.4. Microstructural basis of neuroplasticity

MRI-detectable changes occur in the brain as part of the natural aging process, with increased

use (experience, learning and memory), in response to peripheral nerve injury, and amputation.

However, the neural basis of these changes is not well understood. Pakkenberg and Gundersen

[662] reported that whole brain age-related atrophy in humans is due to neuronal loss and

reduced cell packing density; whereas Peters et al. [685] reported that age-related atrophy in

macaques is related to a decrease in the number of neuropils, not neurones. There are also other

proposed hypotheses to explain mechanisms of gray matter change. For instance, rather than

neural loss, there may be glial death [575]. Recent evidence suggests that gray matter loss may

be related to the density of small dendritic spines [278,608], and the remodeling of neuronal

processes [537]. Alternatively, reversible gray matter changes in chronic pain may be caused by

neuroinflammation [238,381,920], and induce MRI-detectable increases in gray matter. This

mechanism could explain both abnormal age-related increases and maintenance of gray matter

volume/thickness. For the observed gray matter losses, however, we cannot rule out cell death as

a factor in age-related gray matter loss – healthy populations lose neurones as they age, and

persons with neurodegenerative diseases suffer increased rates of atrophy related to cell death.

Changes in measures of white matter integrity, as indicated by FA, MD, RD and λ1, each reflect

a different aspect of white matter microstructure. The most commonly computed and assessed

measure of white matter integrity is FA, because it is scaled between 0 and 1, where 0 represents

complete isotropic diffusion, and 1 represents complete anisotropic diffusion, and studies have

demonstrated that a threshold value of 0.2 for FA value can suitably delineate white matter from

gray matter [820]; however, this needs to be checked for specific populations. There could be

macrostructural factors that contribute to reduced FA, such as increased branching, crossing

fibres or larger tracts (more axons) and/or microstructural changes such as cell swelling

(œdema), changes to protein filaments (neurofilament phosphorylation), disruptions to the cell

membranes, and, to a certain extent, decreased myelin [66,67]. However, as FA is a summary

measure, it is not specific to any microstructural features of white matter tracts [695].

Histological studies have demonstrated that RD and λ1 are related to alterations in specific

117

structures [67]. FA can be increased by any decrease in RD, and increases in λ1, and thus RD and

λ1 provide greater insight into the cellular properties underlying the observed changes in FA, and

presumably, white matter tracts. Specifically, RD is related to membrane integrity, and, to some

extent, myelination; whereas λ1 is related to factors that may disrupt the axon or neurofilament

phosphorylation [67].

Recent studies have suggested that chronic pain is associated with neuroinflammation in the

brain [238,381,920], which can lead to changes in both gray matter and white matter that can be

detected by MRI. Therefore, changes in the measures of white matter microstructure may, in

fact, be related to neuroinflammation rather than increases or decreases in the number or size of

cells and/or axons present in the fields being studied. In this scenario, we would expect that

changes in FA are associated to changes in MD and RD, which are markers of inflammation

and/or œdema [63,66].

118

Chapter 3 Rationale, Specific Aims and Hypotheses

The general aim of this thesis is to determine whether patients with TMD have structural

abnormalities in the brain or CNV, compared to healthy, pain-free controls. To this end, three

MRI-based studies have been conducted to assess gray and white matter abnormalities. The

rationale, specific aims and hypotheses for each of these studies are provided below.

3.1. Study I. Contribution of chronic pain and neuroticism to abnormal forebrain gray matter in TMD patients

TMD can be idiopathic in that there may not be any clear peripheral etiological factors

identifiable [283,286,653] (see Chapter 2). In this scenario, it is thought that the CNS may

initiate and/or maintain the pain [757]. Previous structural MRI studies of chronic pain

populations have found both increases and decreases in brain gray matter (See Section 2.5.2).

There are two main routes by which the CNS may contribute to the development and/or

maintenance of chronic pains such as TMD. One possibility is that persistent or long-term

nociceptive input into the brain induces maladaptive brain plasticity that can play a role in

maintaining pain [12,950]. This maladaptive plasticity could hamper a patient’s ability to

habituate to sustained nociceptive activity, possibly related to a reduced capacity of the brain to

dampen pain by descending (top-down) controls [95]. Indeed, many structural MRI studies have

found that chronic pain-related characteristics (intensity, unpleasantness, or duration) can in part

account for gray matter abnormalities in chronic pain populations (See Section 2.5.2, and Table

2-3). Alternatively, inherent personality-related factors that reduce the brain's capacity to

modulate nociceptive activity may contribute to gray matter abnormalities, and represents a

vulnerability to develop chronic pain. For example, there is evidence that neuroticism may be

associated with pain-related suffering (See Section 2.1.7), nerve injury outcomes and neuropathic

pain [848] and inhibition of negative thoughts [178]. However, not all chronic pain patients have

high neuroticism scores, and not all persons with neuroticism have chronic pain [179]. Therefore,

neuroticism alone is not sufficient to develop chronic pain. Rather, the normal relationship

between neuroticism and brain structure and function may be disrupted within regions involved

in pain modulation, and this could facilitate or maintain chronic pain.

119

Therefore, the main aim of Study I is to examine gray matter abnormalities in patients with

idiopathic TMD and to determine the contribution of pain-related characteristics and

neuroticism.

Specific aims:

1. To determine whether patients with TMD have increased gray matter in regions of the

brain implicated in pain perception, compared to healthy, pain-free controls.

2. To determine whether patients with TMD have reduced gray matter in regions of the

brain implicated in pain modulation and motor function, compared to healthy, pain-free

controls.

3. To determine the relative contribution of TMD pain characteristics (duration, intensity

and unpleasantness) to gray matter abnormalities in the brains of patients with TMD.

4. To determine the relative contribution of neuroticism to gray matter abnormalities in the

brain of patients with TMD.

Specific hypotheses:

Compared to healthy, pain-free controls, patients with TMD will have:

1. increased gray matter in brain areas associated with pain perception;

2. decreased gray matter in areas associated with pain modulation and motor function;

3. a positive correlation between gray matter in pain perception regions and pain intensity,

unpleasantness and TMD duration; but a negative correlation between gray matter in

regions implicated in pain inhibition and pain intensity, unpleasantness and TMD

duration;

4. a positive correlation between gray matter in regions implicated in the affective

dimension of pain and neuroticism, but a negative correlation between gray matter in

regions implicated in pain inhibition and neuroticism.

120

3.2. Study II. Age-related gray matter abnormalities in patients with TMD

The brain undergoes structural changes due to normal conditions such as aging. For example,

normal aging is characterized by cortical gray matter atrophy [84,99,372,585,627,824], although

gains have also been reported in some brain areas [322,749]. Brain plasticity also occurs with

dysfunction, injury, in specific diseases, and with chronic pain (see Section 2.5.2). The

cumulative effect of chronic pain has been shown to interact with normal aging processes. For

example, accelerated age-related whole brain gray matter atrophy has been reported in FMS

[502] and chronic back pain [27]. Most aging studies in chronic pain have assessed global gray

matter but little is known about the interaction between chronic pain and age in specific brain

areas.

Although the functional significance of gray matter changes is not understood, MRI and

histological studies indicate that changes can be stimulus-dependent. For instance, studies have

demonstrated gray matter changes in response to learning [267], training [266], and repetitive

noxious stimulation [852]. These data suggest that abnormal gray matter progression in chronic

pain may be related to pain duration. While it is plausible that age-related changes specific to

chronic pain are the product of cumulative pain exposure, this hypothesis has not been tested

empirically in TMD.

Therefore, the main aim of study II is to determine whether the cumulative effect of idiopathic

TMD interacts with the normal aging process in focal cortical regions associated with

nociceptive processes.

Specific aims:

1) To determine whether patients with TMD exhibit greater rates of whole brain gray matter

atrophy, compared to healthy, pain-free controls.

2) To determine whether TMD is associated with abnormal gray matter aging in focal

cortical regions associated with pain processes.

3) The degree to which the cumulative effects of pain (i.e. TMD duration) contributes to the

observed age effects.

121

Specific hypotheses:

1) Patients with TMD will have increased rates of whole brain gray matter atrophy,

compared to healthy, pain-free controls.

2) Age-related gray matter changes will be increased in brain regions implicated in pain

perception of patients with TMD,

3) Abnormal gray matter changes in patients with TMD will be driven by TMD duration,

rather than age.

3.3. Study III. White matter brain abnormalities in temporomandibular disorder

Abnormal CNS function is thought to initiate or maintain TMD pain [757] based on TMD

symptomology such as persistent pain, allodynia, and hyperalgesia (sometimes extending to

regions distant from the face) [305,314,390,565,756,883], enhanced temporal summation of pain

to repetitive noxious heat stimuli [563], and dysfunctional DNIC [113,491,756]. Abnormalities

of these centrally-mediated processes suggest that ascending nociceptive pathways and/or

descending pain-modulatory pathways [515] are affected in TMD. Additionally, patients with

TMD can exhibit cognitive [370,379,380] and motor dysfunction [837] possibly related to

abnormalities in brain regions associated with these functions [799,800,839,928].

Studies (including this thesis) indicate that patients with TMD have gray matter abnormalities in

sensory, motor and cognitive/limbic regions (see Section 2.5.2 and Chapter 5). Our finding of a

correlation of thalamic gray matter with TMD duration [617] supports the concept of long-term

peripherally-induced central plasticity. Therefore, although clinical observations have not clearly

identified gross peripheral abnormalities in TMD, the CNV may indeed undergo changes due to

abnormal persistent activity resulting in microstructural abnormalities.

Previous studies examining white matter in other clinical conditions with sensory abnormalities

and/or chronic pain [156,356,558,847] found white matter abnormalities in brain areas involved

in sensory, modulatory and cognitive functions, and some of these white matter abnormalities

correlated with clinical findings. Therefore, studying the correlation between TMD

122

characteristics and measures of white matter integrity can provide insight into whether chronic

pain drives changes in white matter microstructure.

Therefore, the main aim study III is to evaluate white matter brain and CNV abnormalities in

patients with idiopathic TMD and to determine the contribution of pain-related characteristics.

Specific aims:

1) To determine whether there are white matter abnormalities along the CNV of patients

with TMD compared to pain-free controls.

2) To determine whether there are abnormalities in white matter tracts in the brain

associated with sensory, cognitive or motor functions, compared to healthy controls.

3) To determine the connectivity of main regions of white matter disruption.

4) To determine whether there is a link between white matter microstructure and clinical

characteristics of TMD.

Specific hypotheses:

Patients with TMD will have:

1) Abnormal white matter along the CNV;

2) Abnormal white matter in sensory, cognitive and motor tracts;

3) Abnormal connectivity of white matter regions with abnormal microstructure;

4) Abnormalities in white matter will be correlated to pain characteristics (TMD duration,

pain intensity and pain unpleasantness).

123

Chapter 4 General Methods

This project was conducted in a cohort of female patients with TMD and an age-matched group

of control subjects (see below).

4.1. Overview of Project

All study participants underwent two experimental sessions:

• Questionnaires and task training session – This session consisted of handedness and

personality questionnaires. All subjects were also trained in an fMRI task that is not part of

this thesis. Furthermore, patients underwent an interview to characterize their pain, and

completed pain questionnaires.

• Brain imaging session – This session included acquisition of structural MRI sequences to

assess gray and white matter structure. This session also included acquisition of Stroop task

fMRI [928] and resting state fMRI that are not part of this thesis.

4.2. Subject Recruitment

Patients were recruited by dentists working in the Facial Pain Program at the Mount Sinai

Hospital Dental Clinic. Patients were diagnosed with non-traumatic TMD based on the Mount

Sinai Hospital Dental Clinics’ TMD diagnostic criteria (that are based on the TMD-RDC).

Specifically, inclusion criteria included: 1) Self-report of TMD pain in the masticatory muscles

of greater or equal to 4/10 (with ‘0’ being equivalent to no pain at all and ‘10’ representing

maximal pain) for at least 3 months or pain that is aggravated by mandibular function; and 2)

moderate pain to palpation and/or persisting pain after examination in at least 3 masticatory

muscle sites and/or moderate pain to palpation of the TMJ region and/or limitation in mandibular

movement (opening less than 40 mm). Exclusion criteria included: 1) serious metabolic,

rheumatoid or vascular disorders and other diseases; 2) other craniofacial pain disorders,

previously diagnosed psychiatric disorders (e.g., depression, schizophrenia and attention-deficit

hyperactivity disorder) or self-reported history of an abnormal neurological examination; 3) any

contraindication to MRI scanning (e.g., claustrophobia, pregnancy and metal); and 4) use of

psychotropic drugs. All participants were female because of the increased prevalence of TMD in

females. If patients were eligible to participate in the study, the study was outlined to them, and

124

any of their questions were answered. If there were interested in participating in the study, they

were provided with a letter outlining the general goals and methods of the study, and inviting the

patients to participate (see Appendix I: Recruitment letter). If patients chose to participate,

contact information was collected, so that they had a week to reconsider whether to participate in

the study. Nineteen patients with TMD were initially included in the study, of which two were

later excluded because they did not meet inclusion criteria.

The final study cohort included 17 patients with TMD (age ± SD: 33.1 ± 11.9 years; range: 18 –

59 years). Additionally, a group of 17 age-and sex-matched healthy controls (age ± SD: 32.2 ±

10.1 years; range: 20 – 50 years) was recruited from the University and Hospital environments.

To recruit control subjects, the study was advertised by postings within the hospital. Control

subjects would contact us by telephone or email, and undergo a basic screening over the

telephone. The exclusion criteria for the healthy control cohort were the same as those mentioned

for the patient cohort with the addition of a history of past chronic pain.

All study participants provided informed written consent to procedures approved by the

University Health Network and Mount Sinai Hospital Research Ethics Boards (see Appendix II:

Consent Forms).

One additional control subject was recruited and consented to REB approved procedures to

replace one subject subsequently removed from the subcortical analysis (see below) because of

poor segmentation in the pre-processing pipeline.

4.3. Questionnaires

4.3.1. Edinburgh Handedness Inventory

Handedness was determined using the Edinburgh handedness inventory [655] (see Appendix III).

This questionnaire consists of ten questions about which hand is used to perform everyday

manual activities. Subjects report whether they use their right or left hand, and whether they

could use the opposite hand if they so desired. To tabulate, a laterality quotient is calculated with

the following equation:

125

where X(i,R) is the number of right-handed responses for each item, and X(i,L) is the number of

left-handed responses for each item. If a subject reports that the other hand cannot be used for a

given task, then their response is weighted as 2 responses. The range of possible scores is

between -100 and +100. If subjects scores between < -40 he or she is categorized as left-handed;

if the subject scores between -40 and +40, then he or she is considered ambidextrous; and if the

subject scores > +40, he or she is deemed right-handed.

4.3.2. NEO-Five Factor Inventory-Short Form

All subjects completed the NEO (Neuroticism Extraversion Openness) Five Factor Inventory

(NEO-FFI; Psychological Assessment Resources, Lutz, Florida, USA) [178]. Participants were

asked to indicate the degree to which they agree with a statement on a five-point Likert scale

(strongly disagree, disagree, neutral, agree and strongly agree), each of which is coded to a

number (0–4). A total of 15 of 60 questions probe for aspects of neuroticism in this

questionnaire. Neuroticism was the only personality dimension that was assessed.

4.3.3. Pain Catastrophizing Scale

The Pain Catastrophizing Scale [836] (see Appendix IV) was also collected for all participants of

this study, but they were not included in the analysis within this thesis.

4.3.4. Patient Interview

Patients were interviewed to characterize their TMD-related pain (Appendix V). Patients were

also asked to provide a numerical pain score for their current pain intensity and pain

unpleasantness verbally and were asked to report their average pain intensity and unpleasantness

over the last month before scanning. They were specifically asked the following questions:

“Please rate your current/ average pain/unpleasantness rating over the last month on a scale of 0

to 10 (0 = no pain, 10 = worst pain imaginable)”. The duration of TMD was also documented for

each patient.

H =100∗ i=1

10

ΣX(i,R)−i=1

10

ΣX(i,L)

i=1

10

ΣX(i,R)+i=1

10

ΣX(i,L)

126

Furthermore, patients completed the McGill Pain Questionnaire [591] (see Appendix VI). The

McGill Pain questionnaire was used to assess the different dimensions of pain, and thus to

specify the qualia of pain. The questionnaire comprised 102 words classified into three classes

and sixteen subclasses. Different pain disorders have been related to a specific set of words from

the questionnaire. For the current thesis, the McGill Pain Questionnaire was only collected as a

qualitative tool, and was not further assessed. Results are presented in Appendix VII.

4.4. MR Imaging

4.4.1. Study design

All subjects participated in a 1-hour imaging session that included: 1) a low-resolution

anatomical scan for co-registration of fMRI scans, 2) three “task” fMRI scans (not included in

this thesis), 3) a high-resolution T1-weighted whole brain anatomical scan, 4) a resting state

fMRI scan (not included in this thesis), and 5) two runs of DWI covering the whole head. After

brain scanning, subjects were asked to complete a questionnaire related to the task fMRI scans.

Of note, the task fMRI required subjects to wear special goggles to view images, and to use a

response button box. These devices can increase the inhomogeneity during structural MRI

acquisition and so the goggles and the response button-box were removed from the MRI prior to

structural imaging.

4.4.2. Imaging parameters

Brain imaging data were acquired using a 3.0 T GE Signa HDx MRI system (General Electric,

Milwaukee, WI) fitted with an eight-channel phased-array head coil. Subjects were placed supine

on the MRI table and each subject's head was padded to reduce movement. Each subject was

given a “panic button” which they could use if they felt uncomfortable and wanted to stop the

scan.

For gray matter analysis, a whole brain three dimensional (3D) high-resolution anatomical scan

was acquired with a T1-weighted 3D IR-FSPGR sequence: 128 axial slices, 0.94 × 0.94 × 1.5

mm3 voxels, 256 × 256 matrix size, field of view = 24 × 24 cm, one signal average, flip

angle=20°, TE =5 ms, TR =12 ms, inversion time=300 ms.

127

For white matter analysis, two DWI scans were performed (TR = 14,500 ms, field of view: 24

cm x 24 cm, 128 x 128 matrix, 1.875mm x 1.875 mm in-plane resolution, 3 mm thick axial

slices, with Array Spatial Sensitivity Encoding Technique (ASSET) with a factor of 2, maximum

gradient strength = 40 mT/m, maximum slew rate = 150 T/m/s) acquired along 23 non-collinear,

isotropic directions (b = 1,000 s/mm2). Additionally, 2 non-diffusion-weighted scans (b = 0

s/mm2; b0) were acquired at the beginning of each run. The scans covered from the top of the

head to the second cervical spinal process (C2).

4.4.3. Gray Matter Analysis

Two analysis methods were used to assess gray matter: CTA for cortical analysis [316,538,559],

and VBM for subcortical analysis [36] . We used masks of specific brain regions (i.e., ROI

analysis) to test our specific a priori hypotheses; this focus also reduced the statistical burden

associated with the need to adjust the α level in multiple comparisons to regions of no interest

(masks specific to the different studies are described in Chapters 5 and 6).

4.4.3.1. Cortical Thickness Analysis

FreeSurfer software v 4.5 (http://surfer.nmr.mgh.harvard.edu) was used to assess cortical

thickness [198,316,317,319,320]. There are two automated pipelines embedded into the

FreeSurfer analysis pipeline: the cortical (surface) stream and the subcortical (volumetric)

stream. The subcortical stream is out of the scope of this thesis, and so it will not be discussed

further. The cortical stream is described in [198]. The pipeline is as follows: (1) T1-weighted

scans were normalized to Talairach stereotaxic space [844]. To do so, transformation parameters

are computed “by using gradient descent at various scales to maximize the correlation between

an individual volume and an average volume composed of a large number of previously aligned

brains”[198]. The resultant transformation matrix (from individual space to Talairach space) is

stored for later steps. (2) Anatomical T1-weighted images have MRI-susceptibility artifacts and

RF-field inhomogeneities which cause varying intensities and contrast across the volume. These

need to be corrected for proper segmentation. To do so, white matter is used to correct

inhomogeneities and artifacts, because it is more spatially extensive than gray matter, and thus

has minimal partial voluming issues, and allows for low spatial frequency inhomogeneities to be

estimated. Inhomogeneities are corrected by adjusting for the variability in peak white matter

signal between different slices. This process is called “estimating a B1 bias field” [198]. (3) The

128

skull is removed from the brain image. To do so, the inner surface of the skull is modeled using a

deformation-based tessellated ellipsoid template [198] (see Figure 4-1).

Figure 4-1: Template deformation process for skull-stripping in FreeSurfer. Reproduced with

permission from [198].

(4) Intensity values are used to classify white matter voxels as “white matter” or “not white

matter”. (5) Planes based on the Talairach location of the corpus callosum and the pons are

applied to separate the two hemispheres, remove the cerebellum and the brainstem (see Figure 4-

2).

Figure 4-2: Cutting planes overlaid on sections through white matter-labeled volume.

Reproduced with permission from [198].

(6) The surface that is the border of white matter and gray matter (i.e., the outer surface of the

white matter) is modeled as an initial surface for each hemisphere (see Figure 4.5). The modeled

surface (called “white surface”; see Figure 4-3 and 4-5) undergoes smoothing based on local

MRI intensity values to better reflect the gray/white border. (7) The white matter surface is then

inflated to the gray/CSF surface (the “pial surface”; see Figure 4-4 and 4-5). Cortical thickness is

defined as the distance between the white surface and the pial surface.

129

Figure 4-3: White surface overlaid on original T1 scan. Reproduced with permission from [198].

Figure 4-4: Pial surface overlaid on original T1 scan. Reproduced with permission from [198].

Figure 4-5: Original (left), gray/white boundary (middle), and pial surface (right)

reconstructions of a left hemisphere. Reproduced with permission from [198].

Each subject’s homologous gyri and sulci were aligned to the standard average brain provided in

FreeSurfer. A Gaussian spatial smoothing kernel of 6 mm full-width half-maximum (FWHM)

130

was applied to compensate for topographical heterogeneity prior to statistical analysis. For this

thesis, the final sampling distance between adjacent points on the cortex (or two vertices) was

0.71 mm.

4.4.3.2. Voxel-Based Morphometry

A VBM analysis was used to measure the gray matter volume of subcortical structures.

Preprocessing of T1-weighted images and statistical analyses were performed in the SPM v5

software package (http://www.fil.ion.ucl.ac.uk/spm/software/spm5/) running under Matlab (The

Mathworks, Inc, Natick, MA). We used the VBM 5.1 toolbox (http://dbm.neuro.uni-

jena.de/vbm) implemented in SPM 5 to perform VBM. The methods are described in [36]. The

pipeline included the following steps: DICOM file format images were converted to the NIfTI-2

file format using the DICOM import tool in SPM. Next, images were normalized to a standard

template (International Consortium for Brain Mapping (ICBM) template), using a 12-parameter

affine transformation. Next, the images were intensity normalized, and segmented based on

tissue types using a Markov random field model and prior probability maps. The segmented gray

matter posterior probability maps, underwent Jacobian modulation. The images were then spatial

smoothed with a 10 mm FWHM Gaussian kernel. An absolute threshold mask of 0.10 (i.e., the

posterior probability of being gray matter must exceed 0.1) was set to restrict the results to gray

matter. All coordinates were converted from Montreal Neurological Institute Stereotaxic Space

(MNI; also the ICBM Standard Template) to Talairach space [844] using Yale BioImage Suite

non-linear MNI to Talairach Coordinate Converter [510]

(http://www.bioimagesuite.org/Mni2Tal/index.html) for results in Chapter 5, and the Lancaster

transform [513] implemented in GingerALE v.2.0.4 (http://www.brainmap.org/ale) in Chapter 6.

4.4.4. White matter analysis

Diffusion-weighted images were converted from DICOM image file format to the NIfTI-1 file

format (for FSL) using the dcm2nii application, as part of Chris Rorden’s MRIcon software

package (http://www.mccauslandcenter.sc.edu/mricro/mricron/dcm2nii.html). The NIfTI-1

diffusion-weighted images were imported into FSL v.4.1.8 (http://www.fmrib.ox.ac.uk/fsl) for

quality control [816]. Pre-processing included eddy current and motion artifact correction using

the FSL diffusion toolbox (FDT) [459]. This was achieved by aligning each 3D volume (each

131

DWI direction forms a 3D image; the image with all of the directions is called a 4D image) to the

b0 (non-diffusion-weighted image) collected at the beginning of each scan. In this thesis, two b0

images were acquired for each run, and the 4D dataset was aligned to the second b0 image. Next,

each subject’s two DWI runs were averaged to a single volume to achieve a greater signal-to-

noise ratio. Then, individual brain masks were created using the Brain Extraction Tool (BET)

[814] to cover the entire brain and exclude the skull, the eyes and other non-brain structures. The

parameters were adjusted for each individual subject to obtain maximal brain coverage, while

excluding as much of the non-brain tissue. These images were then processed through two

different pipelines for (1) voxelwise analysis using TBSS and (2) probabilistic tractography.

4.4.4.1. Tract-based spatial statistics

TBSS v.1.2 [815,817] was used to compare FA between the TMD and healthy control cohorts.

FA maps were created using DTIFIT tool in FDT. For TBSS, FA maps underwent non-linear

registration to a 1 x 1 x 1 mm FA map in standard space (FMRIB58_FA, available in FSL), a

mean image derived from all the subjects was created and thinned to represent the centre of

major white matter tracts common to all subjects, forming a white matter skeleton. Each

subject’s peak FA value perpendicular to the skeleton was then projected onto the skeleton. One

skeleton was created for analysis of group differences that included all study participants, and a

second skeleton based on the patients’ FA maps was created for correlation analyses of TMD

pain characteristics (see below). The mean skeleton images were set at a threshold of FA > 0.2 to

include FA values that are related to white matter [815]. Other DTI metrics were also evaluated

to characterize FA findings (see Figure 7-1). Additionally, a separate patient group analysis was

performed to assess which areas of the skeleton correlate with TMD characteristics (pain

intensity, unpleasantness, duration).

4.4.4.2. TBSS with other DTI metrics

To gain more insight into FA findings, other DTI metrics were assessed in the clusters with

significant group difference. λ1 is thought to reflect diffusion along a tract, and reductions in this

value suggest disruptions along the tract diffusivity. RD is believed to reflect changes in neural

membrane permeability and, to some extent, myelination [66,67]. RD is calculated by averaging

the two radial vectors of the tensor model (λ2 and λ3). Finally, MD was measured because it is

associated to œdema and inflammation. Figure 7-1 provides a schematic of these DTI metrics.

132

DTI metrics (FA, MD, λ1, λ 2 and λ3) are calculated using DTIFIT in the FDT toolbox. The TBSS

skeleton was created based on FA, and peak FA values are projected onto the skeleton. For

statistical analysis of the other DTI metrics, MD, RD and λ1 values from the same voxels are

projected onto the skeleton. To test for group differences, we performed a multivariate analysis

of variance (MANOVA) with each of the DTI metrics as dependent variables, and group as an

independent variable.

4.4.4.3. Probabilistic tractography

We sought to better characterize the corpus callosum cluster identified using cluster-mass

correction (see section 7.3.3) by elucidating its connectivity. Therefore, we performed multi-fibre

probabilistic tractography (http://www.fmrib.ox.ac.uk/fsl/fdt/index.html) [73,74] to determine

pathways passing through this cluster, and to elucidate its connectivity. To do so, the

preprocessed diffusion-weighted images were downsampled in order to create isotropic voxels (3

x 3 x 3 mm). The images were then processed in FDT. Probability density functions (PDF) on up

to 2 principal fibre directions were estimated at each voxel in the brain. Multi-fibre tractography

was used and 5000 samples were drawn from each seed voxel along the PDFs. Significant

clusters from the TBSS group difference analysis were binarised and used as seeds for

tractography. Together, each seed voxel’s streamlines provide an estimate of its connectivity.

When a streamline reaches a voxel with more than one fibre direction, the streamline follows the

direction closest to the direction at which it arrives at the voxel. The pathways generated by the

algorithm represent the number of samples that have passed through a voxel. The pathways

generated in each subject were thresholded at 10 samples (of the 5000 generated from each seed

voxel; i.e., 0.2%) to eliminate spurious connections. These thresholds provided consistent tracts

between subjects. For visualization, each of the subject’s tracts were binarised and overlaid on a

standard brain to produce a probabilistic map of the pathways for controls and patients. The

values in these maps represent the number of subjects who share a common pathway.

133

Chapter 5 STUDY I: Contributions of chronic pain and neuroticism to

abnormal forebrain gray matter in TMD

This study was published in Neuroimage:

Moayedi M, Weissman-Fogel I, Crawley AP, Goldberg MB, Freeman BV, Tenenbaum HC,

Davis KD. Contribution of chronic pain and neuroticism to abnormal forebrain gray matter in

patients with temporomandibular disorder. Neuroimage 2011; 55(1): 277-86

5.1. Introduction

TMD represent the most common orofacial chronic pain disorder, and are more prevalent in

women than in men [721]. TMD can be idiopathic in that there may not be any clear peripheral

etiological factors identifiable [283,286,287,653]. Thus, central mechanisms have been proposed

to account for TMD pain [757].

Although a clear pattern of change has yet to be determined, previous structural MRI studies of

chronic pain populations have found both increases and decreases in gray matter. For instance,

some studies of headache and chronic facial pain populations have found that patients with

chronic pain had gray matter increases in regions likely associated with pain perception

[207,575,650,963]. Additionally, most studies of chronic pain patients have found reduced gray

matter in cortical regions likely associated with antinociception and limbic function

[98,356,575]. Interestingly, some studies have also reported gray matter loss in cortical and

subcortical motor areas [22,575,763]. However, the increases are not limited to regions thought

by some to be associated with pain perception, and the decreases are not limited to pain-

modulatory regions. Although the role of motor regions in pain is not fully established, there is

evidence suggesting these areas play a role in pain modulation [3,117,189,342,345,547]. In

support of this concept are the motor abnormalities that can accompany chronic pain

[471,492,839,928], possibly related to nocifensive behaviour [634].

There are two main routes by which the central mechanisms may contribute to the development

and/or maintenance of chronic pains such as TMD. One possibility is that long-term nociceptive

input into the brain induces maladaptive brain plasticity, which may play a role in maintaining

pain [12,950]. For example, a recent study demonstrated that experimental pain that increased

134

gray matter in nociceptive regions [852], induced pain habituation over time that was

accompanied by decreased activity within nociceptive areas and increased activity within the

antinociceptive system [94]. Chronic pain patients, however, may not be able to adapt in this way

to nociceptive activity. For example, neuroimaging studies of chronic pain have shown

hyperactivity in nociceptive regions, and hypoactivity in antinociceptive regions [23,541].

Chronic pain patients’ inability to habituate to increased nociceptive activity may be related to a

reduced capacity of the brain to dampen pain by descending (top-down) controls [95]. Indeed,

many structural MRI studies have found gray matter differences in chronic pain populations

associated with pain-related characteristics (intensity, unpleasantness or duration)

[27,98,575,741,963].

The second route by which the CNS may contribute to the development and/or maintenance of

chronic pain relates to inherent personality-related factors that reduce the brain’s capacity to

modulate nociceptive input. This poor pain control represents a vulnerability to develop chronic

pain. For example, there is evidence that neuroticism may be associated with pain-related

suffering [408], pain sensitivity [177,373,907], nerve injury outcomes and neuropathic pain

[848] and inhibition of negative thoughts [178]. However, not all chronic pain patients have high

neuroticism scores, and not all persons with neuroticism have chronic pain [179]. Therefore,

neuroticism alone is not sufficient to develop chronic pain. Rather, the normal relationship

between neuroticism and brain structure and function may be disrupted within regions involved

in pain modulation, such as the OFC [954] or the mPFC [245,386] and this could facilitate or

maintain chronic pain.

Thus, in the current study we examined gray matter abnormalities in patients with idiopathic

TMD and focused our investigation on the contribution of pain-related characteristics and

neuroticism. Towards this goal, we measured gray matter in patients who had suffered from

TMD over a range of pain intensities, unpleasantness and for varying durations, and neuroticism

scores. Based on the aforementioned behavioural and neuroimaging studies, we specifically

tested the hypotheses that patients with TMD will have: 1) increased gray matter in areas

associated with pain perception; 2) reduced gray matter in areas associated with pain modulation

and motor function; 3) gray matter positively correlates with pain intensity, unpleasantness and

TMD duration within areas associated with pain perception areas and negatively correlates with

gray matter in areas associated with antinociception; 4) negative correlation between neuroticism

135

and gray matter in regions implicated in pain modulation, and positive correlation in regions

implicated in the affective dimension of pain, because of the interaction between affective

processing and pain modulation in TMD [881].

5.2. Methods

A detailed description of methods is provided in Chapter 4.

5.2.1. Subjects

A group of 17 females with idiopathic TMD (mean age ± SD: 33.1 ± 11.9 years) and 17 healthy

females (mean age ± SD: 32.2 ± 10.1 years) provided informed written consent to procedures

approved by the University Health Network and Mount Sinai Hospital Research Ethics Boards.

All subjects were right-handed. Patients with TMD were screened with the Mount Sinai Hospital

Dental Clinic TMD diagnostic criteria, based on the TMD-RDC [285]. Inclusion criteria

included: 1) TMD pain masticatory muscle region greater or equal to 4/10 for at least 3 months

or pain that is aggravated by mandibular function; 2) moderate pain to palpation and/or persisting

pain after examination in at least 3 muscle sites and/or moderate pain to palpation of the TMJ

region and/or limitation in the mandibular movement (opening less than 40mm). Patients were

asked to remain analgesic-free for 24 hours prior to scanning, as functional data was also being

collected during the scanning session. For both patients and control subjects, exclusion criteria

included: 1) serious metabolic, rheumatoid or vascular disorders and other diseases; 2) other

craniofacial pain disorders, previously diagnosed psychiatric disorders (e.g., depression,

schizophrenia, ADHD) or self-reported history of an abnormal neurological examination; 3) any

contraindication to MRI scanning (e.g., claustrophobia, metal); 4) use of psychotropic drugs. In

addition, healthy controls were not eligible for the study if they had a history of chronic pain.

5.2.2. Questionnaires

Each participant completed the NEO-FFI [178]. The NEO-FFI is a self-report questionnaire

comprising of 60 statements. Participants were asked to indicate the degree to which they agree

with a statement on a five-point scale (strongly disagree, disagree, neutral, agree, strongly agree),

each of which is coded to a number (0-4). A total of 15 of 60 questions probe for aspects of

neuroticism in this questionnaire.

136

Patients were also asked to rate their current pain and unpleasantness and their average pain

intensity and unpleasantness over the last month before scanning with the following instructions:

“Please rate you current/average pain/unpleasantness rating over the last month on a scale of 0 to

10 (0 = no pain, 10 = worst pain imaginable)”. The duration of the patients’ TMD was also

recorded for each patient.

5.2.3. Imaging parameters

Brain imaging data were acquired using a 3T GE MRI system fitted with an eight-channel

phased array head coil. Subjects were placed supine on the MRI table and each subject’s head

was padded to reduce movement. A whole brain three dimensional (3D) high-resolution

anatomical scan was acquired with a T1-weighted 3D IR-FSPGR sequence: 128 axial slices, 0.94

x 0.94 x 1.5 mm3 voxels, 256 x 256 matrix size, field of view = 24 x 24 cm, one signal average,

flip angle = 45°, TE = 5ms, TR = 12ms, TI = 300ms.

5.2.4. Structural brain imaging analysis

Because we were interested in both cortical and subcortical structures, we used two analysis

approaches to measure gray matter from high-resolution MRI images. To evaluate differences in

cortical thickness (measured in mm) we used CTA [316,538,559], and VBM [36] was used to

measure subcortical gray matter density. We used masks of specific brain regions (i.e., ROI

analysis) to test our specific a priori hypotheses; this focus also reduced multiple comparisons to

regions of no interest (see below).

5.2.4.1. Cortical thickness analysis

CTA was carried out using the Freesurfer software (http://surfer.nmr.mgh.harvard.edu); methods

are described in full detail elsewhere (see:[198,316,317,319,320]. Briefly, pre-processing

included intensity normalization, skull stripping, separation of the hemispheres, and gray matter

segmentation. The white matter/gray matter border (i.e., white surface) and gray matter/CSF

border (i.e., pial surface) were identified and modeled as surfaces. The software then calculated

the distance between the two borders at every point on the cortex, for each hemisphere. The gray

matter surface was then warped such that homologous gyri and sulci were aligned across all

subjects. Each individual subject’s cortex underwent automatic anatomical parcellation, and each

sulcus and gyrus was labeled during pre-processing [321]. To restrict our search to regions

137

hypothesized to be involved in TMD chronic pain, we used the cortical parcellations map

implemented into FreeSurfer (aparcs2005) on a standard subject (fsaverage). Two masks were

created: (1) a sensorimotor/pain mask which included S1, S2, M1, and MCC (see Figure 1 a);

and (2) a cognitive/modulatory mask of the frontal lobe that included the OFC, PFC, insula and

the ACC/MCC (see Figure 1 b). The MCC has been implicated in both nociception and pain

modulation [23] and so this region was included in both masks. To compensate for topographical

heterogeneity amongst the different subjects, a 6 mm FWHM Gaussian spatial smoothing kernel

was applied to the data prior to statistical analysis. A vertex represents a point on a two-

dimensional sheet, and, in this study, the distance between two vertices is 0.71 mm.

5.2.4.2. Subcortical analysis with VBM

Image processing and statistical analyses were performed in the SPM5 software package

(http://www.fil.ion.ucl.ac.uk/spm/software/spm5/) running under Matlab (Mathworks). The

VBM analysis used Gaser’s VBM 5.1 toolbox (http://dbm.neuro.uni-jena.de/vbm) within SPM 5;

detailed methodology is described in elsewhere [36]. Briefly, the preprocessing included setting

the origin of the image at the anterior commissure of each subject, spatial normalization to the

ICBM-152 template, gray matter segmentation, Jacobian modulation to adjust for the effects of

spatial normalization, and spatial smoothing with a 10 mm FWHM Gaussian kernel. Total

intracranial volume (TIV) was calculated for each subject from the volume estimates output

during segmentation. An absolute threshold mask of 0.10 was applied to restrict the results to

gray matter only. We constructed a single mask for our subcortical VBM analysis (See figure 1

c). We used prefabricated anatomical masks in WFU Pickatlas

(http://www.nitrc.org/projects/wfu_pickatlas). The subcortical mask was used to investigate

regions that cannot be investigated using CTA, and comprised the basal ganglia, amygdala,

brainstem and thalamus.

5.2.5. Statistical analyses

5.2.5.1. Demographics

Group differences between characteristics of the patients and controls were evaluated using a

multivariate analysis of variance in SPSS 18.0 (http://www.spss.com/). Age and neuroticism

scores were compared between the groups. The significance threshold was set at p < 0.05.

138

5.2.5.2. Gray matter group differences

CTA

To restrict our search to regions hypothesized to be involved in TMD chronic pain, we used the

cortical parcellations map implemented into FreeSurfer (aparcs2005) on a standard subject

(fsaverage). Two masks were created: (1) a sensorimotor/pain mask which included S1, S2, M1,

and the MCC (see Figure 5-1 a); and (2) a cognitive/modulatory mask of the frontal lobe that

included the OFC, PFC, insula and the ACC and MCC (see Figure 5-1 b). The MCC has been

implicated in both nociception and pain modulation [23] and so this region was included in both

masks.

To test our first hypothesis that patients have thicker cortex in perception areas, we ran a general

linear model (GLM) testing for group differences in cortical thickness at every vertex within the

sensorimotor/pain mask (see Figure 5-1a). Age was included as a regressor of no interest. To test

our second hypothesis that patients have thinner cortex in modulatory/cognitive areas, we ran a

second GLM testing for group differences at every vertex within the cognitive/modulatory mask

(see Figure 5-1b). Percent change was calculated using the following formula: ((Mean of

patients’ thickness - mean of controls’ thickness)/mean of controls’ thickness) × 100.

All data were thresholded at an experiment-wide Bonferroni corrected p < 0.05. To do so, we

calculated the corresponding cluster-wide p-value to reach significance and so the experiment-

wide correction was based on conducting 6 tests: (1) thicker cortex in pain perception regions,

(2) thinner cortex in cognitive/modulatory regions, (3) the contribution of pain intensity, (4) the

contribution of pain unpleasantness, (5) the contribution of TMD duration to cortical

abnormalities, and (6) group by neuroticism interactions in cortex. Thus the correction for the

threshold was 0.05/6 = 0.008. Therefore, to achieve a Bonferroni-corrected threshold of p < 0.05,

CTA data were thresholded at image-wide threshold of p < 0.008. Our two masks (see Figure 5-

1) were of different sizes; thus, we used Monte Carlo permutation to calculate an image-wide

threshold for each. For the sensorimotor/pain mask, this was derived from an uncorrected

voxelwise p < 0.01 and 190 contiguous vertices (379 voxels), and for the cognitive/modulatory

mask, this was derived from an uncorrected voxelwise p < 0.01 and 220 contiguous vertices (436

voxels). Monte Carlo simulations were run with 1000 iterations using AlphaSim

(http://afni.nimh.nih.gov/afni/doc/manual/AlphaSim), as we have used previously [847].

139

VBM

We constructed a single mask for our subcortical VBM analysis (See figure 5-1c). We used

prefabricated anatomical masks in WFU Pickatlas (http://www.nitrc.org/projects/wfu_pickatlas).

The subcortical mask was used to investigate regions that cannot be investigated using CTA, and

comprised the basal ganglia, amygdala, brainstem and thalamus. To test both our first and second

hypotheses, we used student’s t-test to test for group differences in each voxel within the

subcortical mask (see Figure 5-1c). The general linear model included age and TIV as covariates

of no interest. All data were thresholded at a voxelwise corrected p < 0.05 using false discovery

rate (FDR), as implemented in SPM 5.

5.2.5.3. Clinical correlates

To assess pain-related plasticity, we correlated pain intensity, pain unpleasantness and TMD

duration to gray matter. For all the correlations, age was included as a regressor of no interest.

However, for the correlation of gray matter to TMD duration we could not include age as a

covariate, as these two explanatory variables (age and TMD duration) are correlated (r = 0.536, p

= 0.027). For the CTA analysis, pain intensity, pain unpleasantness and TMD duration

correlations were performed within the sensorimotor/pain mask and the cognitive/modulatory

mask. For the VBM analysis, we ran correlations in each voxel within our subcortical mask, as

previously described. TIV was included as a covariate of no interest in every model.

5.2.5.4. Neuroticism effects

To determine whether neuroticism may be affected by chronic pain in TMD, we performed a

Pearson correlation between TMD duration and neuroticism scores in the patient group.

To investigate whether neuroticism has an abnormal relationship to gray matter, and thus

contribute to TMD onset, we performed an interaction analysis. Specifically, in CTA, we

performed a group by neuroticism interaction using the FreeSurfer software on a vertex-wise

basis within the cognitive/modulatory mask. In the VBM analysis, to test this hypothesis, we

performed a group by neuroticism interaction in each voxel within our subcortical mask.

140

5.3. Results

5.3.1. Patient demographics

The characteristics of individual patients are shown in Table 5-1. Patients and controls did not

differ in age (patients: mean age ± SD: 33.1 ± 11.9 years; controls: 32.2 ± 10.1 years; p = 0.94)

or in neuroticism scores (patients: mean age ± SD: 19.35 ± 6.84; controls: 18.35 ± 7.95; p =

0.70). Of importance for this study is that there was variability in the durations of TMD, intensity

and unpleasantness of pain that facilitated assessment of these factors in correlation analyses.

The duration of TMD ranged from 0.75 – 30 years (mean ± SD: 9.8 ± 8.2 years), while pain

intensity ranged from 2 – 7 (mean ± SD: 4.3 ± 1.8), and unpleasantness ranged from 1 – 8 (mean

± SD: 5.4 ± 2.1). Interestingly, TMD pain intensity and pain unpleasantness scores were not

significantly correlated (r = 0.40, p = 0.12), unlike the tight correlation of these dimensions in

healthy subjects in acute pain paradigms [717]. Patients reported that their TMD pain was

confined to the TMJ (n = 6), the muscles of mastication (n = 3), or both the joint and muscles (n

= 8). Most patients reported bilateral pain (n = 13), but some reported unilateral pain (three right-

sided, one left-sided). The duration of TMD significantly correlated with the patients’ age (r =

0.536, p = 0.027).

Patient neuroticism scores showed no significant relationship to TMD duration (r = 0.33, p =

0.21).

5.3.2. Group differences: S1 and frontal thickening in TMD

There were prominent group differences (p < 0.05, corrected) in gray matter thickness in the

patients with TMD compared to controls (see Table 5-2, Figure 5-2). In line with our first

hypothesis, the ventrolateral S1 was 22% thicker in the TMD group compared to controls. This

region of thickening was located in a region consistent with the somatosensory representation of

the face [205,630]. However, contrary to our second hypothesis, we did not identify statistically

significant cortical thinning in any pain modulation or motor region. Unexpectedly though, we

identified regions in the left frontal pole and vlPFC that were thicker in the patient group

compared to controls by 17% and 15% respectively (see Table 5-2 and Figure 5-2). The VBM

analysis did not reveal any significant subcortical differences between patients and controls.

141

5.3.3. Effect of chronic pain intensity and unpleasantness

We also examined the relation between pain intensity and pain unpleasantness in gray matter

within our cortical masks. Cortical thinning that correlated with the degree of TMD pain

intensity or unpleasantness was found within three brain regions (see Figure 5-3 and Table 5-3).

A significant negative correlation was found between pain intensity and gray matter thickness in

the aMCC (BA32; r = -0.83), and in the ventrolateral aspect of M1 in a region consistent with the

representation of the face [952] (r = -0.83). Our test of the impact of pain unpleasantness on

cortical thickness within our two masks revealed a significant effect only in left OFC. This

region showed a negative correlation between gray matter thickness and unpleasantness (BA 47;

r = -0.74). Interestingly, in these three brain areas, the patients reporting intense pain had gray

matter that was thinner than the healthy controls, but the individuals with only mild pain had

gray matter either slightly thicker than the controls or within the normal range. Of note is that we

did not identify any significant gray matter subcortical correlations with pain intensity or

unpleasantness.

5.3.4. Effect of TMD chronicity

The subcortical VBM analysis revealed that gray matter in the sensory thalamus was positively

correlated to the length of time that patients had been suffering with TMD (r = 0.91; see Table 5-

4 and Figure 5-4 for details). This region of gray matter increase included the left PO, the left

VL, and bilateral VPM of the thalamus. Interestingly, the patients who had TMD for only a few

years showed gray matter within the normal range (see Figure 5-4), but patients who had TMD

for 7 or more years showed a stronger relationship between duration and thalamic gray matter.

We did not find any significant correlations between TMD duration and gray matter thickness in

the a priori hypothesized cortical ROI.

5.3.5. Effect of neuroticism

Because of the expectation that neuroticism manifests differently in patients than healthy

controls, we tested whether there was neuroticism by group interaction in the

cognitive/modulation mask. The controls showed a negative correlation (r = -0.32) between gray

matter thickness in the left OFC (BA11; Talairach coordinates: -18, 22, -18) and neuroticism

scores, whereas patients show a positive relationship (r = 0.61) in this region. The area of the

interaction in this ROI was 517 mm2 (peak vertex t-score = -2.87; p < 0.05, experiment-wide

142

Bonferroni corrected; see Figure 5-5). We did not identify any significant neuroticism by group

interactions in the subcortical VBM analysis.

143

Table 5-1: Patient demographics.

# Age

TMD

Dur

(yrs)

Pain

Intensity

Pain

Unpl

Pain

Sites

TMD

Laterality

Other

pains Meds

Oral

contraceptive

1 22 2 4 5 J Bilateral Knee pain NSAID Yes

2 20 3 3 3 M, J Bilateral A*, cyc* No

3 24 7 4 8 J Right Headache N, P Yes

4 38 20 7 6 J Bilateral NSAID No

5 42 0.75 2 3 M Bilateral F* No

6 33 4 5 6 M, J Bilateral Shoulder,

Neck

A*, F* No

7 28 17 3 5 M, J Left No

8 34 14 2

8

J Bilateral

A*, F*,

Hy* No

9 50 10 7 6 M Bilateral A, Ch, Di No

10 59 13 7 7 M, J Bilateral Whiplash

(neck and

shoulder),

Gastric

Distress

No

11 18 3 6 8 M, J Bilateral Yes

12 34 15 2 3 M Right Neck,

Sciatica

No

13 52 30 4 5 M, J Bilateral A No

14 31 2 6 5 M, J Right Ovarian

cyst

N No

15 33 17 5 8 M, J Bilateral A, F,

NSAID

No

16 22 1 4 1 J Bilateral Yes

17 23 8 2 5 J Bilateral Yes

Abbreviations: M= masticatory muscles; J=TMJ; NSAID: Non-steroidal anti-inflammatory, over

the counter, as needed; A: Arthrotec; cyc: cyclobenzapine; Ch: Champix; Di: Dixarit

(Clonidine); F: Flexoril; Hy: Hydromorphone; Meds: Medications; N: Naproxen; P: Prevacid;

TMD Dur: TMD duration; Unpl: unpleasantness; yrs: years. The asterisk (*) denotes that

subjects discontinued the use of the drug prior to our study. A and N are also NSAIDs, but were

144

patients were prescribed to take these medications daily- rather than when needed. Gastric

distress was described as “burning in the stomach” by the patient.

145

Table 5-2: Group differences in cortical thickness.

Peak vertex Talairach coordinates are reported. All results are significant at p < 0.05, corrected

for multiple comparisons. Abbreviations: R: right; L: left; BA: Brodmann’s Area; FP: frontal

polar cortex; S1: primary somatosensory cortex; vlPFC: ventrolateral prefrontal cortex.

Contrast Region BA

TAL of Peak Area

(mm2)

Peak

T-score X Y Z

Patients > Controls R S1 2 50 -18 37 1132 5.18

L FP 10 -30 45 6 569 4.87

L vlPFC 9/10 -38 35 4 531 3.41

146

Table 5-3: Cortical thickness negatively correlates with TMD pain intensity or unpleasantness.

Peak vertex Talairach coordinates (TAL) are reported. All results are significant at p < 0.05,

corrected for multiple comparisons. Abbreviations: L: left; aMCC: anterior mid-cingulate cortex;

M1: primary motor cortex; OFC: orbitofrontal cortex.

TMD attribute Region BA

TAL of Peak Area

(mm2)

Peak

T-score X Y Z

Pain Intensity L aMCC 32 -4 18 19 718 - 3.94

L M1 4 -57 -3 13 674 - 4.15

Unpleasantness L OFC 11/47 -27 16 -19 472 - 4.22

147

L R

R L

L R

R L

b)

a)

R L c)

148

Figure 5-1: Masks used to restrict analyses. Three masks were used to restrict the analyses to

regions of interest: a) the sensorimotor/pain mask included cortical sensorimotor regions

(primary and secondary somatosensory cortices (S1, S2), the primary motor cortex (M1), and the

mid-cingulate cortex (MCC), and b) the cognitive/modulatory mask which included the insula,

the prefrontal cortex (PFC), the orbitofrontal cortex (OFC), and the cingulate cortex

(ACC/MCC), c) a subcortical mask used for voxel-based morphometry, included the basal

ganglia, thalamus, amygdala and the brainstem. CTA results are overlaid onto the fsaverage

standard brain in FreeSurfer.

149

Figure 5-2: Cortical thickening in TMD. The CTA GLM group analyses within the cortical

masks identified three cortical regions (shown in blue on the brain images) that showed

significant group differences at a corrected p < 0.05. The top panel shows that the TMD patient

group had thicker right primary somatosensory cortex (S1) (peak T-score = 5.18) compared to

controls, with the effect of age regressed out. The bottom panel shows that the TMD patient

group has thicker ventrolateral (peak T-score = 3.41) and frontal polar cortices (peak T-score =

4.87). All labeled regions are significant at a corrected p < 0.05. The graphs indicate the mean

cortical thickness ± SE (age not regressed out). CTA results are overlaid onto the fsaverage

standard brain in FreeSurfer.

R

L

vlPFC

FP

FP

vlPFC

S1 S1

150

Figure 5-3: TMD pain intensity and unpleasantness are negatively correlated to modulatory

regions. The CTA regressions within the cortical masks in the TMD group identified three

cortical regions (shown in blue on the brain images) that were significantly correlated with

individual pain scores at a corrected p < 0.05. The top panel shows that pain intensity has a

negative correlation to anterior mid-cingulate cortex (aMCC) (r = - 0.83, peak T-score = -3.94).

L

L

M1

aMCC

L

OFC

151

The middle panel shows that pain intensity is negatively correlated to cortical thickness in the

left primary motor cortex (M1) (r = -0.83, peak T-score = -4.15). The bottom panel shows that

pain unpleasantness negatively correlated to cortical thickness in the left orbitofrontal cortex

(OFC; r = -0.75, peak T-score = -4.22), with the effect of age regressed out). All labeled regions

are significant at a corrected p < 0.05. The graphs indicate the mean cortical thickness ± SE for

each subject versus pain intensity (age not regressed out). CTA results are overlaid onto the

fsaverage standard brain in FreeSurfer.

152

Figure 5-4: TMD duration is related to plasticity in the thalamus. The TMD patient group

showed a positive correlation between TMD duration (measured in years) and gray matter in a

cluster in the sensory nuclei in the left thalamus (posterior (PO), ventral lateral (VL), ventral

posterior medial (VPM) (2312 voxels, peak T-score = 7.84, r = 0.91). The finding is significant

at a false-discovery rate corrected p < 0.013. The graph indicates each TMD patient’s gray matter

volume versus her individual TMD duration (in years). The dashed lines represent the mean gray

matter volume of controls ± SE. VBM results are displayed on the single subject standard brain

in SPM.

L

153

Figure 5-5: Patients with TMD showed a positive relationship between OFC thickness and

neuroticism, which is considered to be abnormal. Neuroticism was correlated with changes in the

left orbitofrontal cortex (OFC; peak T-score = 3.97), with the effect of age regressed out. For all

correlations, mean cortical thickness ± SE (age not regressed out) versus the raw neuroticism

score was graphed. CTA results were overlaid onto the fsaverage standard brain in FreeSurfer.

L

OFC

154

5.4. Discussion

This structural imaging study identified striking abnormalities in patients with chronic idiopathic

TMD pain and highlights the contribution of both TMD-related and neuroticism-related factors.

Our key findings were that, compared to controls, patients with TMD had 1) cortical thickening

of the S1, frontal pole and vlPFC, 2) pain intensity-dependent cortical thinning in the aMCC and

M1 and pain unpleasantness-dependent cortical thinning in the OFC, 3) TMD duration-

dependent gray matter increase in the sensory thalamus, and 4) a positive correlation between

cortical thickness in the OFC and neuroticism, in contrast to a normally negative correlation.

5.4.1. S1 and thalamic gray matter increases in TMD

Our finding of S1 thickening in TMD is consistent with other gray matter studies of chronic pain

in the trigeminal system [207]. However, trigeminal chronic pains are heterogeneous, and so may

produce different patterns of gray matter changes specific to the particular symptomology or

pathology. Nonetheless, the convergence of increased S1 thickness suggests that there may be

prolonged or repeated barrage of nociceptive input to the cortex from the thalamus. Evidence for

this hypothesis comes from a study by Teutsch and colleagues [852] that showed persistent

noxious stimulation in healthy subjects can induce increased gray matter in S1. Since we also

found that in patients with TMD there was a positive correlation between thalamic gray matter

and TMD duration, it is possible that sustained trigeminothalamic nociceptive activity over time

leads to increased gray matter in sensory thalamus. The patients included in our study have a

larger range of TMD durations (0.75-30 years) compared to the study by Younger et al. [962] (0-

11 years). Thalamic changes were more apparent in patients with longer reports of TMD

duration, which may explain why our findings are different than those reported by Younger and

colleagues [963]. Increased firing to the thalamus may lead to increased thalamocortical activity

to S1 and plasticity [950]. The cellular basis of cortical thickening (plasticity) is not yet

established but is thought to include increases in synaptic boutons, dendritic branching, glial

cells, and/or neurones [575,608].

5.4.2. Cortical thickening in cognitive and modulatory regions in TMD

We found that patients with TMD had cortical thickening in the frontal pole. These findings

diverge from those of Younger et al. [963] study. A factor that may have contributed to the

155

different findings is that their study examined patients with myofascial TMD, whereas our TMD

group consisted of a somewhat more clinically representative group of mixed muscular and/or

joint TMD.

Activation of the frontal pole has been reported during spontaneous pain in patients with chronic

pain [47]. The frontal pole has also been implicated in a number of complex executive cognitive

functions such as learning behavioural routines [457,496,834], cognitive branching (the ability to

put a pending task on hold to execute an ongoing one) [497], behavioural flexibility/adaptability

[105] and post-hoc monitoring or evaluating decisions based on feedback [875]. Therefore, the

frontal pole may process the cognitive dimension of pain, which suggests that pain has a

cognitive load. We propose that the cognitive load of pain may require constant engagement of

the frontal polar cortex for cognitive branching. That is, chronic pain may need to be put “on

hold” in order for a patient to engage in another competing cognitive load. However, if the

cognitive branching system is limited, a chronic pain load may impact the ability to properly

complete tasks [529]. In support of this, we have recently shown that patients with TMD have

sluggish reaction times to low and high conflict cognitive interference tasks, but not to a simple

sensorimotor task [928]. Further evidence for cognitive branching of pain comes from studies

showing that an increase in cognitive load modulates acute pain perception in healthy controls

[935]. Therefore, one possibility for the thickening of the frontal polar cortex is that chronic

TMD pain bears a cognitive load that the brain needs to ‘put on hold’ in order to address more

immediate environmental demands.

The other frontal area we found to be thicker in TMD than controls was the vlPFC. This region

has been implicated in pain anticipation [752], pain modulation [909], and behavioural inhibition

[32]. There is evidence that patients with TMD are hypervigilant [435,808], which is related to

increased pain anticipation. Therefore, vlPFC thickening may be related to patients’

hypervigilance and increased anticipation to pain.

5.4.3. Relationship between chronic pain intensity and cortical thickness in M1 and aMCC

We found that gray matter in the orofacial region of M1 showed a negative correlation to pain

intensity. Several studies have reported abnormal stimulus-evoked fMRI activity in motor

regions of chronic pain patients, although the implication of these abnormalities is not often

156

discussed (see: [23]). One interpretation of our finding is that there is an adaptive mechanism

that responds to sustained, intense nociceptive activity by dampening motor output to prevent

further damage to the affected region. Evidence for this concept comes from studies of both

acute and chronic orofacial pains . For instance, TMD pain inhibits craniofacial motor function

[839], and dampens motor neurone output [556]. Similarly, acute pain stimuli applied to the

orofacial region can decrease M1 excitability [3,109]. Interestingly, Kirveskari et al. [492]

reported that patients with CRPS had weaker M1 reactivity with increased spontaneous pain

ratings, that is to say that there is a negative correlation between pain intensity and M1

activation. Another possibility is that patients with TMD make fewer jaw movements to avoid

eliciting pain. This reduced jaw activity may lead to atrophy of the orofacial region of the motor

cortex. Conversely, there is evidence to suggest that M1 may play a role in central processing of

pain [11,244,476,785] and pain modulation [189]. For instance, motor cortex stimulation has

been shown to have some, albeit limited, analgesic effect in chronic pain [342,689]. Therefore,

another interpretation of our finding is that M1 may be, in part, involved in descending

modulation of pain.

The other gray matter region that we identified that was negatively correlated to pain intensity is

the aMCC. This finding had also been identified in a previous study of gray matter in TMD

[963]. Experimental acute pain in healthy subjects is associated with activity in the MCC (caudal

BA 24) [690] and NS neurones have been identified in caudal BA 24 [440]. The top-down

descending antinociceptive system is thought to arise from the aMCC (possibly triggered by

activity in cingulate nociceptive neurones) [95]. The aMCC also shows increased rCBF during

analgesic thalamic stimulation [228]. Furthermore, this region shows opioid-related pain

modulation [342].

Therefore, both the M1 and the aMCC have been implicated in descending pain modulation

[689]. We found that patients with thicker cortex in both the aMCC and M1 report lower pain

intensity, whereas patients with thinner cortex in these regions report higher pain intensity.

Therefore, thicker cortex in these regions may provide a greater capacity to modulate TMD pain,

whereas thinner cortex may be related to an impairment of this descending modulation system.

157

5.4.4. Interaction between neuroticism and group in prefrontal gray matter

We investigated two factors that have not been previously examined in imaging studies of TMD:

neuroticism and TMD pain unpleasantness. Neuroticism is described as a personality trait

associated with heightened sensitivity and/or processing of negative affective stimuli

[178,179,907]. We found that in patients with TMD there is a positive correlation between

neuroticism and cortical thickness in the left ventromedial PFC (vmPFC; part of the OFC); an

area that normally shows a negative correlation between gray matter and neuroticism [954].

Neuroticism is considered to be a stable trait across the lifespan [178]. Therefore, our

observation highlights the contribution of neuroticism to abnormal gray matter in the OFC of

patients with TMD and, perhaps, to the development of TMD. An alternative interpretation is

that chronic pain may have modified the relationship between neuroticism and OFC thickness.

However, the lack of correlation of TMD duration and neuroticism scores in our patients with

TMD does not support this hypothesis.

We also found that unpleasantness is negatively correlated to gray matter in an adjacent region of

the OFC. The OFC has been implicated in cognitive reappraisal and emotional regulation [723]

related to interoception and somatoviscero stimuli, for directing our behaviour appropriately

[57,935]. The OFC is also thought to play a role in mental flexibility and adaptability (for a

critical review, see: [770]. Therefore, our finding suggests that patients with TMD may have

abnormal emotional regulation and reappraisal, which may lead to pain behaviours and

exaggerated affective responses to pain.

5.4.5. Study limitations

A few limitations in study design and results should be considered in interpretation of our data.

First, it is not known whether medications used by some of the patients have an effect on gray

matter. Second, some of our patients had co-morbid chronic pains (see Table 5-1), and these

pains (mostly related to TMD) may have contributed to our findings. Third, we cannot

disassociate an age effect from a pure TMD duration effect because of the strong correlation

between age and TMD duration (see Chapter 6). Furthermore, duration (and intensity) may not

fully describe the effect of pain on gray matter structure. Due to the nature of TMD pain, other

factors that we did not collect, such as the number of days of pain in a month may have more

158

explanatory variance. Our patients did not self-report any previous diagnosis of depression or

anxiety, although these exist in some patients with TMD [283,808,850]. Thus we cannot exclude

the contribution of these factors in white matter abnormalities in TMD in general. Finally, we

cannot totally rule out a contribution of ventricular volume to the VBM analysis of subcortical

structures. Despite these limitations, our study does provide considerable evidence in line with

previous findings in the literature about gray matter abnormalities in chronic pain. Further study

limitations are discussed in Section 8.6.

5.5. Conclusions

This study provides evidence for gray matter abnormalities in both the ascending pain and

descending antinociceptive systems as well as motor and cognitive areas in TMD. Further, we

have shown that the personality trait neuroticism and TMD intensity and duration can affect gray

matter, suggesting the presence of both personality-based and chronic pain-related abnormalities.

This structural imaging study identified striking abnormalities in patients with chronic idiopathic

TMD pain and highlights the contribution of both TMD-related and neuroticism-related factors.

Our key findings were that, compared to controls, patients with TMD had 1) cortical thickening

of the S1, frontal pole and vlPFC, 2) pain intensity-dependent cortical thinning in the aMCC and

M1 and pain unpleasantness-dependent cortical thinning in the OFC, 3) TMD duration-

dependent gray matter increase in the sensory thalamus, and 4) a positive correlation between

cortical thickness in the OFC and neuroticism, in contrast to a normally negative correlation.

159

Chapter 6 STUDY II: Age-related gray matter abnormalities in

temporomandibular disorder

This study has been published in Brain Research:

Moayedi M, Weissman-Fogel I, Salomons TV, Crawley AP, Goldberg MB, Freeman BV,

Tenenbaum HC, Davis KD. Abnormal gray matter aging in chronic pain patients. Brain

Research 2012; 1456: 82–93.

6.1. Introduction

There is no doubt that the brain undergoes structural changes as part of a normal aging process.

For example, normal aging is characterized by cortical gray matter atrophy

[84,99,372,585,627,824], although gains have also been reported in some brain areas [322,749].

Brain plasticity also occurs with dysfunction, injury, or specific disease. Chronic pain in

particular is associated with brain plasticity, including reduced gray matter in regions associated

with pain modulation and limbic function, such as the ACC, insula, dlPFC

[98,225,575,577,617,779,780,963], and gray matter increases in regions associated with pain

perception, such as S1 [209,577,617]. Additionally, some studies of chronic pain populations

have identified gray matter decreases in the basal ganglia [741,764-766,772,963] and pain-

related gray matter decreases in other motor regions, such as M1 (see Chapter 5). We have

previously reported that thalamic gray matter volume in patients with TMD is correlated to the

duration of their pain [617] (see Chapter 5). However, it is unclear whether these changes are

caused by the cumulative effect of TMD pain over time, or whether age contributes to the

observed abnormalities.

Chronic diseases such as pain may interact with normal aging processes. For example,

accelerated age-related whole brain gray matter atrophy has been reported in FMS [502] and

chronic back pain [27]. However, most aging studies in chronic pain have assessed global gray

matter and little is known about the interaction between chronic pain and age in gray matter

volume/thickness of specific brain areas.

Although the functional significance of gray matter changes is not understood, MRI and

histological studies indicate that changes can be stimulus-dependent, as has been demonstrated in

160

studies of learning [267], training [266], and repetitive noxious stimulation [852]. These data

suggest that the aforementioned findings of abnormal gray matter progression in chronic pain

may be related to pain duration. While it is plausible that age-related changes specific to chronic

pain are the product of cumulative pain exposure, this hypothesis has not been tested empirically.

Therefore, the aims of this study were to determine 1) whether chronic pain in TMD is associated

with abnormal gray matter aging in focal cortical regions associated with nociceptive processes,

and 2) the degree to which the cumulative effects of pain contributes to age effects.

Prolonged nociceptive activity may disrupt or even reverse normal gray matter atrophy in

nociceptive and motor regions, and increase rates of atrophy in pain-modulatory regions.

Therefore, we hypothesized that normal age-related gray matter changes would be enhanced in

brain regions implicated in pain perception (e.g., the thalamus, S1, posterior insula and MCC,

and suppressed in motor regions (e.g., the M1, PMC, SMA and the basal ganglia) and in regions

implicated in pain modulation (e.g., ACC and the anterior insula) of patients with chronic pain.

6.2. Methods

A detailed description of methods is provided in Chapter 4.

6.2.1. Subjects

Seventeen patients with non-traumatic TMD and 17 pain-free healthy subjects with no prior

history of chronic pain were recruited and provided informed written consent to procedures

approved by the local Research Ethics Boards. All subjects were right-handed. Structural MRI

data in this cohort unrelated to age have been presented [617] (See Chapter 5). Patients with

TMD were screened with the Mount Sinai Hospital Dental Clinic TMD diagnostic criteria, based

on the TMD-RDC [285]. Patient demographics and specific screening criteria have been

previously described elsewhere [617,928] (see Chapter 5 and Table 5-1). All patients reported

the number of years of TMD symptoms, herein described as TMD duration. Exclusion criteria

(for all study participants) included: a history of serious diseases (metabolic, rheumatoid,

vascular), concurrent craniofacial pain disorders, or any contraindication to MRI scanning (e.g.,

claustrophobia, metal).

161

6.2.2. Imaging & Analysis

Each study participant was placed in a 3T GE MRI (Signa HDx) system fitted with an eight-

channel phased array head coil, lying supine with foam padding around their head to reduce

movement. The 3D brain scan was acquired with a T1-weighted scan (IR-prep 3D-FSPGR with

128 axial slices, 0.94 x 0.94 x 1.5 mm3 voxels, 256 x 256 matrix size, field of view = 24 x 24

cm, one signal average, flip angle = 20°, TI = 300 ms, TE = 5 ms, TR = 12 ms).

6.2.3. Global age effects

To assess group differences in global age effects, we estimated whole brain gray matter volume

from tissue segmentation performed in the SPM5 software package

(http://www.fil.ion.ucl.ac.uk/spm/software/spm5/) running under Matlab v.7.0.4 (Mathworks).

We used a general linear model to test for age-by-group interactions for whole brain gray matter.

All statistical analyses were performed in SPSS v.19.0 (www.spss.com).

6.2.4. Age-by-group effects in gray matter

CTA

FreeSurfer software (http://surfer.nmr.mgh.harvard.edu) was used to assess cortical thickness.

Detailed methods are described elsewhere [198,316,317,319,320]. Briefly, T1-weighted scans

underwent intensity normalization, skull stripping and segmentation of each hemisphere based

on tissue type (gray matter, white matter, CSF). Within each hemisphere, the boundary between

each tissue type was modeled as a surface so that the distance between each surface could be

calculated at each point on the cortex. Each subject’s homologous gyri and sulci were aligned to

the standard average brain provided in FreeSurfer. A Gaussian spatial smoothing kernel of 6 mm

FWHM was applied to compensate for topographical heterogeneity prior to statistical analysis.

For this study, the distance between two points on the cortex (or two vertices) was 0.71 mm.

To restrict our search to regions hypothesized to be involved in TMD chronic pain, a single mask

using FreeSurfer’s cortical parcellations was constructed [321] that included the OFC, PFC,

insula, M1, SMA, cingulate cortex, postcentral gyrus and sulcus (including S1 and S2) and the

posterior parietal cortex [23]. Age-by-group interactions in cortical thickness were tested at each

vertex on the cortex within the cortical mask. To correct CTA results for multiple comparisons,

we used a Monte Carlo simulation with 1000 permutations with the AlphaSim software package

162

(http://afni.nimh.nih.gov/afni/doc/manual/AlphaSim), as done previously [98,847]. Thus, we

calculated a mask-wise corrected threshold of p < 0.05 from an uncorrected voxelwise p < 0.01

and a cluster threshold of 178 contiguous vertices (349 voxels).

VBM

VBM analysis was used to measure subcortical structures. One additional control subject was

recruited and consented to REB approved procedures to replace one subject subsequently

removed from the subcortical analysis because of poor segmentation in the pre-processing

pipeline.

Preprocessing of T1-weighted images and statistical analyses were performed in the SPM5

software package (http://www.fil.ion.ucl.ac.uk/spm/software/spm5/) running under Matlab

(Mathworks). We used the VBM 5.1 toolbox (http://dbm.neuro.uni-jena.de/vbm) implemented in

SPM 5 to perform VBM (for details see: [36]). Briefly, images were normalized to a standard

template (ICBM-152), tissue types were then segmented using a Markov random field model,

and underwent Jacobian modulation, followed by spatial smoothing of gray matter with a 10 mm

FWHM Gaussian kernel. We set an absolute threshold mask of 0.10 to restrict the results to gray

matter. All coordinates were converted from MNI space to Talairach space [844] using the

Lancaster transform [513] implemented in GingerALE v.2.0.4 (http://www.brainmap.org/ale).

A gray matter mask of subcortical regions was constructed using the WFU Pickatlas toolbox

(http://www.nitrc.org/projects/wfu_pickatlas). The mask was constructed using the “Sub-lobar”

label mask and the restricted by the “Gray Matter” label mask. The final mask included the

bilateral basal ganglia, amygdala and thalamus. Age-by-group interactions in gray matter volume

were tested for each voxel within the subcortical mask. All significant results are reported at a

voxelwise false discovery rate (FDR) [357] corrected p < 0.05, as implemented in SPM 5.

6.2.5. Contribution of TMD duration to age-related gray matter abnormalities

We examined the degree to which any of the observed age effects are attributed to cumulative

duration of chronic pain. To do this, we performed a forward model multiple linear regression,

which measures the degree to which one independent variable correlates to the dependent

variable. Additional independent variables are added to the equation and the degree to which the

163

independent variable predicts the dependent variable. Age and TMD duration were entered as

explanatory (independent) variables and each of the significant findings as dependent variables.

6.3. Results

6.3.1. Patient characteristics

The mean age of subjects in the patient group (mean ± SD: 33 ± 12 years) was not significantly

different than the control group (CTA analysis cohort: 33 ± 9.8 years, p = 0.94; VBM analysis

cohort: 32 ± 10.1 years, p = 0.81). The range of the controls ages was 20 to 50 years old, and the

age range for patients was 18-59 years old. The distribution of controls’ and patients’ ages is

presented in Table 6-1. Patient characteristic details are found in Table 5-1 [617,928]. Patients

reported having TMD for durations of 0.75- 30 years (mean ± SD: 9.8 ± 8.3 years). Of

importance for this study is that there was variability in the durations of TMD pain that

facilitated assessment of this factor, compared to age.

6.3.2. Global age effects

Both the control and TMD groups showed age-related whole brain gray matter atrophy, and there

were no significant group differences in whole brain gray matter volume (p = 0.88, see Figure 6-

1a). Specifically, there was a significant negative correlation between whole brain gray matter

and age for the controls (r = -0.55, p = 0.024, slope (m) = -3.46 cm3 /year) and for the patients ( r

= -0.72; p = 0.001, m = -5.11 cm3/year). However, there was an accelerated overall whole brain

aging effect in the patients with a significant group interaction of the slopes of the gray

matter/age curves (i.e., rate of change of gray matter with age) (p = 0.0002; see Figure 6-1b).

There was a significant correlation between patients’ duration of TMD and their age (r = 0.54, p

= 0.026; see Figure 6-1c), however TMD duration was not significantly correlated to gray matter

volume in the patient group (r = -0.37, p = 0.14, see Figure 6-1d).

6.3.3. Focal age effects

A significant age-by-group interaction (p < 0.05, corrected for multiple comparisons) was

localized to two focal regions within the right cortex. In one region, on the border of aMCC and

pgACC (BA32), patients had cortical thinning with age (r = -0.54, m = -0.022 mm/year),

whereas controls had age-related cortical thickening (r = 0.76, m = 0.033 mm/year). In another

164

region, the PMC, controls had age-related cortical thinning (r = -0.87, m = -0.035 mm/year),

whereas patients did not have normal atrophy, but rather had a very modest age-related cortical

thickening (r = 0.11, m = 0.002 mm/year) (see Table 6-2 and Figure 6-2 for details).

Three subcortical clusters were identified that had a significant age-by-group interaction (p <

0.05, FDR corrected; see Figure 6-3): the left PO thalamus and the right and left dorsal striatum.

The cluster in the right dorsal striatum extends to VPL, VPM and VL thalamus. Interestingly,

Figure 6-3 illustrates the strong positive correlation between gray matter volume with age in the

TMD group (r = 0.65) in contrast to the weak correlation in the control group in the thalamus (r =

-0.27). Figure 6-3 also illustrates the group differences in the dorsal striatum where the control

group showed a clear age-related gray matter loss (left: r = -0.78; right: r = -0.80), in contrast to

the age effect observed in the TMD group (left: r = 0.42, right: r = 0.33) (see Table 6-2, Figure 6-

3).

6.3.4. The contribution of TMD duration to gray matter age effects

A schematic of the focal aging effects in control and TMD groups is shown in Figure 6-4a (also

see discussion). The contribution of age and duration to the observed age-by-group interactions

was examined because of the observed correlation between age and duration. We performed a

multiple regression analysis with age and duration as independent variables to parse out the

relative contributions of these two factors. To describe our results, we used the standard

annotation for partial correlations, e.g., age • duration, such that age is being correlated to the

dependent variable while regressing out the variance related to duration. The outcomes of these

analyses are shown schematically in Figure 6-4b. In the right thalamus, the partial correlation

coefficient between age and gray matter volume was no longer significant when controlling for

duration (rage = 0.65, p = 0.005; rage • duration = 0.45, p = 0.082), whereas duration remained

significant when age was included in the model (rduration = 0.70, p = 0.002; rduration • age = 0.56 p =

0.026). In the aMCC/pgACC, the age-by-group interaction was driven by the shared variance

between age and duration. When the age was regressed out of the correlation between duration

and cortical thickness, the relationship remained insignificant (rduration = -0.45, p = 0.073 to rduration

• age= -0.22, p = 0.409). Further, when duration was regressed out of the correlation between age

and cortical thickness in the aMCC/pgACC, the relationship was no longer significant (rage = -

0.54, p = 0.027 and rage • duration = -0.39, p = 0.133) (see Table 6-3).

165

In some cases, duration did not contribute to the observed age-by-group interaction. For instance,

in the bilateral dorsal striatum, the partial correlations between age and gray matter in these

regions did not show much change when duration was regressed out. Similarly, when age was

regressed out of the correlation between duration and gray matter, the partial correlations were

not significant (see Table 6-3). In the case of the PMC, we found a suppressive effect [170], i.e.,

when duration was included in the regression model, age became a better predictor of the

interaction as the correlation became more significant. That is, the partial correlation coefficient

of age and thickness in the PMC, controlling for duration, was larger than the zero-order

correlation (zero-order correlation: 0.11, partial correlation: 0.38). Similarly, when correlating

thickness to TMD duration and controlling for the effect of age, the correlation coefficient for the

PMC decreased from -0.36 to -0.50 (see Table 6-3).

166

Table 6-1: Distribution of subject ages

Age range

18-20 21-25 26-30 31-35 36-40 41-45 46-50 51-55 56-60 TOTAL

#Con 1 5 2 2 3 2 2 0 0 17

#Pat 2 4 1 5 1 1 1 1 1 17

Abbreviations: #Con – Number of control subjects; #Pat – Number of patients with TMD

167

Table 6-2: Age-related group differences in cortical thickness and subcortical gray matter

volume. Shown are the statistically significant (p<0.05) group differences in correlation

coefficients of gray matter against age. Peak vertex/voxel Talairach coordinates (TAL) are

reported.

Interaction Region # Voxels Correlation

(age vs. thickness)

Peak TAL Peak

t-score

rpatients rcontrols X Y Z

Cortical

(CTA)

C > P R aMCC/pgACC 451 -0.54 0.76 14 30 20 4.42

C < P R PMC 423 0.11 -0.87 8 12 54 -5.18

Subcortical

(VBM)

C < P L dorsal Striatum 1537 0.42 -0.78 -18 10 10 -4.99

R dorsal Striatum 4415 0.33 -0.80 11 -3 18 -4.65

L Thalamus 264 0.65 -0.27 -21 -23 14 -3.36

Abbreviations: aMCC – anterior mid-cingulate cortex; pgACC – perigenual anterior cingulate cortex; CTA – Cortical thickness analysis; PMC – premotor cortex; VBM – voxel-based morphometry.

168

Table 6-3: Contributions of TMD duration to the age-gray matter relationships. Zero (0) order

correlations and partial correlations are presented for age and TMD duration versus gray matter

thickness/volume. The p-values in the age and duration columns represent the significance of the

correlations.

Significant group

differences

Age Duration

r p r age • duration p r p r duration • age p

R PMC 0.11 0.677 0.38 0.145 -0.36 0.159 -0.50 0.051

R aMCC/ pgACC -0.54 0.027 -0.39 0.133 -0.45 0.073 -0.22 0.409

L Dorsal Striatum 0.42 0.096 0.42 0.109 0.13 0.639 -0.12 0.650

R Dorsal Striatum 0.33 0.204 0.33 0.216 0.09 0.742 -0.10 0.706

R Thalamus 0.65 0.005 0.45 0.082 0.70 0.002 0.56 0.026

Abbreviations: aMCC – anterior mid-cingulate cortex; pgACC – perigenual anterior cingulate cortex; PMC – premotor cortex; S1 – primary somatosensory cortex

169

Figure 6-1: Total gray matter volume and correlations with age and duration: (a) Total gray

matter volume was not significantly different in the TMD group versus control group (p = 0.88).

(b) Significant gray matter volume decreases with age for both controls (3.46cm3/year) and TMD

170

group (5.11cm3/year) with a significant age-by-group interaction between the groups’ slopes (p =

0.0002). (c) TMD duration significantly correlated with patients’ age (r = 0.54, p = 0.026), (d)

but not to total gray matter volume (r = -0.37, p = 0.139). n.s. = not statistically significant

(p>0.05); asterisks (*) = p < 0.05; m = slope (rate of gray matter change in cm3/year).

R

171

Figure 6-2: Age-by-group interactions in cortical thickness. Patients with TMD had age-related

thinning (0.022 mm/year) in the anterior mid-cingulate cortex/perigenual anterior cingulate

cortex (aMCC/pACC), whereas controls show age-related thickening (0.033 mm/year) in this

region. In the premotor cortex (PMC), only the controls had age-related thinning (0.035

mm/year).

172

Figure 6-3: Group differences in age effects within subcortical gray matter volume. Age-by-

group interactions in the thalamus and dorsal striatum (p < 0.05, FDR). The graphs indicate gray

matter volume (corrected for total intracranial volume; TIV) versus age for each subject by

group. Colour bar shows t-statistic values.

173

Figure 6-4: Summary of age-related gray matter abnormalities (a) Schematic diagram of gray

matter regions with normal aging in healthy controls and abnormal aging in TMD. (b) Schematic

diagram summarizing the relative contribution of age and TMD duration to the regions of gray

matter with abnormal aging. The line thickness of each arrow depicts the relative contribution of

age and duration and the plus (+) and minus (-) signs depicts the direction of the relationship.

Abbreviations: aMCC/pACC – anterior mid-cingulate cortex/perigenual anterior cingulate

cortex; PMC – premotor cortex

174

6.4. Discussion

This study is the first to show that chronic pain associated with TMD in females is associated

with abnormal gray matter aging in focal cortical regions associated with pain and motor

processes. We found that patients with TMD have accelerated whole brain gray matter atrophy,

compared to pain-free controls. We also identified three types of aberrant patterns of gray matter

aging in five focal brain regions (Figure 6-4): 1) in the thalamus, patients with TMD gray matter

volume increased with age, whereas gray matter in controls was relatively sustained; 2) in the

aMCC/pgACC, patients with TMD had a progressive loss of cortical thickness with age, whereas

the controls had age-related cortical thickening; 3) in the dorsal striatum and PMC, there was

little change with age in the patients with TMD, whereas the controls had age-related atrophy.

TMD duration added to the age effects in the thalamus and the aMCC/pgACC, suppressed the

age effect in the PMC, but did not contribute to age effects in the dorsal striatum. Abnormal gray

matter aging in TMD may thus be due to the progressive impact of TMD-related factors as well

as inherent factors in individuals with this chronic pain.

6.4.1. Whole brain gray matter atrophy

Our finding of accelerated gray matter atrophy with age in TMD is consistent with previous

studies of gray matter in FMS [502] and chronic back pain [27]. These previous studies

attributed increased rate of gray matter loss to excitotoxicity and inflammatory molecules.

Specifically, they suggest that, because chronic pain is inherently harmful to the body and is

associated with negative affect and increased stress, the inflammatory response is upregulated

centrally, inducing cell death. These processes have been implicated in chronic age-related

diseases (Mattson, 2003), and therefore, it is plausible that they are implicated in age-related gray

matter loss in chronic pain states.

6.4.2. Use-dependent plasticity

We have demonstrated that TMD duration added to the age effects in the thalamus and cingulate

cortex. This form of plasticity is in line with the concept of use-dependent plasticity which

comprises structural and functional changes in the brain in response to increased or decreased

neuronal input [348,758]. For instance, studies in healthy human subjects have found that

training [266] and learning [267] can increase gray matter in the brain. Conversely, limb

175

amputation, and other forms of sensory loss can induce reorganization of the cortical map that

represents the affected limb [601,602] and gray matter loss [268,847]. Of particular interest are

studies that report reversible gray matter changes in nociceptive and antinociceptive regions of

the brain in response to repeated noxious stimulation in healthy subjects [93,852]. Furthermore,

patients with chronic pain show age-independent gray matter changes in both nociceptive and

pain-modulatory regions [98,356,558,575,617,741,772,963], some of which can be reversible

[384,741]. Our finding of progressive thinning in the cingulate cortex and increasing thalamic

gray matter are consistent with the concept of use-dependent plasticity. That is to say that

increased input, activity or use will increase gray matter, and decreased use or activity is

associated with decreased brain gray matter. Therefore, the observed age-by-group interaction in

the thalamus and aMCC/pgACC may, in part, be driven by prolonged nociceptive activity, which

could suppress the normal pattern of aging. These findings are consistent with our previous

report that gray matter in the thalamus is positively correlated with TMD duration, and may be

due to increased nociceptive activity, as previously discussed (see [617]).

In the aMCC/pgACC, we found that the shared variance of both age and TMD duration

contributes to progressive atrophy in the patients. This region receives orofacial nociceptive

input from thalamic nuclei that are part of the STT and TTT [189,276]. The aMCC/pgACC is a

complex, multimodal region that has been implicated in a number of functions [69]. For instance,

this region has been identified as a node in the salience network,!and has been implicated in

aspects of salience [213,229,262,263,774,849,927],!and pain

[220,230,231,259,265,440,508,528,632,933]. The cingulate cortex has also been implicated in

the cognitive and affective processing of pain [218,230,529,716,776,777,934,936], and the MCC

is involved in action selection and modulation of motor output in response to aversive stimuli

[790,898,904]. Therefore, it is possible that the age-related thinning in the MCC is related to

abnormalities in TMD with regard to cognitive and attentional processes related to prolonged

TMD pain [928].

The suppressive relationship identified in the PMC is of particular interest. The PMC is a region

that is often activated in neuroimaging studies of experimental pain [304,511]. In the current

study, age and TMD duration uniquely and non-redundantly predicted variance in gray matter.

When we included duration in the model, both variables (age and duration) better predicted the

progression of cortical thickness over time. Furthermore, age and duration had differential effects

176

on the PMC. Specifically, patients had sustained cortical thickness with age, whereas TMD

duration was related to cortical thinning in the PMC. The PMC receives nociceptive input from

the ventral caudal portion of the MD thalamus [276]. Therefore, we would expect that a barrage

of nociceptive input from the thalamus over an extended period of time could induce gray matter

plasticity in the PMC, as we have observed in the thalamus, rather than the observed

normalization. These paradoxical findings warrant further study to better understand the

relationship of the PMC and aging in the context of chronic pain. It is therefore apparent that

there is abnormal gray matter aging in TMD, independent of how long patients have had TMD.

The findings reported here are in contrast to previous findings that in some chronic pain

conditions central gray matter changes are related to pathology where there exists a peripheral

aetiology, such as osteoarthritis. It has been reported that once the peripheral cause of the pain

has been resolved, gray matter changes resolve [384,741]. There is evidence suggesting that both

central and peripheral mechanisms contribute to pain in TMD [561,757]. Therefore the source of

age-related variance may derive from multiple mechanisms in addition to (dis)use-dependent

plasticity. Interestingly, recent studies have reported that development of some brain regions is

tightly regulated by genes, rather than the environment [680,854]. There is also evidence

suggesting that there is a genetic predisposition to functional chronic pain syndromes [247-

249,809]. It is therefore feasible that there is a genetic contribution to the observed abnormal

aging effects. These findings provide a genetic source of plasticity [137,858] that may co-exist

with other forms of plasticity.

6.4.3. Cellular basis of changes in gray matter

The cellular and molecular basis of MRI-detectable gray matter changes remains to be explained.

However, several hypothetical mechanisms have been postulated, such as neuronal and/or glial

death [575], but recent evidence suggests that, to some extent, gray matter losses are likely

related to density of small dendritic spines [278,608], and remodeling of neuronal processes

[537]. Alternatively, reversible gray matter changes in chronic pain may be caused by

neuroinflammation [238,381,920], and induce MRI-detectable increases in gray matter. This

mechanism could explain both abnormal age-related increases and maintenance of gray matter

volume/thickness. For the observed gray matter losses, however, we cannot rule out that cell

death does not occurring in age-related gray matter losses – healthy populations lose neurones as

177

they age, and persons with neurodegenerative diseases suffer increased rates of atrophy related to

cell death.

6.4.4. Study limitations

One limitiation of this study is that it is not possible to control for the impact of long-term use of

medications in some of the patients on gray and white matter plasticity. Although the effect of

such medications is not well understood, recent evidence suggests that medications commonly

used to manage pain, such as anti-inflammatory medications and opiates, can impact brain

structure [916,962]. Considering that thirteen of seventeen of the patients included in this study

were taking NSAIDs, it is possible that the observed age-related findings (especially those of

gray matter maintenance) may be related to the effect of these drugs. Only one subject was

taking an opiate, hydromorphone, and so this is of less concern to this study.

Our patients did not self-report any previous diagnosis of depression or anxiety, although these

exist in some patients with TMD [283,808,850]. Thus we cannot exclude the contribution of

these factors in white matter abnormalities in TMD in general.

Another limitation of these findings is that it is a cross-sectional examination of structural

abnormalities in TMD. Therefore, the interpretations of “age-driven” and “TMD duration-

driven” abnormalities are limited to parsing out the relative variance that is accounted by each

factor. Therefore, the results warrant further study, and should be interpreted with caution.

6.5. Conclusion

In sum, our findings provide novel evidence that chronic TMD pain patients have abnormal age-

related gray matter changes in cognitive, motor and nociceptive brain regions. Our study

highlights the importance of understanding the effects of age and TMD duration in structural

studies of chronic pain, as progressive changes in gray matter may require differential

therapeutic approaches.

178

Chapter 7 STUDY III: White matter brain and trigeminal abnormalities in

TMD

The findings presented in this chapter are in press in Pain:

Moayedi M, Weissman-Fogel I, Salomons TV, Crawley AP, Goldberg MB, Freeman BV,

Tenenbaum HC, Davis KD. White matter brain and trigeminal nerve abnormalities in

temporomandibular disorder, Pain 2012; (in press).

7.1. Introduction

TMD comprise clinical problems involving the structures of and around the TMJ, the

masticatory musculature, or both [1,418]. TMD represent the most common orofacial chronic

pain disorder, prevalent in ~3-20% of the United States’ adult population [269,539,540,562].,

mostly in women [285,312,539,721]. In idiopathic TMD there is no clear aetiology

[283,286,287,653]. Central mechanisms are thought to initiate or maintain TMD pain [757]

based on TMD symptomology such as persistent pain, allodynia, and hyperalgesia, sometimes

extending to regions distant from the face [305,314,390,565,756,883], enhanced temporal

summation of pain to repetitive noxious heat stimuli [563], and dysfunctional DNIC

[113,491,756]. Abnormalities of these centrally-mediated processes suggest that ascending

nociceptive pathways and/or descending pain-modulatory pathways [515] are affected.

Additionally, patients with TMD can exhibit cognitive [370,379,380] and motor dysfunction

[837] possibly related to abnormalities in brain regions associated with these functions

[799,800,839,928].

Structural brain imaging provides an opportunity to delineate anatomical substrates of CNS

abnormalities in TMD. We reported that patients with TMD have increased cortical thickness in

the orofacial region of the primary somatosensory cortex (S1), the ventrolateral prefrontal cortex

(vlPFC) and the frontal pole [617]. Similarly, Younger and colleagues [963] reported increased

gray matter volume in motor, limbic and sensory regions, including trigeminal brainstem nuclei,

in myofascial TMD. These studies indicate that TMD patients have gray matter abnormalities in

sensory, motor and cognitive/limbic regions. Our finding of a correlation of thalamic gray matter

with TMD duration [617] supports the concept of long-term nociceptive-induced central

179

plasticity, possibly due to increased peripheral nociceptive activity, or dysfunctional

antinociceptive systems. Therefore, the former possibility suggests that although clinical

observations have not clearly identified gross peripheral abnormalities in TMD, the CNV may

indeed undergo changes due to abnormal persistent activity resulting in microstructural

abnormalities.

Previous studies examining white matter in other clinical conditions with sensory abnormalities

and/or chronic pain [156,356,558,847] found white matter abnormalities in brain areas involved

in sensory, modulatory and cognitive functions, and some of these white matter abnormalities

correlated with clinical findings. Therefore, studying the correlation between TMD

characteristics and measures of white matter integrity can provide insight into whether chronic

pain drives changes in white matter microstructure.

The main aim of this study was to evaluate the white matter abnormalities in TMD. Specifically,

we determined whether there are abnormalities along the CNVs or in white matter tracts in the

brain associated with sensory, cognitive or motor functions, compared to healthy controls.

Towards this goal, we used DTI, a brain imaging technique sensitive to the inherent diffusion of

water molecules that allows us to assess white matter microstructure. FA is a DTI metric that

reflects white matter tract integrity or structure [52,59,66,626]. We tested for abnormalities in FA

along the CNV of patients with TMD. Further, we used two approaches to assess white matter in

the CNS: First, we quantified brain FA abnormalities in TMD with TBSS [815], and second we

used probabilistic tractography [74] to determine the connectivity of main regions of white

matter disruption. Finally, we determined whether there is a link between white matter

microstructure and clinical characteristics of TMD.

7.2. Methods

A detailed description of methods is provided in Chapter 4.

7.2.1. Subjects

The study consisted of a cohort of 17 right-handed females with idiopathic TMD (mean age ±

SD: 33.1 ± 11.9 years) and 17 healthy right-handed females (mean age ± SD: 32.8 ± 9.8 years).

Informed written consent was obtained from all study participants for procedures approved by

the University Health Network and Mount Sinai Hospital Research Ethics Boards. Patients were

180

screened with the Mount Sinai Hospital Dental Clinic TMD diagnostic criteria, based on the

TMD-RDC [285]. Inclusion criteria included: 1) pain in the muscles of mastication rated

verbally as at least 4/10 for at least 3 months at the time of evaluation, or pain that is aggravated

by mandibular function; 2) moderate (on a scale of no (0), low (1), moderate (2), or severe (3))

pain to palpation and/or pain persisting post-examination in at least 3 muscle sites and/or

moderate pain to palpation of the TMJ and/or limited mandibular movement (opening < 40mm).

Patients were analgesic-free for 24 hours prior to scanning. Exclusion criteria for all subjects

included: 1) left-handedness; 2) self-report of metabolic, rheumatoid or vascular diseases or

disorders, or any other serious diseases; 3) self-report of commonly co-morbid functional chronic

pain disorders (IBS and FMS); 4) self-report of psychiatric disorders (e.g., depression,

schizophrenia, attention deficit hyperactivity disorder); and 5) self-report history of an abnormal

neurological examination; 6) contraindication to MRI scanning; 7) self-report of substance

abuse. Additionally, healthy controls were excluded if they had a history of chronic pain and

patients were excluded if they had a pain disorder other than TMD.

7.2.2. Questionnaires

Patients provided verbal ratings of their current pain and unpleasantness, as well as their average

pain intensity and unpleasantness ratings over the last month before scanning on a scale of 0 to

10 (0 = no pain, 10 = worst pain imaginable). Specifically, patients provided a numerical pain

score for pain intensity and pain unpleasantness by answering the following questions: 1) “Please

rate the intensity of your average pain over the last month, 0 being no pain and 10 being the

worst pain imaginable?” and 2) “Please rate the unpleasantness of your average pain over the last

month, 0 being not unpleasant and 10 being the most unpleasant pain imaginable?”. The patients

also reported the duration of their TMD symptoms.

7.2.3. Imaging parameters

All subjects underwent scanning in a 3-Tesla GE MRI system fitted with an eight-channel phased

array head coil. Each subject’s head was padded to reduce movement. The DTI acquisition

consisted of two runs of diffusion-weighted scans (TR = 14,500 ms, field of view: 24 cm x 24

cm, 128 x 128 matrix, 1.875mm x 1.875 mm in-plane resolution, 3 mm thick axial slices, with

ASSET with a factor of 2, maximum gradient strength = 40 mT/m, maximum slew rate = 150

T/m/s) acquired along 23 non-collinear, isotropic directions (b = 1,000 s/mm2). Additionally, 2

181

non-diffusion weighted scans (b = 0 s/mm2; b0) were acquired at the beginning of each run. The

scans covered from the top of the head to the second cervical spinal process (C2).

7.2.4. DTI pre-processing

The diffusion-weighted images were imported into FSL 4.1.8 (http://www.fmrib.ox.ac.uk/fsl) for

quality control [816]. Pre-processing included eddy current and motion artifact correction using

FDT [459]. Each subject’s two DTI runs were averaged to a single volume to achieve a greater

signal-to-noise ratio. Then, individual brain masks were created using BET [814]. These images

were then processed through two different pipelines for (1) voxelwise analysis and (2)

tractography.

The preprocessed images were fit with a diffusion tensor model using DTIFIT in FDT v. 2.0. We

then calculated voxelwise values of FA, and other DTI metrics (see below)

7.2.4.1. Other DTI metrics

To gain more insight into FA findings, we also assessed other DTI metrics in the clusters with

significant group difference. λ1 is thought to reflect diffusion along a tract, and reductions in this

value suggest disruptions along the tract diffusivity. RD is believed to reflect changes in

membrane permeability and, to some extent, myelination [66,67]. RD is calculated by averaging

the two radial vectors of the tensor model (λ2 and λ3). Finally, MD was measured because it is

associated to oedema and inflammation. Figure 7-1 provides a schematic of these DTI metrics.

DTI metrics (FA, MD, λ1, λ 2 and λ3) are calculated using DTIFIT in the FDT toolbox. The TBSS

skeleton is created based on FA, and peak FA values are projected onto the skeleton. For

statistical analysis of the other DTI metrics, MD, RD and λ1 values from the same voxels are

projected onto the skeleton. To test for group differences, we performed a multivariate analysis

of variance (MANOVA) with each of the DTI metrics as dependent variables, and group as an

independent variable.

7.2.5. Assessment of CNV

To assess the CNV, we manually drew a ROI (3.75 mm x 3.75 mm x 3 mm) in the axial plane on

each subject’s CNV root in the cisternal space in native space (see Figure 7-1). The nerves were

identified on color orientation maps of the primary eigenvector (λ1) image overlaid onto the FA

182

map. FA, MD, RD, λ1 values (see Figure 7-1) were calculated for each of the DTI volumes and

extracted from the ROI. Independent samples t-tests were used to assess group differences in FA

for each CNV. Furthermore, a paired t-test was utilized to compare left and right CNV FA values

within the patient group. Lastly, FA values derived from the each of patients’ nerve were

correlated with TMD characteristics (pain intensity, unpleasantness and duration). Statistical

significance was set at p < 0.05. All statistical tests were performed in SPSS v. 19.0

(http://www.spss.com/).

Figure 7-1: Schematic of diffusion tensor imaging, and summary of eigenvectors and metrics.

Lambda (λ) represents each of the eigenvectors that make up the tensor model. The equations to

calculate MD, fractional anisotropy (FA) and radial diffusivity (RD) from these eigenvectors are

also provided.

7.2.6. Tract-based spatial statistics

TBSS v.1.2 [815,817] was used to compare FA between the TMD and healthy control cohorts.

Briefly, FA maps underwent non-linear registration to a 1 x 1 x 1 mm FA map in standard space

(FMRIB58_FA, available in FSL), a mean image derived from all the subjects was created and

thinned to represent the centre of major white matter tracts common to all subjects, forming a

white matter skeleton. Each subject’s peak FA value perpendicular to the thinned track was then

projected onto the skeleton. One skeleton was created for analysis of group differences that

183

included all study participants, and a second patient only skeleton was created for correlation

analyses of TMD pain characteristics (see below). The mean skeleton images were set at a

threshold of 0.2 to include FA values that are related to white matter [815]. Other DTI metrics

were also evaluated to characterize FA findings (see Figure 7-1, and below). Additionally, we

performed a separate patient group analysis to assess which areas of the skeleton correlate with

TMD characteristics (pain intensity, unpleasantness and duration).

7.2.7. Probabilistic tractography

Probabilistic tractography was used to assess the connectivity of findings from the TBSS analysis

(http://www.fmrib.ox.ac.uk.fsl/fdt/index.html)[74,76]. First, we downsampled the preprocessed

diffusion-weighted images to create isotropic voxels (3 x 3 x 3 mm). The images were then

processed in FDT. Probability density functions (PDF) on up to two principal fibre directions

were estimated at each voxel in the brain. We then used multi-fibre tractography and drew 5000

samples from each seed voxel along the PDFs. The seed voxels were binarised images of the

significant clusters. Together, each seed voxel’s streamlines provide an estimate of its

connectivity. When a streamline reaches a voxel with more than one fibre direction, the

streamline follows the direction closest to the direction at which it arrives at the voxel. The

pathways generated by the algorithm represent the number of samples that have passed through a

voxel. To eliminate spurious connections, each computed pathway in each subject was

thresholded at 10 samples (of the 5000 generated from each seed voxel) that passed through the

voxels. These thresholded tracts were consistent between subjects. To visualize the findings,

each of the subject’s tracts were binarised and overlaid on a standard brain to produce a

probabilistic map of the pathways for controls and patients. The values in these maps reflect the

number of subjects who share a pathway.

7.2.8. Statistical analyses

7.2.8.1. Whole brain white matter

We assessed group differences in global (whole brain) FA and the white matter skeleton FA. To

do so, we extracted mean FA values across every voxel in the brain and every voxel in the

skeleton for each subject and performed a two-tailed t-test.

184

7.2.8.2. Mask analysis

A mask was created to restrict the analysis to a priori white matter ROI, including the white

matter regions containing the pathways subserving nociceptive, antinociceptive, motor and

cognitive functions. Thus, the mask included the brainstem white matter (including the white

matter that contains the TTT and the CST, TCT (including tracts to S1 and the anterior corona

radiata), the IC containing the CST, the cingulum, the rostrum, genu and the body of the corpus

callosum, the uncinate fasciculus, and the EC/ExC. The mask also included the thalamus, as

white matter courses through this region. The ROI were identified with the Johns Hopkins

University white matter Tractography Atlas, ICBM-DTI-81 white matter Atlas [625,910], the

Harvard-Oxford Cortical Atlas and the Harvard-Oxford Subcortical Atlas, available within FSL

(http://www.cma.mgh.harvard.edu/fsl_atlas.html). The final mask corresponded to voxels within

the FA skeleton.

A between group voxel-wise t-test within the skeleton mask was performed to identify regions

with significant group differences in FA, with age included as a covariate of no-interest. We used

two complementary statistical thresholding methods with 5000 permutations testing in FSL’s

randomise toolbox: (1) threshold-free cluster enhancement (TFCE; correct cluster p < 0.01)

[819], which is sensitive to spatially extensive areas of significant difference; and (2) cluster-

mass correction (p < 0.05), which requires clusters to meet a specific height threshold and,

therefore, is only sensitive to t-values above the threshold, which we set at t > 2.3.

An additional set of post-hoc of analyses was run to characterize regions with significant group

differences in FA. First, we determined whether the findings were related to specific

characteristics of TMD. To do so, we extracted each subject’s mean FA values from significant

group difference clusters. The patients’ FA values from each of these clusters then were

correlated with the TMD pain intensity, unpleasantness and duration in SPSS v. 19.0

(http://www.spss.com/). Statistical significance was set at p < 0.05. Second, we extracted the

mean values of mean diffusivity MD, RD and λ1 for each of the clusters and assessed whether

there were significant group differences between patients with TMD and controls using a

multivariate analysis of variance (MANOVA) with each of the DTI metrics as dependent

variables and group as an independent variable. Statistical significance was set at p < 0.05.

185

All significant TBSS results are thickened with the “tbss_fill” tool implemented in FSL for ease

of visualization. This tool uses a 3mm Gaussian smoothing kernel to thicken significant results to

the white matter tract (as defined by FA > 0.2).

7.2.8.3. Quantitative Tractography: connection probability

We identified that patients and controls had differential connectivity between the corpus

callosum seed and the dlPFC, and between the corpus callosum seed and the frontal pole.

However, these findings were only qualitative, and we therefore sought to quantify these group

differences. To do so, we performed a second tractographic analysis with specified targets based

on the observed differences. Specifically, we thresholded the tractography maps for each group

at 5 of 17 subjects (~30%) (see Figure 7-7A). Subsequently, we binarised the resultant image for

each group, and subtracted them one from another to determine the regions of difference between

the groups. We extracted the tracts for each hemisphere and restricted them to gray matter of the

target region. For example, the tract projecting to the frontal pole was restricted to the frontal

polar cortex. This was done by multiplying the binarised subtraction image to a binarised mask

of the frontal pole (from the Harvard-Oxford Cortical Structural Atlas;

http://www.cma.mgh.harvard.edu/fsl_atlas.html). Similarly, the tracts projecting to the dlPFC

were restricted to the middle frontal gyrus. We then performed tractography between the corpus

callosum seed and these 4 targets (the left and right frontal pole, and the left and right dlPFC).

Next, we extracted the number of samples that project to each target in each subject, and the

number of voxels in the seed that project to each of the targets. Group differences were tested

using two-tailed, independent samples t-tests. Significance was set at p < 0.05

7.3. Results

7.3.1. Patient characteristics

The ages of the patient and control groups were not significantly different (patients: mean age ±

SD: 33.1 ± 11.9 years; controls: 32.8 ± 9.8 years; p = 0.94). Details of the individual patient

demographics have been tabulated in our previous study of this cohort (see Table 5-1; and

[617]). Of importance for this study is that there was variability in the durations of TMD,

intensity and unpleasantness of pain that facilitated assessment of these factors in correlation

analyses. The duration of TMD ranged from 0.75 – 30 years (mean ± SD: 9.8 ± 8.2 years), while

pain intensity ranged from 2 – 7 (mean ± SD: 4.3 ± 1.8), and unpleasantness ranged from 1 – 8

186

(mean ± SD: 5.4 ± 2.1). Patients reported that their TMD pain was confined to the TMJ (n = 6),

the muscles of mastication (n = 3), or both the joint and muscles (n = 8). Most patients reported

bilateral pain (n = 13), but some reported unilateral pain (three right-sided, one left-sided). The

duration of TMD significantly correlated with the patients’ age (r = 0.536, p = 0.027).

7.3.2. Trigeminal nerve FA

We tested the CNV roots, in the cisternal space (see Figure 7-2), for group differences in FA,

MD, RD and λ1. We found that patients with TMD had significantly lower FA in both CNVs

(right and left: p < 0.001) (see Figure 7-3A). FA in the right CNV was also negatively correlated

with TMD duration (r = -0.53, p = 0.028) (See Figure 7-3B). The left CNV was not significantly

correlated with TMD characteristics. MD and RD were significantly higher in both CNVs (right:

p < 0.01; left: p < 0.05), compared to controls (See Figure 7-3C), but there were no significant

group differences in λ1.

7.3.3. Patients have lower white matter FA

To investigate global differences in white matter microstructure, we tested for group differences

in FA across the whole brain and within the white matter skeleton. Compared to controls, the

TMD patient group showed a 3.8% reduction in whole brain FA (mean ± SD: controls = 0.264 ±

0.012; patients = 0.254 ± 0.008; p = 0.003) and 3.6% reduction in whole brain skeletonised FA

(controls = 0.449 ± 0.018; patients = 0.433 ± 0.001; p = 0.006) (see Figures 7-4 and 7-5). We

also evaluated MD, RD and λ1 across the white matter skeleton and found that there was a

significant group difference (Pillai’s trace = 0.416; F(3,30) = 7.117; p< 0.001), and Bonferroni

post-hoc tests revealed that all three DTI metrics were significantly higher in the patient group

compared to the control group (all post-hoc tests: p < 0.05; see Table 7-1 for details).

The mask-wide TBSS analysis (p < 0.01; TFCE corrected) identified white matter clusters that

had 6.6-16.8% lower FA in the patient cohort compared to controls (see Table 7-2 and Figure 7-

5A). Two of these clusters were localized to the right IC, and two were in the right

external/extreme capsule adjacent to the insula. Other significant clusters were at the junction of

the right internal and external/extreme capsules, adjacent to the vlPFC, and in the white matter

adjacent to right S1 and M1. The largest cluster included the bilateral anterior body of the corpus

callosum, and extended to the left cingulum, the bilateral anterior corona radiata, the bilateral

187

ICs, the left external/extreme capsules, the fornix, the left and right thalamus and the brainstem.

Although the thalamus is mostly a gray matter structure, white matter does course through it and

we identified a cluster within the right thalamus with significantly lower FA in patients with

TMD, compared to controls. Individual clusters are shown in Figure 7-6.

Because one of our findings encompassed a large, diffuse region, we performed a secondary

analysis to test for more focal, highly significant group differences in FA. To do so, we used a

cluster-mass correction that requires significant clusters to reach a specified threshold, in this

case t(32)= 2.30. This analysis revealed a single focal cluster in the corpus callosum (peak voxel

MNI coordinates: -4, 26, 9, t(32) = 2.31, cluster size 353 voxels) (see Figure 7-5B).

We also determined whether the white matter tracts with lower FA in patients with TMD also

showed changes in MD, RD and λ1, indicative changes in axonal diffusion, permeability and

myelination. We found that there was a significant between-group difference in these

microstructural variable (Pillai’s trace = 0.958; F(7,26) = 6.182; p = 0.009). Bonferroni post-hoc

tests revealed that all of the clusters with lower FA showed significantly increased MD and RD

(p < 0.05), but there were no significant differences in λ1 (see Table 7-3).

7.3.4. Probabilistic tractography

Because the corpus callosum is a large structure that contains interhemispheric fibres we used

tractography to elucidate the connectivity of our finding of lower FA in the corpus callosum. We

evaluated the relationship between FA and connectivity by determining the connectivity of a

region with lower FA in patients. To do so, we used multi-fibre probabilistic tractography to

determine pathways passing through the corpus callosum finding from the cluster-mass

correction analysis (see Figure 7-5B and 7-7A). We first performed a qualitative analysis to

determine the connectivity of the corpus callosum finding within each group. As shown in Figure

7-7A, there were denser connections between the corpus callosum and the bilateral dlPFC in

controls, compared to patients. Conversely, connections between the corpus callosum and the

bilateral frontal pole were quite dense in patients, but sparse in controls. Quantification of these

connection probabilities verified that significantly fewer voxels from the corpus callosum

connect with the left and right frontal poles in controls, compared to patients with TMD (see

Figure 7-7B and C). We also found that patients with TMD have significantly higher number of

samples (i.e., connection probability) between the corpus callosum and the left frontal pole (p <

188

0.05). There were no significant group differences in the connection probabilities of the corpus

callosum and the right frontal pole. Additionally, we found that patients had significantly lower

connection probability between the corpus callosum and the right dlPFC (p < 0.05), but not the

left dlPFC (p > 0.05) (see Figure 7-7C).

7.3.5. White matter FA related to TMD pain characteristics

We tested whether each of the clusters with significant group differences in FA was correlated

with TMD characteristics (TMD pain intensity, unpleasantness and duration). We found that

three of these clusters showed significant FA-TMD correlations. Specifically, the FA within the

cluster at the junction of the internal and external/extreme capsules (adjacent to the vlPFC) was

significantly negatively correlated with TMD pain intensity (r = -0.49, p = 0.046) (see Figure 7-

8A). We also identified a significant negative correlation between FA in the thalamic cluster and

TMD pain intensity (r = -0.59, p = 0.013) (see Figure 7-8B). Finally, we identified a significant

negative correlation between a cluster in the right IC and both pain intensity (r = -0.72, p =

0.001) and pain unpleasantness (r = -0.63, p = 0.007) (See Figure 7-8C).

189

Table 7-1: Group differences in whole brain skeletonised white matter. Group means ± standard

deviations (SD) are presented for each group. Bonferroni post-hoc p-values are calculated based

on independent-samples t-tests.

Controls Patients

p-value

Mean

(x10-4 mm2/s) SD

Mean

(x10-4 mm2/s) SD

MD 7.3 0.16 7.6 0.12 < 0.05

RD 5.4 0.19 5.6 0.10 <0.05

λ1 11.2 0.22 11.4 0.23 <0.05

Abbreviations: MD – mean diffusivity; RD – radial diffusivity; λ1 – axial diffusivity

190

Table 7-2: White matter regions in patients with TMD with significantly lower fractional

anisotropy compared to controls. Mean and standard deviations of FA values, T-scores, percent

decrease and MNI coordinates are reported for the peak voxel.

Regions

FA

# Voxels T

% decrease

MNI

Controls TMD

x y z Mean St Dev Mean St Dev

Thalamus 0.371 0.0326 0.32 0.0002 305 4.08 13.8 17 -15 12

WM adjacent to S1/M1 0.428 0.0476 0.356 0.0012 129 3.51 16.8 32 -22 38

ICAL 0.564 0.0613 0.526 0.0046 3 3.06 -11.1 14 12 -2

EC/ExC 0.373 0.0335 0.339 0.0059 4 2.90 9.1 33 6 3

EC/ExC 0.399 0.0283 0.366 0.0073 3 2.82 8.3 32 9 1

Internal capsule 0.599 0.0454 0.558 0.0132 581 2.60 6.8 11 1 3

ICAL and EC/ExC 0.513 0.0501 0.456 0.0332 55 2.27 6.6 26 24 11

Diffuse1 0.853 0.0521 0.786 0.0018 10283 3.46 7.8 -5 26 10

1Diffuse cluster includes white matter in the corpus callosum, corticospinal tracts, internal

capsule, external/extreme capsules, fornix, cingulum, anterior corona radiata, cerebellar

peduncle, and pontine tracts.

All clusters are significant at p < 0.05, voxel-wise, mask-wide, FWE corrected after threshold-

free cluster enhancement. Voxels are 1mm3.

Abbreviations: IC – internal capsule; WM – white matter; S1 – primary somatosensory cortex;

M1 – primary motor cortex; ICAL – anterior limb of the internal capsule; EC/ExC –

external/extreme capsules.

191

Table 7-3: Group differences in mean diffusivity (MD), radial diffusivity (RD) and axial

diffusivity (λ1) in clusters with significant group differences in FA. Mean DTI metric values and

p-values are shown for two-tailed t-tests for MD, RD and λ1 values in each cluster from Table 7-2

and the cluster-mass corrected cluster. n.s. indicates that the t-test did not reach significance (p ≥

0.05).

MD (x104 mm2/s) RD (x104 mm2/s) λ1 (x103 mm2/s)

Controls TMD p Controls TMD p Controls TMD p

ICAL 8.0 8.5 <0.001 5.3 6.0 <0.001 1.3 1.3 n.s.

EC/ExC 7.5 7.9 <0.005 6.0 6.5 <0.001 1.0 1.1 n.s.

EC/ExC 7.6 8.0 <0.001 6.2 6.7 <0.001 1.0 1.1 n.s.

ICAL/EC/ExC 7.2 7.6 <0.01 5.0 5.4 <0.001 1.2 1.2 n.s.

S1/M1 6.7 6.9 <0.05 4.8 5.1 <0.001 1.1 1.0 n.s.

Thalamus 7.5 7.8 <0.001 6.0 6.4 <0.001 1.0 1.0 n.s.

IC 7.0 7.2 <0.005 4.1 4.5 <0.001 1.3 1.3 n.s.

Diffuse1 7.6 7.9 <0.001 4.8 5.3 <0.001 1.3 1.3 n.s.

Corpus callosum2 8.2 8.8 <0.05 3.8 4.6 <0.005 1.7 1.7 n.s.

1see Table 7-2 for description of the diffuse cluster.

2Corpus callosum cluster is from the cluster-mass correction.

Abbreviations: IC – internal capsule; ICAL- anterior limb of the internal capsule; EC/ExC –

external/extreme capsules; S1 – primary somatosensory cortex; M1 – primary motor cortex

192

Figure 7-2: Colour orientation maps of the trigeminal nerves. Axial images of three controls and

three patients with the primary eigenvector of the tensor model (λ1) within each voxel. The

eigenvector is colour-coded (green - anterior-posterior, red - left-right, blue - in the inferior-

superior plane) and modulated to the FA map. The location of the trigeminal nerves is indicated

by the yellow arrowheads.

193

Figure 7-3: Trigeminal nerve fractional anisotropy abnormalities in TMD. (A) The trigeminal

nerve roots (within the blue boxes) at the pontine level are shown on an axial slice from a high-

resolution T1- weighted MRI scan. The magnified view of the right nerve is from a diffusion-

weighted scan. The direction of the primary vector of the tensor model (λ1) within each voxel is

color-coded (green - anterior-posterior, red - left-right, blue - in the inferior-superior plane) and

194

the primary vector shown within each voxel. (B) patients with TMD have lower fractional

anisotropy in bilateral trigeminal nerves, compared to controls. (C) Fractional anisotropy is

negatively correlated with TMD duration. (D) Group differences in trigeminal nerve mean

diffusivity (MD), radial diffusivity (RD) and axial diffusivity (λ1). * = p < 0.05; ** = p < 0.01

*** = p < 0.005. CNV = trigeminal nerve.

195

Figure 7-4: Group differences in mean fractional anisotropy (FA) between patients with TMD

and controls. Patients had significantly lower whole brain FA (left panel) and skeleton FA (right

panel). Asterisks (*) indicate p < 0.05.

196

Figure 7-5: Group differences in fractional anisotropy between patients with TMD and controls.

(A) Blue regions indicate areas showing significant reduction in fractional anisotropy in TMD

compared to controls using threshold-free cluster enhancement (TFCE, corrected p < 0.01) and

overlaid on the white matter skeleton (shown in green) (significant clusters have been thickened

for enhanced visualization). Widespread abnormalities are observed in bilateral internal (IC) and

external/extreme capsules (EC/ExC), corpus callosum (CC), cingulum (Cing), thalamic and

brainstem white matter. (B) 3D rendering of the region of reduced FA in the corpus callosum of

patients compared to controls using cluster-mass thresholding (t > 2.3, corrected p < 0.05), the

cluster-mass corrected result in the corpus callosum (CC). A = anterior.

197

Figure 7-6: Clusters with significant group differences in fractional anisotropy. Blue clusters

(filled for visualization purposes) indicate regions of decreased FA in patients compared to

controls. Significant clusters (TFCE, p < 0.01) are displayed on separate brains in the (A)

anterior limb of the internal capsule (ICAL), (B and C) external/capsules (EC/ExC), (D) junction

of the ICAL and the EC/ExC, (E) white matter adjacent to the primary somatosensory and

primary motor cortices (S1/M1), (F) white matter in the thalamus, (G) internal capsule (IC) and

(H) in a diffuse cluster that includes brainstem white matter (cerebellar peduncles, pontine tracts,

corticospinal tracts), IC, EC/ExC, corpus callosum (CC), cingulum (Cing), the fornix and white

matter within the orbitofrontal cortex (OFC).

198

Figure 7-7: Abnormal white matter connectivity in TMD. Qualitative analysis of probabilistic

tractography of the cluster-mass corrected (t > 2.3, p < 0.05) cluster in the left corpus callosum

revealed that (A) this abnormal white matter region has different connections (yellow arrowhead)

199

in TMD and controls. The right panel shows a subtraction map that reveals that patients have

sparser connections between the corpus callosum seed and the frontal pole (blue box), and denser

connections to the dorsolateral prefrontal cortex (dlPFC; green box). The colour bar indicates the

number of subjects (between 5 and 17) contributing to the cluster at each voxel (controls are in

blue-light blue, and patients are in red-yellow). Quantitative tractography (B and C) revealed that

more voxels from the seed region (in green) in the corpus callosum project to the frontal pole in

the patients, compared to controls. The colour bar in (B) represents the proportion of subjects

with projections to the frontal pole in each voxel (controls are in blue-light blue, and patients are

in red-yellow). (C) Also, controls have a higher connection probability between the corpus

callosum and the right dlPFC, whereas patients have a higher probability of connection between

the corpus callosum and the left frontal pole. Graphs show mean number of samples (± SE) that

reach the target in each group (top panel), and the mean number of voxels (± SE) in the seed

mask that have samples that project to the target masks. * = p < 0.05.

200

Figure 7-8: Regions with group differences in fractional anisotropy are also correlated with

TMD characteristics: (A) the junction between the right internal and external/extreme capsules

(ICAL/EC/ExC) adjacent to the ventrolateral prefrontal cortex (vlPFC), (B) thalamic WM (white

matter) and the (C) right internal capsule are negatively correlated with TMD pain intensity. The

right internal capsule is also negatively correlated with TMD unpleasantness. Controls’ mean ±

SE FA for each region is shown in green. * = p < 0.05; ** = p < 0.01

201

7.4. Discussion

This study is the first report to show that both peripheral nerve and CNS white matter

abnormalities contribute to TMD pain. In support of a peripheral contribution, we found that

patients with TMD had lower FA in both CNVs and FA in the right CNV was negatively

correlated with TMD duration. In support of a central component to TMD, we found that patients

with TMD had (1) widespread abnormalities in the microstructure of white matter tracts related

to sensory, motor, cognitive and pain functions, including a focal area of the corpus callosum; (2)

stronger connectivity between the corpus callosum and the frontal pole, but sparser connectivity

with the dlPFC, compared to controls; (3) FA abnormalities that correlated with TMD

characteristics, and (4) reduced brain FA that were associated with greater MD and RD, markers

of inflammation and œdema.

7.4.1. Trigeminal nerve abnormalities

This is the first study to assess the CNV white matter integrity in TMD. We found that patients

with TMD had significantly lower FA and higher MD and RD in both CNVs. We also found that

the FA in the right CNV was negatively correlated with TMD duration. Previously, TMD was

considered idiopathic because there were no observable or major peripheral abnormalities in the

TM joint or muscles [757]. However, the DTI-based imaging technology used here clearly shows

microstructural abnormalities in the CNV. In line with peripheral abnormalities in TMD, one

study has demonstrated that the jaw-jerk reflex is abnormal on the painful side in TMD [332],

and another study found reduced cortical LEPs and reflexive masticatory muscular activity when

the CNV was stimulated in TMD, compared to controls [743]. Therefore, increased nociceptive

firing either from the periphery or from aberrant firing patterns (see below) over time could

affect the microstructure of the CNV and could also contribute to central abnormalities along the

ascending nociceptive system. This concept is supported by findings of increased pain sensitivity

within and outside the trigeminal region [313,491,563-565,841]. Alternatively, there is evidence

of abnormalities in orofacial motor function in TMD [46]. Recent evidence suggests that

nociceptive input can modulate motor output (see Section 2.2.3.2). Therefore, it is possible that

the microstructural abnormalities in the trigeminal nerve may be related to altered motor output

from the CNS.

202

7.4.2. Abnormal sensorimotor tracts in TMD

In addition to the peripheral nerve findings, we also found that patients have lower FA in the

brainstem, white matter coursing through the thalamus, the IC, and tracts adjacent to S1/M1.

Therefore, decreases in FA and increases in MD and RD along the ascending nociceptive

pathways could be induced or maintained by aberrant peripheral input from the CNV, or reduced

antinociceptive activity in the CNS. In support of this concept, are our previous study showing

that in patients with TMD there is a positive correlation between gray matter volume in the

thalamus and TMD duration, and thicker cortex in the orofacial region of S1 (See Chapter 5)

[617], suggesting that increased activity may lead to structural brain changes over time.

Given that DTI cannot distinguish between ascending and descending tracts [63], it is possible

that the abnormal white matter tracts contain descending projection fibres (the corticofugal

tracts) and corticothalamic tracts. The corticofugal tracts comprise the CST, corticobulbar,

corticoreticular, corticopontine tracts [625]. These tracts largely contain motor efferents and so

abnormalities in these tracts may underlie motor abnormalities in TMD, such as the masticatory

muscle hyperactivity, and abnormal jaw motor function under cortical control [837,839].

Furthermore, patients with TMD have been shown to have increased cortical activity in motor

regions and sluggish reaction times during a cognitive task [928].

7.4.3. Cognitive interference of pain and decreased modulation in TMD

We also identified lower FA in the cingulum bundle (adjacent to the mid-cingulate cortex), white

matter tracts coursing to the OFC and sgACC, tracts adjacent to the vlPFC, the anterior corona

radiata (in the ICAL, which projects to the prefrontal cortex) and the EC/ExC adjacent to the right

mid-insula. The so-called medial pain system is comprised of ascending nociceptive fibres that

project to the medial thalamus and further to the cingulate cortex, the insula and the PFC

[234,505,868]. It is believed that these regions contribute to the cognitive-affective dimension of

pain. The insula, mid-cingulate cortex and vlPFC have been implicated in pain perception,

negative emotion, and cognitive function

[23,94,108,116,183,221,220,222,227,304,447,752,790,898]. The OFC and the sgACC are

thought to contribute to the descending pain modulation [95]. Patients with TMD have

widespread pain [305,563,565,883], abnormal DNIC [113,491,756], and have been shown to

203

have sluggish reaction times to a cognitive task [928]. Therefore, the observed microstructural

abnormalities in both the ascending medial pain system and descending modulatory system

provide a plausible neural substrate for abnormal cognitive and anti-nociceptive function in

TMD.

7.4.4. Abnormal connectivity in TMD

Another main finding of this study is that we identified a significant decrease in FA in the corpus

callosum (identified using cluster-mass correction). This is a region that connects

interhemispheric prefrontal regions [966], and this portion of the corpus callosum is also

connected to the cingulate cortex. Similarly, a study examining white matter abnormalities in

complex regional pain syndrome reported decreased FA in a similar region of the corpus

callosum of patients, compared to controls [356].

In our study, probabilistic tractography revealed that patients had differential connectivity

between the corpus callosum and two regions that have been implicated in executive control and

pain perception and/or modulation: the frontal polar cortex and the dlPFC

[19,47,105,144,173,403,404,457,496,497,503,552,620,834,875,876,908,922]. In line with these

findings, we have previously identified structural and functional abnormalities in these regions

[617,928]. Further, studies have shown that pain interferes with cognitive processes [529]. Given

the possibility that cognitive resources are limited, these competing demands can interfere with

one another, and impede cognitive performance. Evidence for the cognitive interference of pain

comes from studies that have shown that acute pain can modulate performance in a cognitive

task [291,529,687,697]. Further evidence comes from studies that have reported that chronic pain

patients have slower reaction times in cognitive tasks [370,379,526,850,928]. Within the context

of these findings, our results may represent that FA differences in the corpus callosum are related

to differential connectivity in the brain, and potentially related to interaction between executive

cognitive function and pain in TMD.

7.4.5. Intense and prolonged TMD may drive plasticity in white matter

We have previously reported that both pre-existing and chronic pain-driven factors contribute to

gray and white matter brain abnormalities in patients with IBS (see: [98,156]. In the current

study, the intensity and unpleasantness of TMD pain also correlated with regions with lower FA

204

in patients. There are two possible explanations of the finding of reduced FA correlated with pain

intensity. Firstly, it is conceivable that abnormalities in white matter existed prior to the onset of

TMD and, therefore, could influence the degree of TMD pain. Secondly, more intense,

unpleasant or longer TMD pain may induce white matter plasticity, reflected as changes in FA.

Evidence that pain induces or is associated with structural brain plasticity comes from recent

studies that have reported that gray matter abnormalities in chronic pain are reversible, once the

source of the pain has been resolved [384,741,782], which suggests that pain can potentially

induce gray matter plasiticity. In this study, decreased FA in the right CNV was related to the

duration of TMD, which provides support for the latter possibility – that is, pain-induced

plasticity. Similarly, the relationship between regions with abnormal FA in the brain (i.e., tracts

adjacent to the vlPFC, within the IC and coursing through the thalamus) and TMD pain intensity

suggest that TMD pain may be driving the decrease in FA, and the observed group difference. As

mentioned above, these findings suggest that increased peripheral firing or abnormal firing in the

CNV of patients with TMD or decreased dysfunctional central antinociceptive systems may alter

its microstructure over time. This represents a particularly novel finding given the emphasis of a

central ætiology for TMD [757].

7.4.6. Cellular basis of changes in FA

The diffuse pattern of white matter abnormality is likely due to the multidimensional nature of

chronic pain that impacts several anatomical pathways. However, our diffuse white matter

findings need to be interpreted with caution due to the global difference in FA we identified

between patients and controls.

The factors that contribute to reduced FA may be macrostructural, such as increased branching,

more crossing fibres or larger tracts (more axons) and/or microstructural changes such as cell

swelling (œdema), changes to protein filaments (neurofilament phosphorylation), disruptions to

the cell membranes, and, to a certain extent, decreased myelin [66,67]. Recent studies have

suggested that chronic pain may induce neuroinflammation in the brain [238,381,920], which can

lead to changes in both gray matter and white matter that can be detected by MRI. Therefore, the

observed changes in white matter may, in fact, be related to neuroinflammation rather than

increases or decreases in the number or size of cells and/or axons present in the fields being

studied. Our study has examined other measures of white matter microstructure (MD, RD and

205

λ1) to better characterize group differences in FA. The neuroinflammation model is supported by

our finding that regions with lower FA in TMD have higher MD and RD, which are markers of

inflammation and/or œdema [63,66].

7.4.7. Study limitations

The patients in this study patients did not self-report any previous diagnosis of depression or

anxiety, although these exist in some patients with TMD [283,808,850]. Thus, the contribution of

these factors in white matter abnormalities in TMD in general cannot be excluded. Another

limitation of the current study is that microstructural abnormalities in white matter tracts cannot

definitively be attributed to functional abnormalities, as this study only investigates structural

abnormalities. For example, the CNV is a mixed cranial nerve, with motor and sensory fibres. It

was not possible to disentangle whether the structural abnormalities are related to sensory

abnormalities, motor abnormalities, or both. The resolution of DTI analysis based on voxels that

are much larger than the width of a single axon. The current methods allow generalized

inferences to be made only about the local microstructure within a voxel, but cannot differentiate

which tracts are affected.

7.5. Conclusion

This is the first study to identify abnormalities in the CNVs of patients with TMD, suggesting a

peripheral contribution to TMD. We have also demonstrated widespread white matter brain

abnormalities in patients with TMD in sensory, motor and cognitive-affective regions, consistent

with a central component to TMD. Finally, we have identified that patients with TMD have

abnormal connectivity of cognitive-affective regions of the brain. In sum, we provide structural

findings consistent with observed behavioural and functional finding in TMD, supporting a

model of central abnormalities. while also present novel evidence of a potential peripheral

contribution to TMD.

206

Chapter 8 General Discussion

This is the first comprehensive study of both gray and white matter structural abnormalities in

idiopathic TMD, and the first to investigate focal age-related gray matter abnormalities in

chronic pain. The main findings of this thesis are that patients with TMD show:

1) Abnormal gray matter in brain areas implicated in somatosensory, motor, cognitive and

modulatory functions, including a) increased cortical thickness in the S1 and frontal

cortex, b) chronic pain intensity or unpleasantness-dependent cortical thinning in the

cingulate, motor, and orbitofrontal cortices, c) pain chronicity-dependent thalamic gray

matter increases, d) abnormal impact of neuroticism on orbitofrontal gray matter;

2) Abnormal gray matter aging in focal brain regions of pain and motor systems, some of

which is impacted by chronic pain chronicity;

3) Widespread brain white matter abnormalities, including a) in tracts containing

connections for sensory, motor, cognitive and pain functions, b) prominently stronger

corpus callosum-frontal pole connectivity and weaker dlPFC connectivity, c) chronic

pain/unpleasantness-dependent abnormalities, d) abnormally high MD and RD levels,

markers of brain neuroinflammation; and

4) Trigeminal nerve abnormalities, exacerbated by TMD chronicity.

Together, these findings suggest that idiopathic TMD are complex, and are accompanied by both

central and peripheral abnormalities, some of which may be pre-existing abnormalities that could

predispose persons to developing TMD.

Several factors can contribute to the observed structural abnormalities in the brains of patients

with TMD. The first possibility is that these changes are “all-or-none”, where having TMD is

accompanied with these brain abnormalities. Another possibility is that the observed changes

occur over time, the prolonged effect of pain progressively affects brain structure. Another

possibility is also related to duration, but different in that the magnitude of pain (intensity or

unpleasantness) can affect the magnitude of change in the brain: higher pain intensity drives

changes in the structure of the brain. Together, these possibilities may be referred to as pain-

207

driven abnormalities. Finally, in contrast to pain-driven abnormalities, there could be pre-

existing brain abnormalities, possibly related to stable personality traits that can predispose

persons to developing TMD. The relative contribution of these factors are discussed below.

8.1. Pain-driven abnormalities along nociceptive pathways

In this thesis, structural abnormalities were identified at several locations along the ascending

nociceptive pathway. This pathway begins with the CNV in the periphery. Nociceptive

information from the TMJ and the muscles of mastication is carried to the CNS via the CNV.

Nociceptive C- and Aδ-fibres project to second-order NS and WDR neurones in the VBSNC

[784]. These cells ascend to various brainstem nuclei, and the thalamus via the TTT, where they

project to third-order neurones, which form thalamocortical tracts [784]. This thesis has reported

decreased FA, and increased MD, and RD along the CNV of patients with idiopathic TMD.

These data suggest that there is increased permeability in the membrane or decreased myelin

insulation [67] (see below). Further, these abnormalities are negatively correlated to TMD

duration in the right CNV– patients with relatively shorter TMD durations have relatively normal

FA values, and these values deviate from normal values as TMD duration increases. These data

suggest that chronicity of TMD pain contributes to changes in the microstructure of the nerve.

Therefore, there may be increased nociceptive input from the periphery, or

decreased/dysfunctional central antinociceptive systems, or progressive abnormalities along the

nerve root, which may cause aberrant firing patterns. In line with the revised pain adaptation

model [431], changes in the recruitment of motor units, over time, can themselves become a

source of pain. Therefore, it is possible that increased nociceptive activity over time alters facial

motor activity. Prolonged alterations in motor activity from the orofacial region can itself

become pathological and a source of nociceptive input. This prolonged input, in turn, may alter

CNV microstructure and structures along the ascending nociceptive pathways.

In support of this hypothesis is the finding that gray matter volume in the thalamus positively

correlated to TMD duration. Similar to our finding in the CNV, with increased TMD chronicity,

patients showed increased gray matter in the thalamus. The age-by-group study confirmed the

finding that thalamic gray matter changes are driven by duration. It is therefore likely that

ascending nociceptive pathways undergo progressive change with long-term TMD. Similarly,

several studies reported pain-driven gray matter structural abnormalities in chronic pain

208

[27,98,206,268,303,356,490,502,740,764-766,772,888,889,919,963] (see Table 2-3). Support for

the concept of chronic pain-driven gray matter changes comes from a study by Teutsch and

colleagues [852], who reported increased S1 and MCC gray matter volume after eight

consecutive days of 20-minute paradigm of acute nociceptive stimulation to the volar forearm.

Subjects had thicker S1 and MCC. Additionally, this thesis reports abnormalities along white

matter tracts connecting the various levels of the ascending trigeminal nociceptive system – the

TTT, white matter coursing through the thalamus, TCT coursing through the somatomotor region

of the IC and adjacent to the somatosensory and motor cortices. Furthermore, the white matter

tracts within the thalamus and the motor region of the IC are related to pain intensity and/or

unpleasantness. For a summary schematic of abnormalities along the ascending sensory (and

nociceptive) pathway, see Figure 8-1.

In sum, this thesis has demonstrated structural abnormalities at several points along the

ascending nociceptive pathway – including key gray matter regions implicated in nociception,

and the tracts that connect them.

209

Figure 8-1: Proposed model of neural basis of sensory abnormalities in TMD. Altered

nocicieptive activity in the periphery may be related to decreased FA in CNV, and TTT in the

brainstem. This altered activity also leads to progressive increases in thalamic GM volume,

reduced FA in the TCT and increased cortical thickness in the orofacial region of S1. Brain

outline modified from the PHT500Y laboratory manual, University of Toronto.

8.2. Abnormalities in motor regions

The pain and motor systems interact at several levels within the nervous system (see Section

2.1.6.). This thesis reports pain-related structural abnormalities in gray matter and white matter

associated to motor function in patients with TMD. Specifically, the thickness of the orofacial

region of M1 was negatively correlated to pain intensity in patients with TMD. Similarly,

patients with TMD had lower FA in white matter tracts associated to motor regions, and FA in

these regions was negatively correlated to pain intensity. Furthermore, our lab has previously

reported that the same group of patients with TMD had increased cortical activity in motor

regions and sluggish reaction times during a cognitive task [928]. Finally, the age-by-group

analysis revealed abnormal aging of the PMC and the dorsal striatum in TMD, compared to

healthy controls.

210

There are three possible mechanisms that may explain the observed abnormalities in motor

regions and tracts (see Figure 8-2): (1) nociceptive input dampens motor output to prevent

further damage; (2) motor cortex modulates pain intensity; or (3) reduced movement to avoid

pain may affect motor cortex thickness. Future work should specifically address the pain-motor

interaction in TMD to determine which of these mechanisms is occurring in TMD.

Previous studies have reported that nociceptive input inhibited motor cortex excitability

[3,157,301,302,638,856], and that motor cortex stimulation modulated the perceived intensity of

pain in chronic pain [103,117,118,301,331,342,344,345,535,548,657,689]. Furthermore, a study

by Boudreau and colleagues [109] reported that task training-related motor cortex plasticity was

inhibited by noxious stimulation of the intraoral cavity. Additionally, several studies have

reported abnormal stimulus-evoked fMRI activity in motor regions of chronic pain patients,

although the implication of these abnormalities is not often discussed (see: [23]). For instance,

Kirveskari et al. [492] reported that patients with CRPS had weaker M1 reactivity with increased

spontaneous pain ratings, and a negative correlation between pain intensity and M1 activation.

Therefore, studies in experimental pain paradigms and in chronic pain patients have

demonstrated that nociceptive input modulates M1 function. Furthermore, evidence suggests that

M1 activity can modulate pain perception.

Alternatively, the observed relationship between pain intensity and M1 thickness may be related

to the possibility that patients with TMD make fewer jaw movements to avoid eliciting pain.

Based on the concepts of use-dependent plasticity (or in this case disuse-dependent plasticity),

this reduction in jaw activity may lead to atrophy of the orofacial region of the motor cortex

[576].

211

Figure 8-2: Proposed model for the neural basis of motor abnormalities in TMD. Altered

nociceptive input inhibits motor cortex output in an intensity dependent manner, and reduces FA

in corticocortical tracts connecting S1 and M1. Decreased motor output causes disuse-related

plasticity along corticofugal tracts (including corticobulbar tracts) and the motor portion of the

CNV.

8.3. Abnormalities in cognitive-modulatory brain regions

This thesis has demonstrated structural abnormalities in brain areas that have been implicated in

cognitive/modulatory functions, providing a structural framework that could underlie the

negative impact of chronic pain on cognitive abilities, including concentration, learning,

memory, and attentional focus [26,250,275,365,369,455,489,666,928]. Specifically, we have

reported thicker cortex in the vlPFC. Furthermore, patients with TMD had reduced measures of

white matter integrity in white matter adjacent to the vlPFC. The vlPFC has been implicated in

various aspects of pain, including anticipation [752] and modulation [909], as well as

behavioural inhibition [32]. Furthermore, studies have reported hypervigilance in patients with

TMD [435,808], which is related to increased pain anticipation. Therefore, the observed

structural abnormalities in the vlPFC may be related to patients’ hypervigilance and increased

anticipation to pain (see Figure 8-3).

212

Another key finding of this thesis was that the left frontal polar cortex was thicker in patients

with TMD, than in controls. Additionally, probabilistic tractography revealed that the frontal

polar cortices were more connected interhemispherically – the right and left frontal poles were

more connected in TMD than in pain free controls. The frontal pole can be activated during

spontaneous pain in patients with chronic pain [47]. Additionally, the frontal pole has also been

implicated in a number of complex executive cognitive functions such as learning behavioural

routines [457,496,834], cognitive branching (the ability to put a pending task on hold to execute

an ongoing one) [497], behavioural flexibility/adaptability [105] and post-hoc monitoring or

evaluating decisions based on feedback [875]. The concept of cognitive branching would suggest

that there is a limited resource of working memory, and that tasks that are more behaviourally-

relevant (or salient) are prioritized [44,45,155,495,497,498]. Specifically, the frontal polar cortex

monitors the relevant salience (or value) of a given stimulus and select behavioural outcomes

accordingly. Given that pain is an inherently salient stimulus [265,530,933], it is conceivable that

pain engages the frontal pole. Further, a study by Harman and Ruyak [409] reported that people

with persistent low-level pain related to their work habits had cognitive deficits, indicative of an

interference effect of pain. Therefore, the frontal pole may process the cognitive dimension of

pain, which suggests that pain has a cognitive load. In the short term, as in cases of acute pain,

the frontal pole may direct behaviour to protect the subject. However, in chronic pain conditions,

this region may be continuously engaged. Therefore, when a person with chronic pain is

distracted, or performing a demanding cognitive task, the frontal pole may be “putting the pain

on hold”, allocating resources to the other, competing cognitive demands (see Figure 8-3).

Accordingly, Eccleston and colleagues [289] showed that chronic pain patients could briefly be

disengaged from pain during attentional switches. Further support for this hypothesis comes from

studies that demonstrate that distraction can reduce pain perception

[123,430,584,588,687,864,888,891,935,955]. These studies have demonstrated that distraction-

based modulation of pain requires the recruitment of antinociceptive brain regions. However,

these systems may not be effective in patients with TMD in two ways: (1) there is evidence of

dysfunctional antinociceptive systems (see Section 2.2.3.1) [305,491], and (2) the cognitive

branching model suggests that cognitive resources are limited – there are only so many tasks that

can be juggled [497]. Therefore, the cognitive load of chronic pain could impact the a patient’s

ability to properly complete tasks [529]. In support of this idea, we have recently shown that

patients with TMD have sluggish reaction times to low and high conflict cognitive interference

213

tasks, but not to a simple sensorimotor task [928]. Therefore, in the context of use-dependent

plasticity [576], one possible explanation for the observed thickening of the frontal pole and the

increased white matter connectivity between the bilateral frontal poles is that chronic TMD pain

bears a cognitive load that the brain needs to ‘put on hold’ in order to address more immediate

environmental demands.

Figure 8-3: Proposed model of the neural basis for cognitive abnormalities in TMD. Altered

nociceptive activity alters the structure and connectivity of cognitive brain regions, including the

vlPFC and the frontal pole.

Additionally, this thesis found that patients with TMD have abnormal gray matter aging in the

aMCC/pgACC. Furthermore, these patients have lower FA in the cingulum bundle (adjacent to

the MCC). This region receives orofacial nociceptive input from thalamic nuclei that are part of

the STT and TTT systems [189,276]. The aMCC/pgACC is a complex, multimodal region that

has been implicated in a number of functions [69]. For instance, this region has been identified as

214

a node in the salience network,!and has been implicated in aspects of salience

[213,229,262,263,774,849,927],!and pain [230,231,259,265,440,508,528,632,933]. Furthermore,

the cingulate cortex has also been implicated in the cognitive and affective processing of pain

[218,230,529,716,776,777,934,936], and the MCC is involved in action selection and modulation

of motor output in response to aversive stimuli [790,898,904]. Therefore, it is possible that these

structural abnormalities in the cingulate cortex are related to cognitive and attentional

abnormalities in TMD.

Patients with TMD also had widespread white matter abnormalities in white matter tracts related

to pain-modulatory brain regions. Specifically, patients had lower FA and increased MD and RD

in white matter tracts coursing to the OFC and sgACC, the anterior corona radiata (in the ICAL,

which projects to the prefrontal cortex) and the external/extreme capsule adjacent to the right

mid-insula. The insula, mid-cingulate cortex have been implicated in pain perception, negative

emotion, and cognitive function [23,94,108,116,183,222,227,304,447,752,790,898]. The OFC

and sgACC are thought to contribute to descending pain modulation [95]. Furthermore, patients

with TMD had abnormal connectivity to the dlPFC, a region implicated in pain modulation

[552], selecting information about the sensory environment (attention, self-monitoring,

personality, working memory) and decision-making [54,700] (see Figure 8-4). Patients with

TMD have widespread acute pain hypersensitivity [305,563,565,883], abnormal DNIC

[113,491,756], while having sluggish reaction times to a cognitive task [928].

In sum, this thesis provides a potential structural neural basis for cognitive deficits and

dysfunctional pain-modulatory circuits in patients with idiopathic TMD.

8.4. Neuroticism and TMD

It is possible that some factors predispose persons to develop TMD. A recent prospective study

by Fillingim and colleagues [315] reported that certain psychological traits could, to some extent,

predict TMD onset, including neuroticism. In the current thesis, we report that in patients with

TMD there is a positive correlation between neuroticism and cortical thickness in the left vmPFC

(part of the OFC); an area that normally shows a negative correlation between gray matter and

neuroticism [954]. Neuroticism has been defined as a stable personality trait associated with

heightened sensitivity and/or processing of negative affective stimuli [178,179,907]. Therefore, it

is not surprising that patients with high neuroticism scores have lower pain threshold and pain

215

tolerance, and are more distressed by pain in experimental settings and chronic pain [408].

Neuroticism has been associated with pain-related suffering [35,408], pain sensitivity

[177,373,907], nerve injury outcomes and the development of neuropathic pain [848], and TMD

[315]. However, the mechanism by which inherent personality-related factors reduce the brain’s

capacity to effectively modulate nociceptive input and pain perception, remains unclear.

Furthermore, not all chronic pain patients have high neuroticism scores, and not all persons with

neuroticism have chronic pain [179]. Therefore, neuroticism alone is not sufficient to develop

chronic pain, however it can contribute to the severity of pain in chronic pain patients. Therefore,

our observation highlights the contribution of neuroticism to abnormal gray matter in the OFC of

patients with TMD and, perhaps, to the development of TMD (see Figure 8-4). Similarly, in IBS

patients, neuroticism correlated to FA in the left MD thalamus, but not in controls [156]. An

alternative interpretation is that chronic pain may have modified the relationship between

neuroticism and OFC thickness. However, the lack of correlation of TMD duration and

neuroticism scores in our patients with TMD does not support this hypothesis.

We also found that unpleasantness is negatively correlated to gray matter in an adjacent region of

the OFC. The OFC has been implicated in cognitive reappraisal and emotional regulation [723]

related to interoception and somatovisceral stimuli, for directing our behaviour appropriately

[57,935]. The OFC is also thought to play a role in mental flexibility and adaptability (for a

critical review, see: [770]. Therefore, our finding suggests that patients with TMD may have

abnormal emotional regulation and reappraisal, which may lead to pain behaviours and

exaggerated affective responses to pain.

216

Figure 8-4: Proposed model for the neural basis of modulatory system abnormalities in TMD.

The relationship of neuroticism and cortical thickness in the OFC alters the structure and

antinociceptive function in pain modulatory brain regions, including the sgACC. The dlPFC is

also has reduced white matter connectivity in patients with idiopathic TMD, which may be

another substrate for dysfunctional pain modulation networks in TMD.

8.5. Basis of structural brain abnormalities

The cellular and molecular bases of MRI-detectable gray and white changes in chronic pain are

not known but there is now evidence to support several hypotheses. Mechanisms of gray matter

change include neuronal and/or glial death [575], but recent evidence suggests that, to some

extent, gray matter losses are likely related to density of small dendritic spines [278,608], and the

remodeling of neuronal processes [537]. Alternatively, reversible gray matter changes in chronic

pain may be caused by neuroinflammation [238,381,920], and induce MRI-detectable increases

in gray matter. This mechanism could explain abnormal thicknening (or maintenance) of gray

matter volume/thickness. For the observed gray matter losses, however, we cannot rule out that

cell death is not occurring – healthy populations lose neurones as they age, and persons with

neurodegenerative diseases suffer increased rates of atrophy related to cell death.

217

The factors that contribute to reduced FA may be macrostructural, such as increased branching,

more crossing fibres or larger tracts (more axons) and/or microstructural changes such as cell

swelling (œdema), changes to protein filaments (neurofilament phosphorylation), disruptions to

the cell membranes, and, to a certain extent, decreased myelin [66,67]. Furthermore, the

neuroinflammation hypothesis for gray matter changes can also apply to white matter changes.

Therefore, the observed changes in white matter may, in fact, be related to neuroinflammation

rather than increases or decreases in the number or size of cells and/or axons present in the fields

being studied. This thesis has examined other measures of white matter microstructure (MD, RD

and λ1) to better characterize group differences in FA. The neuroinflammation model is

supported by the finding that regions with lower FA in TMD have higher MD and RD, which are

markers of inflammation and/or œdema [63,66].

8.6. Study limitations

A few limitations in study design and results should be considered in interpretation of the data

reported in this thesis. Some of the limitations specific to each study have been discussed within

the relevant chapters.

First, the patients that were recruited for this project were deemed to have TMD based on

longstanding criteria used by the TMD clinicians at the Mount Sinai Hospital Pain Clinic and

represent a spectrum of idiopathic TMD seen clinically. However, this categorization does not

strictly adhere to the standard TMD-RDC, as described by Dworkin and LeResche [285].

Second, it is not known whether there is an impact of long-term use of medications in some of

the patients on gray and white matter plasticity. Recent evidence suggests that medications

commonly used to manage pain, such as anti-inflammatory medications and opiates, can impact

brain structure [916,962]. Only one subject was taking an opiate, hydromorphone, and so this is

of less concern to this study, but considering that thirteen of seventeen of the patients included in

this study were taking NSAIDs, it is possible that the observed age-related findings – especially

those of gray matter maintenance, may be related to the effect of these drugs..

Furthermore, it is not possible to totally rule out a contribution of ventricular volume to the VBM

analysis of subcortical structures. Additionally, some of the patients had co-morbid chronic pains

(see Table 5-1), and these pains (mostly related to TMD) may have contributed to our findings.

218

Another limitation is that duration (and intensity) may not fully describe the effect of pain on

gray and white matter structure. Due to the nature of TMD pain, other factors that were not

collected, such as the number of days of pain in a month may have more explanatory variance.

Patients also did not self-report any previous diagnosis of depression or anxiety, although these

exist in some patients with TMD [283,808,850]. Thus the contribution of these factors to gray

and white matter abnormalities in TMD cannot be ruled out.

Most of the patient sample had bilateral pain, but our findings tended to be unilateral. These

inconsistencies can simply be due to statistical thresholds. However, future studies should limit

patient recruitment to unilateral pain to determine the laterality of CNS abnormalities.

Another limitation of this thesis is that it is a cross-sectional examination of structural

abnormalities in TMD. Therefore, the interpretations of “pain-driven” and “pre-existing”

abnormalities are limited, and should be taken with caution. Furthermore, quantitative sensory

testing was not performed on the patient group recruited for this study, and so we cannot directly

relate our findings of structural abnormalities to these behavioural abnormalities.

Despite these limitations, this thesis does provide considerable evidence in line with previous

findings in the literature about gray and white matter abnormalities in chronic pain.

8.7. Future directions

This thesis has clearly demonstrated structural gray and white matter abnormalities in the brain

of patients with TMD. Future studies could build on these findings to provide more insight into

the potential underlying mechanisms. For example, behavioural measure of pain sensitivity in

these patients (obtained through quantitative sensory testing) would be useful to link the

structural abnormalities with perceptual abnormalities. Another hypothesis put forward by this

thesis is that patients with TMD (and potentially all chronic pain patients) have deficient

cognitive abilities due to the cognitive load of pain overloading the cognitive branching system.

Therefore, future studies could conduct cognitive tests that have been used to establish the

presence of a cognitive branching system in the brain to assess whether patients with TMD do

have deficient branching.

It was also demonstrated that pain intensity is related to abnormal structure in the motor cortex.

However, whether this abnormality is related to function has not been determined. Therefore, in

219

the future, motor function could be tested in the patients with TMD in the orofacial region and

elsewhere in the body. Future studies could also investigate the functional brain abnormalities

related to abnormal motor behaviour. In line with this experiment, we have clearly demonstrated

abnormalities along the CNV of patients with TMD. Nerve conduction studies could also be used

to test whether the sensory or motor aspect of the nerve is affected, and how these structural

abnormalities affect nerve function.

Furthermore, other imaging modalities can be used to assess the molecular basis of structural

abnormalities in the brain. For instance, magnetic resonance spectroscopy can be used to assess

whether there are differences in markers of specific neural structure and viability, and relate

these to structural abnormalities.

Animal models of TMD should be used to investigate the relationship between MRI-detectable

structural changes in the brain, and pain-related behaviours. These structural abnormalities could

be further characterized with genetic, cellular, electrophysiological and molecular methods.

Functional brain imaging with arterial-spin labeling, an imaging modality that allows baseline

imaging, like PET but not invasive, could be used to assess spontaneous pain in patients with

TMD. This would provide insight into which brain regions are implicated in pain processing, and

abnormally activated in TMD pain.

Prospective studies of TMD could also provide insight about the risk factors and development of

TMD. Patients with TMD could be followed over time to assess whether they respond to

treatment, or not. The factors that contribute and/or predict whether patients are treatment-

responders and non-responders could be assessed. Neuroimaging at several time points could

elucidate how the structure and function of the brain are affected by the cumulative effect of

pain, and how these abnoramltities may vary between the groups.

Fewer patients with idiopathic TMD respond to treatment, compared to patients with post-

traumatic TMD. Therefore, another study could compare whether the same functional and

structural abnormalities occur in patients with idiopathic and post-traumatic TMD. The study

could provide insight about peripheral pain-driven abnormalities (in post-traumatic TMD) and

central mechanisms that contribute to TMD.

220

8.8. Conclusions

In sum, this thesis provides novel evidence for both peripheral nerve and CNS abnormalities that

contribute to TMD pain. In support of a peripheral contribution, it was found that patients with

TMD had lower FA in both CNVs and FA in the right CNV was negatively correlated with TMD

duration. In support of CNS abnormalities, structural abnormalities in gray and white matter

were found along the ascending nociceptive, descending pain-modulatory, motor and cognitive

pathways. Furthermore, it was demonstrated that some, but not all of the observed structural

abnormalities are pain-driven. Other brain abnormalities do not show a relationship to pain

characteristics, but rather to factors independent of pain, such as age and neuroticism, a stable

personality trait. The latter findings suggest the presence of pre-existing abnormalities that may

be a predisposing risk factor to the development of TMD.

221

References

[1] National Institutes of Health Technology Assessment Conference on Management of

Temporomandibular Disorders. Bethesda, Maryland, April 29-May 1, 1996. Proceedings.

Oral Surg Oral Med Oral Pathol Oral Radiol Endod 1997;83:49-183.

[2] Adachi K, Lee JC, Hu JW, Yao D, Sessle BJ. Motor cortex neuroplasticity associated with

lingual nerve injury in rats. Somatosens Mot Res 2007;24:97-109.

[3] Adachi K, Murray GM, Lee J-C, Sessle BJ. Noxious lingual stimulation influences the

excitability of the face primary motor cerebral cortex (face MI) in the rat. J Neurophysiol

2008;100:1234-1244.

[4] Adachi K, Murray GM, Lee JC, Sessle BJ. Noxious lingual stimulation influences the

excitability of the face primary motor cerebral cortex (face MI) in the rat. J Neurophysiol

2008;100:1234-1244.

[5] Adams JE. Naloxone reversal of analgesia produced by brain stimulation in the human. Pain

1976;2:161-166.

[6] Adrian ED. Afferent discharges to the cerebral cortex from peripheral sense organs. J Physiol

1941;100:159-191.

[7] Afif A, Hoffmann D, Minotti L, Benabid AL, Kahane P. Middle short gyrus of the insula

implicated in pain processing. Pain 2008;138:546-555.

[8] Ahdab R, Ayache SS, Brugieres P, Goujon C, Lefaucheur JP. Comparison of "standard" and

"navigated" procedures of TMS coil positioning over motor, premotor and prefrontal

targets in patients with chronic pain and depression. Neurophysiol Clin 2010;40:27-36.

[9] Aigner M, Robert Lukas J, Denk M, Ziya-Ghazvini F, Kaider A, Mayr R. Somatotopic

organization of primary afferent perikarya of the guinea-pig extraocular muscles in the

trigeminal ganglion: a post-mortem DiI-tracing study. Exp Eye Res 2000;70:411-418.

[10] Aimone LD, Gebhart GF. Stimulation-produced spinal inhibition from the midbrain in the

rat is mediated by an excitatory amino acid neurotransmitter in the medial medulla. J

Neurosci 1986;6:1803-1813.

222

[11] Al-Chaer ED, Feng Y, Willis WD. Visceral pain. A disturbance in the sensorimotor

continuum? Pain Forum 1998;7:117-125.

[12] Albanese MC, Duerden EG, Rainville P, Duncan GH. Memory Traces of Pain in the Human

Cortex. J Neurosci 2007;27:4612-4621.

[13] Aldridge JW, Anderson RJ, Murphy JT. The role of the basal ganglia in controlling a

movement initiated by a visually presented cue. Brain Res 1980;192:3-16.

[14] Alexander AL. Deterministic white matter tractography. In: DK Jones, editor. Diffusion

MRI: Theory, Methods and Applications. New York: Oxford University Press, 2011. pp.

383-395.

[15] Alexander DC, Seunarine KK. Mathematics of crossing fibres. In: DK Jones, editor.

Diffusion MRI: Theory, Methods and Applications. New York: Oxford University Press,

2011. pp. 451-464.

[16] Allison T, McCarthy G, Wood CC, Darcey TM, Spencer DD, Williamson PD. Human

cortical potentials evoked by stimulation of the median nerve. I. Cytoarchitectonic areas

generating short-latency activity. J Neurophysiol 1989;62:694-710.

[17] Amano N, Hu JW, Sessle BJ. Responses of neurons in feline trigeminal subnucleus caudalis

(medullary dorsal horn) to cutaneous, intraoral, and muscle afferent stimuli. J

Neurophysiol 1986;55:227-243.

[18] Aminoff MJ. Brown-Séquard and His Work on the Spinal Cord. Spine 1996;21:133-140.

[19] Andersen E. Periaqueductal gray and cerebral cortex modulate responses of medial thalamic

neurons to noxious stimulation. Brain Res 1986;375:30-36.

[20] Andrew D, Krout KE, Craig AD. Differentiation of lamina I spinomedullary and

spinothalamic neurons in the cat. J Comp Neurol 2003;458:257-271.

[21] Apkarian AV. Thalamic anatomy and physiology of pain perception: connectivity, somato-

visceral convergence and spatio-temporal dynamics of nociceptive information coding.

In: JM Besson, G Guilbaud, H Ollat, editors. Forebrain areas involved in pain processing.

Paris: John Libbey Eurotext, 1995. pp. 93-118.

[22] Apkarian AV, Baliki MN, Geha PY. Towards a theory of chronic pain. Prog Neurobiol

2009;87:81-97.

223

[23] Apkarian AV, Bushnell MC, Treede RD, Zubieta J. Human brain mechanisms of pain

perception and regulation in health and disease. Eur J Pain 2005;9:463-484.

[24] Apkarian AV, Hodge CJ. Primate spinothalamic pathways: III. Thalamic terminations of the

dorsolateral and ventral spinothalamic pathways. J Comp Neurol 1989;288:493-511.

[25] Apkarian AV, Shi T. Squirrel monkey lateral thalamus. I. Somatic nociresponsive neurons

and their relation to spinothalamic terminals. J Neurosci 1994;14:6779-6795.

[26] Apkarian AV, Sosa Y, Krauss BR, Thomas PS, Fredrickson BE, Levy RE, Harden RN,

Chialvo DR. Chronic pain patients are impaired on an emotional decision-making task.

Pain 2004;108:129-136.

[27] Apkarian AV, Sosa Y, Sonty S, Levy RM, Harden RN, Parrish TB, Gitelman DR. Chronic

back pain is associated with decreased prefrontal and thalamic gray matter density. J

Neurosci 2004;24:10410-10415.

[28] Apkarian AV, Stea RA, Manglos SH, Szeverenyi NM, King RB, Thomas FD. Persistent

pain inhibits contralateral somatosensory cortical activity in humans. Neurosci Lett

1992;140:141-147.

[29] Applebaum AE, Beall JE, Foreman RD, Willis WD. Organization and receptive fields of

primate spinothalamic tract neurons. J Neurophysiol 1975;38:572-586.

[30] Applebaum AE, Leonard RB, Kenshalo DR, Jr., Martin RF, Willis WD. Nuclei in which

functionally identified spinothalamic tract neurons terminate. J Comp Neurol

1979;188:575-586.

[31] Arendt-Nielsen L, Brennum J, Sindrup S, Bak P. Electrophysiological and psychophysical

quantification of temporal summation in the human nociceptive system. EurJApplPhysiol

1994;68:266-273.

[32] Aron AR, Fletcher PC, Bullmore ET, Sahakian BJ, Robbins TW. Stop-signal inhibition

disrupted by damage to right inferior frontal gyrus in humans. Nat Neurosci 2003;6:115-

117.

[33] Arvidsson J. Somatotopic organization of vibrissae afferents in the trigeminal sensory nuclei

of the rat studied by transganglionic transport of HRP. J Comp Neurol 1982;211:84-92.

224

[34] Asaki S, Sekikawa M, Kim YT. Sensory innervation of temporomandibular joint disk. J

Orthop Surg (Hong Kong) 2006;14:3-8.

[35] Asghari A, Nicholas MK. Personality and pain-related beliefs/coping strategies: a

prospective study. Clin J Pain 2006;22:10-18.

[36] Ashburner J, Friston KJ. Voxel-based morphometry--the methods. Neuroimage

2000;11:805-821.

[37] Ashburner J, Friston KJ. Voxel-based morphometry. In: KJ Friston, editor. Statistical

parametric mapping the analysis of funtional brain images. Amsterdam ; Boston:

Elsevier/Academic Press, 2007. pp. 92-98.

[38] Atweh SF, Kuhar MJ. Autoradiographic localization of opiate receptors in rat brain. III. The

telencephalon. Brain Res 1977;134:393-405.

[39] Augustine JR. The insular lobe in primates including humans. Neurological Res 1985;7:2-

10.

[40] Augustine JR. Circuitry and functional aspects of the insular lobe in primates including

humans. Brain Res Rev 1996;22:229-244.

[41] Avivi-Arber L, Martin R, Lee JC, Sessle BJ. Face sensorimotor cortex and its

neuroplasticity related to orofacial sensorimotor functions. Arch Oral Biol 2011.

[42] Ayesh EE, Jensen TS, Svensson P. Somatosensory function following painful repetitive

electrical stimulation of the human temporomandibular joint and skin. Exp Brain Res

2007;179:415-425.

[43] Azkue JJ, Mateos JM, Elezgarai I, Benitez R, Lazaro E, Streit P, Grandes P. Glutamate-like

immunoreactivity in ascending spinofugal afferents to the rat periaqueductal grey. Brain

Res 1998;790:74-81.

[44] Badre D, D'Esposito M. Is the rostro-caudal axis of the frontal lobe hierarchical? Nat Rev

Neurosci 2009;10:659-669.

[45] Badre D, Hoffman J, Cooney JW, D'Esposito M. Hierarchical cognitive control deficits

following damage to the human frontal lobe. Nat Neurosci 2009;12:515-522.

225

[46] Bakke M, Hansdottir R. Mandibular function in patients with temporomandibular joint pain:

a 3-year follow-up. Oral surgery, oral medicine, oral pathology, oral radiology, and

endodontics 2008;106:227-234.

[47] Baliki MN, Chialvo DR, Geha PY, Levy RM, Harden RN, Parrish TB, Apkarian AV.

Chronic pain and the emotional brain: specific brain activity associated with spontaneous

fluctuations of intensity of chronic back pain. J Neurosci 2006;26:12165-12173.

[48] Baliki MN, Geha PY, Apkarian AV. Parsing pain perception between nociceptive

representation and magnitude estimation. J Neurophysiol 2009;101:875-887.

[49] Baliki MN, Geha PY, Apkarian AV, Chialvo DR. Beyond feeling: chronic pain hurts the

brain, disrupting the default-mode network dynamics. J Neurosci 2008;28:1398-1403.

[50] Ballantine HT, Jr., Cassidy WL, Flanagan NB, Marino R, Jr. Stereotaxic anterior

cingulotomy for neuropsychiatric illness and intractable pain. J Neurosurg 1967;26:488-

495.

[51] Ballantine HT, Jr., Giriunas I. Treatment of intractable psychiatric illness and chronic pain

by stereotactic cingulotomy. In: HH Schmidek, WH Sweet, editors. Operative

neurosurgical techniques, Vol. Second Edtion. Toronto: Grune and Stratton, 1988. pp.

1069-1075.

[52] Bammer R. Basic principles of diffusion-weighted imaging. Eur J Radiol 2003;45:169-184.

[53] Bantli H, Bloedel JR, Thienprasi P. Supraspinal interactions resulting from experimental

dorsal column stimulation. J Neurosurg 1975;42:296-300.

[54] Barbas H. Anatomic basis of cognitive-emotional interactions in the primate prefrontal

cortex. Neurosci Biobehav Rev 1995;19:499-510.

[55] Barcelo AC, Filippini B, Pazo JH. The Striatum and Pain Modulation. Cell Mol Neurobiol

2011.

[56] Baron R, Baron Y, Disbrow E, Roberts TP. Brain processing of capsaicin-induced

secondary hyperalgesia: a functional MRI study. Neurology 1999;53:548-557.

[57] Barrett LF, Mesquita B, Ochsner KN, Gross JJ. The experience of emotion. Ann Rev

Psychol 2007;58:373-403.

226

[58] Basbaum AI, Fields HL. Endogenous pain control systems: brainstem spinal pathways and

endorphin circuitry. Annu Rev Neurosci 1984;7:309-338.

[59] Basser PJ, Jones DK. Diffusion-tensor MRI: theory, experimental design and data analysis -

a technical review. NMR Biomed 2002;15:456-467.

[60] Basser PJ, Mattiello J, LeBihan D. Estimation of the effective self-diffusion tensor from the

NMR spin echo. J Magn Reson B 1994;103:247-254.

[61] Basser PJ, Mattiello J, LeBihan D. MR diffusion tensor spectroscopy and imaging. Biophys

J 1994;66:259-267.

[62] Basser PJ, Ozarslan E. Anisotropic diffusion: from the apparent diffusion coefficient to the

apparent diffusion tensor. In: DK Jones, editor. Diffusion MRI : theory, methods, and

application. New York: Oxford University Press, 2011. pp. 79-91.

[63] Basser PJ, Pierpaoli C. Microstructural and physiological features of tissues elucidated by

quantitative-diffusion-tensor MRI. J Magn Reson B 1996;111:209-219.

[64] Baumgartner U, Iannetti GD, Zambreanu L, Stoeter P, Treede RD, Tracey I. Multiple

somatotopic representations of heat and mechanical pain in the operculo-insular cortex: a

high-resolution fMRI study. J Neurophysiol 2010;104:2863-2872.

[65] Baumgartner U, Tiede W, Treede RD, Craig AD. Laser-evoked potentials are graded and

somatotopically organized anteroposteriorly in the operculoinsular cortex of anesthetized

monkeys. J Neurophysiol 2006;96:2802-2808.

[66] Beaulieu C. The basis of anisotropic water diffusion in the nervous system - a technical

review. NMR Biomed 2002;15:435-455.

[67] Beaulieu C. The biological basis of diffusion. In: H Johansen-Berg, TEJ Behrens, editors.

Diffusion MRI. London: Elsevier, 2009. pp. 105-126.

[68] Becerra L, Harris W, Grant M, George E, Boas D, Borsook D. Diffuse optical tomography

activation in the somatosensory cortex: specific activation by painful vs. non-painful

thermal stimuli. PLoS One 2009;4:e8016.

[69] Beckmann M, Johansen-Berg H, Rushworth MF. Connectivity-based parcellation of human

cingulate cortex and its relation to functional specialization. J Neurosci 2009;29:1175-

1190.

227

[70] Beckstead RM. An autoradiographic examination of corticocortical and subcortical

projections of the mediodorsal-projection (prefrontal) cortex in the rat. J Comp Neurol

1979;184:43-62.

[71] Behbehani MM, Fields HL. Evidence that an excitatory connection between the

periaqueductal gray and nucleus raphe magnus mediates stimulation produced analgesia.

Brain Res 1979;170:85-93.

[72] Behrens TE, Jbabdi S. MR diffusion tractography. In: H Johansen-Berg, TE Behrens,

editors. Diffusion MRI. London: Academic Press, 2009. pp. 334-351.

[73] Behrens TE, Johansen-Berg H, Jbabdi S, Rushworth MFS, Woolrich M. Probabilistic

diffusion tractography with multiple fibre orientations. What can we gain? Neuroimage

2007;23:144-155.

[74] Behrens TE, Johansen-Berg H, Woolrich MW, Smith SM, Wheeler-Kingshott CA, Boulby

PA, Barker GJ, Sillery EL, Sheehan K, Ciccarelli O, Thompson AJ, Brady JM, Matthews

PM. Non-invasive mapping of connections between human thalamus and cortex using

diffusion imaging. Nat Neurosci 2003;6:750-757.

[75] Behrens TE, Woolrich MW, Jenkinson M, Johansen-Berg H, Nunes RG, Clare S, Matthews

PM, Brady JM, Smith SM. Characterization and propagation of uncertainty in diffusion-

weighted MR imaging. Magn Reson Med 2003;50:1077-1088.

[76] Behrens TEJ, Woolrich MW, Walton ME, Rushworth MFS. Learning the value of

information in an uncertain world. Nat Neurosci 2007;10:1214-1221.

[77] Beitz AJ. The organization of afferent projections to the midbrain periaqueductal gray of the

rat. Neuroscience 1982;7:133-159.

[78] Belforte JE, Pazo JH. Striatal inhibition of nociceptive responses evoked in trigeminal

sensory neurons by tooth pulp stimulation. J Neurophysiol 2005;93:1730-1741.

[79] Bell C, Shaw A. Reprint of the "Idea of a New Anatomy of the Brain," with Letters, &c. J

Anat Physiol 1868;3:147-182.

[80] Bengtsson SL, Nagy Z, Skare S, Forsman L, Forssberg H, Ullen F. Extensive piano

practicing has regionally specific effects on white matter development. Nat Neurosci

2005;8:1148-1150.

228

[81] Benoliel R, Sharav Y. Masticatory myofacial pain, and tension-type and chronic daily

headache. In: Y Sharav, R Benoliel, editors. Orofacial pain & Headache. New York:

Elsevier, 2008. pp. 109-148.

[82] Bereiter DA, Hargreaves KM, Hu JW. Trigeminal mechanisms of nociception: peripheral

and brainstem organization. In: A Basbaum, MC Bushnell, editors. Science of Pain. San

Diego: Elsevier, 2009. pp. 434-460.

[83] Bereiter DA, Hirata H, Hu JW. Trigeminal subnucleus caudalis: beyond homologies with

the spinal dorsal horn. Pain 2000;88:221-224.

[84] Bergfield KL, Hanson KD, Chen K, Teipel SJ, Hampel H, Rapoport SI, Moeller JR,

Alexander GE. Age-related networks of regional covariance in MRI gray matter:

Reproducible multivariate patterns in healthy aging. Neuroimage 2009;49:1750-1759.

[85] Berkley KJ. Spatial relationships between the terminations of somatic sensory and motor

pathways in the rostral brainstem of cats and monkeys. I. Ascending somatic sensory

inputs to lateral diencephalon. J Comp Neurol 1980;193:283-317.

[86] Berlucchi G. The origin of the term plasticity in the neurosciences: Ernesto Lugaro and

chemical synaptic transmission. J Hist Neurosci 2002;11:305-309.

[87] Bernard C, Magendie F. Leçon d’ouverture du Cours de Médecine du Collège de France.

Paris: Baillière, 1856.

[88] Bernard JF, Huang GF, Besson JM. Nucleus centralis of the amygdala and the globus

pallidus ventralis: Electrophysiological evidence for an involvement in pain processes. J

Neurophysiol 1992;68:551-569.

[89] Bessou P, Perl ER. Response of cutaneous sensory units with unmyelinated fibers to

noxious stimuli. J Neurophysiol 1969;32:1025-1043.

[90] Bhalang K, Sigurdsson A, Slade GD, Maixner W. Associations among four modalities of

experimental pain in women. J Pain 2005;6:604-611.

[91] Biedenbach MA, Van Hassel HJ, Brown AC. Tooth pulp-driven neurons in somatosensory

cortex of primates: role in pain mechanisms including a review of the literature. Pain

1979;7:31-50.

229

[92] Biemond A. The conduction of pain above the level of the thalamus opticus. AMA Arch

Neurol Psychiatry 1956;75:231-244.

[93] Bingel U, Herken W, Teutsch S, May A. Habituation to painful stimulation involves the

antinociceptive system--a 1-year follow-up of 10 participants. Pain 2008;140:393-394.

[94] Bingel U, Schoell E, Herken W, Buchel C, May A. Habituation to painful stimulation

involves the antinociceptive system. Pain 2007;131:21-30.

[95] Bingel U, Tracey I. Imaging CNS modulation of pain in humans. Physiology 2008;23:371-

380.

[96] Birbaumer N, Lutzenberger W, Montoya P, Larbig W, Unertl K, Tîpfner S, Grodd W, Taub

E, Flor H. Effects of regional anesthesia on phantom limb pain are mirrored in changes in

cortical reorganization. J Neurosci 1997;17:5503-5508.

[97] Blackburn TP, Cross AJ, Hille C, Slater P. Autoradiographic localization of delta opiate

receptors in rat and human brain. Neuroscience 1988;27:497-506.

[98] Blankstein U, Chen J, Diamant NE, Davis KD. Altered brain structure in irritable bowel

syndrome: potential contributions of pre-existing and disease-driven factors.

Gastroenterology 2010;138:1783-1789.

[99] Blinkov SM, Glezer II. The human brain in figures and tables: a quantitative handbook.

New York: Plenum Press, 1968.

[100] Blits KC. Aristotle: form, function, and comparative anatomy. Anat Rec 1999;257:58-63.

[101] Blomqvist A, Zhang ET, Craig AD. Cytoarchitectonic and immunohistochemical

characterization of a specific pain and temperature relay, the posterior portion of the

ventral medial nucleus, in the human thalamus. Brain 2000;123 Pt 3:601-619.

[102] Boivie J. An anatomical reinvestigation of the termination of the spinothalamic tract in the

monkey. J Comp Neurol 1979;186:343-369.

[103] Bonicalzi V, Canavero S. Motor cortex stimulation for central and neuropathic pain (Letter

regarding Topical Review by Brown and Barbaro). Pain 2004;108:199-200; author reply

200.

230

[104] Bookstein FL. "Voxel-based morphometry" should not be used with imperfectly registered

images. Neuroimage 2001;14:1454-1462.

[105] Boorman ED, Behrens TE, Woolrich MW, Rushworth MF. How green is the grass on the

other side? Frontopolar cortex and the evidence in favor of alternative courses of action.

Neuron 2009;62:733-743.

[106] Boring EG. Sensation and perception in the history of experimental psychology. New

York, London,: D. Appleton-Century Company, 1942.

[107] Borsook D, Upadhyay J, Chudler EH, Becerra L. A key role of the basal ganglia in pain

and analgesia--insights gained through human functional imaging. Mol Pain 2010;6:27.

[108] Botvinick MM, Cohen JD, Carter CS. Conflict monitoring and anterior cingulate cortex: an

update. Trends Cogn Sci 2004;8:539-546.

[109] Boudreau S, Romaniello A, Wang K, Svensson P, Sessle BJ, Arendt-Nielsen L. The effects

of intra-oral pain on motor cortex neuroplasticity associated with short-term novel

tongue-protrusion training in humans. Pain 2007;132:169-178.

[110] Boyd D, Butler M, Carr DB, Cohen MJ, Devor M, Dworkin R, Greenspan JD, Jensen TS,

King S, Koltzenburg M, Loeser JD, Merskey H, Okifuji A, Paice J, Serra J, Treede RD,

Woda A. IASP Pain Terminology. An update on the IASP Taskforce on Taxonomy. Part

III: Pain Terms, A Current List with Definitions and Notes on Usage. In: JD Loeser,

editor. IASP Taxonomy, Vol. 2011. Seattle: IASP Press, 2011.

[111] Boyke J, Driemeyer J, Gaser C, Buchel C, May A. Training-induced brain structure

changes in the elderly. J Neurosci 2008;28:7031-7035.

[112] Bozzali M, Franceschi M, Falini A, Pontesilli S, Cercignani M, Magnani G, Scotti G,

Comi G, Filippi M. Quantification of tissue damage in AD using diffusion tensor and

magnetization transfer MRI. Neurology 2001;57:1135-1137.

[113] Bragdon EE, Light KC, Costello NL, Sigurdsson A, Bunting S, Bhalang K, Maixner W.

Group differences in pain modulation: pain-free women compared to pain-free men and

to women with TMD. Pain 2002;96:227-237.

231

[114] Brighina F, Piazza A, Vitello G, Aloisio A, Palermo A, Daniele O, Fierro B. rTMS of the

prefrontal cortex in the treatment of chronic migraine: a pilot study. J Neurol Sci

2004;227:67-71.

[115] Brody H. Organization of the cerebral cortex. III. A study of aging in the human cerebral

cortex. J Comp Neurol 1955;102:511-516.

[116] Brooks JC, Tracey I. The insula: a multidimensional integration site for pain. Pain

2007;128:1-2.

[117] Brown JA, Barbaro NM. Motor cortex stimulation for central and neuropathic pain: current

status. Pain 2003;104:431-435.

[118] Brown JA, Pilitsis JG. Motor cortex stimulation for central and neuropathic facial pain: a

prospective study of 10 patients and observations of enhanced sensory and motor

function during stimulation. Neurosurgery 2005;56:290-297.

[119] Brown MA, Semelka RC. MRI : basic principles and applications. Hoboken, N.J.: Wiley-

Blackwell, 2010.

[120] Bryan CP. The Papyrus Ebers. Translated from the German version by Cyril P. Bryan ...

With an introduction by Professor G. Elliot Smith. [With plates.]. London: Geoffrey Bles,

1930.

[121] Buckalew N, Haut MW, Morrow L, Weiner D. Chronic pain is associated with brain

volume loss in older adults: preliminary evidence. Pain Med 2008;9:240-248.

[122] Budge EATW. The book of medicines : ancient Syrian anatomy, pathology and theraputics

[sic]. London: Kegan Paul, 2002.

[123] Buhle J, Wager TD. Performance-dependent inhibition of pain by an executive working

memory task. Pain 2010;149:19-26.

[124] Burgess PR, Perl ER. Myelinated afferent fibres responding specifically to noxious

stimulation of the skin. J Physiol 1967;190:541-562.

[125] Burgmer M, Gaubitz M, Konrad C, Wrenger M, Hilgart S, Heuft G, Pfleiderer B.

Decreased gray matter volumes in the cingulo-frontal cortex and the amygdala in patients

with fibromyalgia. Psychosom Med 2009;71:566-573.

232

[126] Burton H, Craig AD, Jr. Distribution of trigeminothalamic projection cells in cat and

monkey. Brain Res 1979;161:515-521.

[127] Burton H, Craig AD, Poulos DA, Molt JT. Efferent projections from temperature sensitive

recording loci within the marginal zone of the nucleus caudalis of the spinal trigeminal

complex in the cat. J Comp Neurol 1979;183:753-777.

[128] Bush G, Luu P, Posner MI. Cognitive and emotional influences in anterior cingulate

cortex. Trends Cogn Sci 2000;4:215-222.

[129] Bushnell MC, Duncan GH. Sensory and affective aspects of pain perception: Is medial

thalamus restricted to emotional issues. Exp Brain Res 1989;78:415-418.

[130] Bushnell MC, Duncan GH, Hofbauer RK, Ha B, Chen JI, Carrier B. Pain perception: is

there a role for primary somatosensory cortex? Proc Natl Acad Sci U S A 1999;96:7705-

7709.

[131] Cadden SW, Van der Glas HW, Lobbezoo F, Van der Bilt A. Effects of remote noxious

stimulation on exteroceptive reflexes in human jaw-closing muscles. Brain Research

1996;726:189-197.

[132] Cadden SW, Van Der Glas HW, Van Der Bilt A. Modulation of jaw reflexes by remote

noxious stimulation and mental state: possible association with psychological

measurements of mental stress and occupation. J Oral Rehabil 1999;26:952-961.

[133] Cairns BE, Sessle BJ, Hu JW. Evidence that excitatory amino acid receptors within the

temporomandibular joint region are involved in the reflex activation of the jaw muscles. J

Neurosci 1998;18:8056-8064.

[134] Cajal SRy. Textura des sistema nerviosa del hombre y de Los Vertebrados. S.l.: Nicolas

Moya, 1904.

[135] Cameron AA, Khan IA, Westlund KN, Cliffer KD, Willis WD. The Efferent projections of

the periaqueductal gray in the rat: A phaselous vulgaris -leucoagglutinin study. I.

Ascending projections. J Comp Neurol 1995;351:568-584.

[136] Campbell JN, LaMotte RH. Latency to detection of first pain. Brain Research

1983;266:203-208.

233

[137] Cannon TD, van Erp TG, Bearden CE, Loewy R, Thompson P, Toga AW, Huttunen MO,

Keshavan MS, Seidman LJ, Tsuang MT. Early and late neurodevelopmental influences in

the prodrome to schizophrenia: contributions of genes, environment, and their

interactions. Schizophr Bull 2003;29:653-669.

[138] Capra NF, Dessem D. Central connections of trigeminal primary afferent neurons:

topographical and functional considerations. Crit Rev Oral Biol Med 1992;4:1-52.

[139] Carlson CR, Reid KI, Curran SL, Studts J, Okeson JP, Falace D, Nitz A, Bertrand PM.

Psychological and physiological parameters of masticatory muscle pain. Pain

1998;76:297-307.

[140] Carpenter MB, Hanna GR. Fiber projections from the spinal trigeminal nucleus in the cat. J

Comp Neurol 1961;117:117-131.

[141] Carstens E, Kuenzler N, Handwerker HO. Activation of neurons in rat trigeminal

subnucleus caudalis by different irritant chemicals applied to oral or ocular mucosa. J

Neurophysiol 1998;80:465-492.

[142] Casey KL. Unit analysis of nociceptive mechanisms in the thalamus of awake squirrel

monkey. J Neurophysiol 1966;29.

[143] Casey KL. Concepts of pain mechanisms: the contribution of functional imaging of the

human brain. Prog Brain Res 2000;129:277-287.

[144] Casey KL. Central pain: distributed effects of focal lesions. Pain 2004;108:205-206.

[145] Catani M. From hodology to function. Brain 2007;130:602-605.

[146] Catani M. The functional anatomy of white matter: from postmortem dissections to in vivo

virtual tractography. In: DK Jones, editor. Diffusion MRI: Theory, Methods and

Applications. Oxford: Oxford University Press, 2011. pp. 5-18.

[147] Catani M, ffytche DH. The rises and falls of disconnection syndromes. Brain

2005;128:2224-2239.

[148] Cauna N, Mannan G. The structure of human digital pacinian corpuscles (corpus cula

lamellosa) and its functional significance. J Anat 1958;92:1-20.

234

[149] Cauna N, Ross LL. The fine structure of Meissner's touch corpuscles of human fingers. J

Biophys Biochem Cytol 1960;8:467-482.

[150] Cavada C, Compa§y T, Tejedor J, Cruz-Rizzolo RJ, Reinoso-Su†rez F. The anatomical

connections of the macaque monkey orbitofrontal cortex. A review. Cereb Cortex

2000;10:220-242.

[151] Cercignani M. Strategies for patient-control comparison of diffusion MR data. In: DK

Jones, editor. Diffusion MRI : theory, methods, and application. New York: Oxford

University Press, 2011. pp. 485-499.

[152] Cercignani M, Iannucci G, Rocca MA, Comi G, Horsfield MA, Filippi M. Pathologic

damage in MS assessed by diffusion-weighted and magnetization transfer MRI.

Neurology 2000;54:1139-1144.

[153] Cervero F, Laird JMA. Mechanisms of touch-evoked pain (allodynia): A new model. Pain

1996;68:13-23.

[154] Chapman WP, Rose AS, Solomon HC. Measurements of heat stimulus producing motor

withdrawal reaction in patients following frontal lobotomy. Association For Research In

Nervous and Mental Disease 1948;XXVII:754-768.

[155] Charron S, Koechlin E. Divided representation of concurrent goals in the human frontal

lobes. Science 2010;328:360-363.

[156] Chen J, Blankstein U, Diamant N, Davis KD. White matter abnormalities in irritable bowel

syndrome and relation to individual factors. Brain Res 2011;1392:121-131.

[157] Cheong JY, Yoon TS, Lee SJ. Evaluations of inhibitory effect on the motor cortex by

cutaneous pain via application of capsaicin. Electromyogr Clin Neurophysiol

2003;43:203-210.

[158] Chiang CY, Chen IC, Dostrovsky JO, Sessle BJ. Anterior pretectal nucleus-induced

modulatory effects on trigeminal brainstem somatosensory neurons. Neurosci Lett

1992;134:233-237.

[159] Chiang CY, Dostrovsky JO, Iwata K, Sessle BJ. Role of glia in orofacial pain.

Neuroscientist 2011;17:303-320.

235

[160] Christensen BN, Perl ER. Spinal neurons specifically excited by noxious or thermal

stimuli: Marginal zone of the dorsal horn. Journal of Neurophysiology 1970;33:293-307.

[161] Chudler EH, Anton F, Dubner R, Kenshalo DR, Jr. Responses of nociceptive S1 neurons in

monkeys and pain sensation in humans elicited by noxious thermal stimulation: Effect of

interstimulus interval. J Neurophysiol 1990;63:559-569.

[162] Chudler EH, Dong WK. The role of the basal ganglia in nociception and pain. Pain

1995;60:3-38.

[163] Chudler EH, Lu Y. Nociceptive behavioral responses to chemical, thermal and mechanical

stimulation after unilateral, intrastriatal administration of 6-hydroxydopamine. Brain Res

2008;1213:41-47.

[164] Chudler EH, Sugiyama K, Dong WK. Nociceptive responses in the neostriatum and globus

pallidus of the anesthetized rat. J Neurophysiol 1993;69:1890-1903.

[165] Chudler EH, Sugiyama K, Dong WK. Multisensory convergence and integration in the

neostriatum and globus pallidus of the rat. Brain Res 1995;674:33-45.

[166] Clark FM, Proudfit HK. Projections of neurons in the ventromedial medulla to pontine

catecholamine cell groups involved in the modulation of nociception. Brain Res

1991;540:105-115.

[167] Cody FW, Lee RW, Taylor A. A functional analysis of the components of the

mesencephalic nucleus of the fifth nerve in the cat. J Physiol 1972;226:249-261.

[168] Coffield JA, Bowen KK, Miletic V. Retrograde tracing of projections between the nucleus

submedius, the ventrolateral orbital cortex, and the midbrain in the rat. J Comp Neurol

1992;321:488-499.

[169] Coghill RC, Sang CN, Maisog JM, Iadarola MJ. Pain Intensity Processing Within the

Human Brain: A Bilateral, Distributed Mechanism. J Neurophysiol 1999;82:1934-1943.

[170] Cohen J, Cohen P, West SG, Aiken LS. Applied Multiple Regression/Correlation Analysis

for the Behavioural Sciences. Mawhaw: Lawrence Erlbaum Associates, Inc, 2003.

[171] Colloca L, Benedetti F. Placebos and painkillers: is mind as real as matter? Nat Rev

Neurosci 2005;6:545-552.

236

[172] Colloca L, Benedetti F, Porro CA. Experimental designs and brain mapping approaches for

studying the placebo analgesic effect. Eur J Appl Physiol 2008;102:371-380.

[173] Condes-Lara M, Omana Zapata I, Leon-Olea M, Sanchez-Alvarez M. Dorsal raphe and

nociceptive stimulations evoke convergent responses on the thalamic centralis lateralis

and medial prefrontal cortex neurons. Brain Res 1989;499:145-152.

[174] Conti PC, dos Santos CN, Kogawa EW, de Castro Ferreira Conti A, de Araujo Cdos R.

The treatment of painful temporomandibular joint clicking with oral splints: a

randomized clinical trial. J Am Dent Assoc 2006;137:1108-1114.

[175] Cooper SJ. Anaesthetisation of prefrontal cortex and response to noxious stimulation.

Nature 1975;254:439-440.

[176] Corbin KB, Harrison F. Function of Mesencephalic Root in Fifth Cranial Nerve. J

Neurophysiol 1940;3.

[177] Costa PT. Influence of the Normal Personality Dimension of Neuroticism on Chest Pain

Symptoms and Coronary Artery Disease. Am J Cardiol 1987;60:J26-33.

[178] Costa PT, McCrae RR. Revised NEO personality (NEO-PI-R) and NEO five factor

inventory (NEO-FFI) professional manual. Book Revised NEO personality (NEO-PI-R)

and NEO five factor inventory (NEO-FFI) professional manual. City: Psychology

Assessment Resources, 1992.

[179] Costa PT, McCrae RR, Zonderman AB, Barbano HE, Lebowitz B, Larson DM. Cross-

Sectional Studies of Personality in a National Sample: 2. Stability in Neuroticism,

Extraversion, and Openness. Psych Aging 1986;1:144-149.

[180] Costello NL, Bragdon EE, Light KC, Sigurdsson A, Bunting S, Grewen K, Maixner W.

Temporomandibular disorder and optimism: relationships to ischemic pain sensitivity and

interleukin-6. Pain 2002;100:99-110.

[181] Costen JB. A syndrome of ear and sinus symptoms dependent upon disturbed function of

the temporomandibular joint. Ann Otol Rhinol Laryngol 1934;43:1-15.

[182] Coulter JD, Maunz RA, Willis WD. Effects of stimulation of sensorimotor cortex on

primate spinothalamic neurons. Brain Res 1974;65:351-356.

237

[183] Craig AD. How do you feel? Interoception: the sense of the physiological condition of the

body. Nat Rev Neurosci 2002;3:655-666.

[184] Craig AD. Pain mechanisms: labeled lines versus convergence in central processing. Ann

Rev Neurosci 2003;26:1-30.

[185] Craig AD. Distribution of trigeminothalamic and spinothalamic lamina I terminations in

the macaque monkey. J Comp Neurol 2004;477:119-148.

[186] Craig AD. Human feelings: why are some more aware than others? Trends Cogn Sci

2004;8:239-241.

[187] Craig AD. Once an island, now the focus of attention. Brain Struct Funct 2010;124:395-

396.

[188] Craig AD, Bushnell MC, Zhang ET, Blomqvist A. A thalamic nucleus specific for pain and

temperature sensation. Nature 1994;372:770-773.

[189] Craig AD, Dostrovsky JO. Processing of nociceptive information at supraspinal levels. In:

TL Yaksh, et al., editors. Anesthesia: Biologic foundations. Philadelphia: Lippincott-

Raven Publishers, 1997. pp. 625-642.

[190] Craig ADB. How do you feel--now? The anterior insula and human awareness. Nat Rev

Neurosci 2009;10:59-70.

[191] Creasey H, Rapoport SI. The aging human brain. Ann Neurol 1985;17:2-10.

[192] Crombez G, Baeyens F, Eelen P. Sensory and temporal information about impending pain:

the influence of predictability on pain. Behav Res Ther 1994;32:611-622.

[193] Crombez G, Eccleston C, Baeyens F, Eelen P. The disruptive nature of pain: an

experimental investigation. Behav Res Ther 1996;34:911-918.

[194] Crombez G, Eccleston C, Baeyens F, Eelen P. Habituation and the interference of pain

with task performance. Pain 1997;70:149-154.

[195] Cruccu G, Frisardi G, Pauletti G, Romaniello A, Manfredi M. Excitability of the central

masticatory pathways in patients with painful temporomandibular disorders. Pain

1997;73:447-454.

238

[196] Cushing H. A note upon the faradic stimulation of the postcentral gyrus in conscious

patients. Brain 1909;32:44-53.

[197] Dahlstrom L. Electromyographic studies of craniomandibular disorders: a review of the

literature. J Oral Rehabil 1989;16:1-20.

[198] Dale AM, Fischl B, Sereno MI. Cortical surface-based analysis. I. Segmentation and

surface reconstruction. Neuroimage 1999;9:179-194.

[199] Dallenbach KM. Pain: history and present status. Am J Psychol 1939;52:331-347.

[200] Dancause N, Barbay S, Frost SB, Plautz EJ, Chen D, Zoubina EV, Stowe AM, Nudo RJ.

Extensive cortical rewiring after brain injury. J Neurosci 2005;25:10167-10179.

[201] Dao TT, LeResche L. Gender differences in pain. J Orofac Pain 2001;14:169-195.

[202] Darian-Smith I, Johnson KO, Dykes R. "Cold" fiber population innervating palmar and

digital skin of the monkey: responses to cooling pulses. J Neurophysiol 1973;36:325-346.

[203] Darwin C. The Expressions of the Emotions in Man and Animals. London: J. Murray,

1901.

[204] Darwin E, Darwin RW. Zoonomia; or, The laws of organic life. London,: J. Johnson, 1794.

[205] DaSilva AF, Becerra L, Makris N, Strassman AM, Gonzalez RG, Geatrakis N, Borsook D.

Somatotopic activation in the human trigeminal pain pathway. J Neurosci 2002;22:8183-

8192.

[206] DaSilva AF, Becerra L, Pendse G, Chizh B, Tully S, Borsook D. Colocalized structural

and functional changes in the cortex of patients with trigeminal neuropathic pain. PLoS

One 2008;3:e3396.

[207] DaSilva AF, Granziera C, Snyder J, Hadjikhani N. Thickening in the somatosensory cortex

of patients with migraine. Neurology 2007;69:1990-1995.

[208] DaSilva AF, Granziera C, Tuch DS, Snyder J, Vincent M, Hadjikhani N. Interictal

alterations of the trigeminal somatosensory pathway and periaqueductal gray matter in

migraine. Neuroreport 2007;18:301-305.

[209] DaSilva AFM, Granziera C, Snyder J, Hadjikhani N. Thickening in the somatosensory

cortex of patients with migraine. Neurology 2007;69:1990-1995.

239

[210] Davatzikos C. Why voxel-based morphometric analysis should be used with great caution

when characterizing group differences. Neuroimage 2004;23:17-20.

[211] Davidson RM, Gale EN. Cutaneous sensory thresholds from skin overlying masseter and

forearm in MPD patients and controls. J Dent Res 1983;62:555-558.

[212] Davis KD. Studies of pain using fMRI. In: MC Bushnell, KL Casey, editors. Pain Imaging,

Vol. 18. Seattle: IASP Press, 2000. pp. 195-210.

[213] Davis KD. Neuroimaging of pain: what does it tell us? Curr Opin Support Palliat Care

2011;5:116-121.

[214] Davis KD, Dostrovsky JO. Activation of trigeminal brain-stem nociceptive neurons by

dural artery stimulation. Pain 1986;25:395-401.

[215] Davis KD, Dostrovsky JO. Effect of trigeminal subnucleus caudalis cold block on the

cerebrovascular-evoked responses of rostral trigeminal complex neurons. Neurosci Lett

1988;94:303-308.

[216] Davis KD, Dostrovsky JO. Responses of feline trigeminal spinal tract nucleus neurons to

stimulation of the middle meningeal artery and sagittal sinus. J Neurophysiol

1988;59:648-666.

[217] Davis KD, Hutchison WD, Lozano AM, Dostrovsky JO. Altered pain and temperature

perception following cingulotomy and capsulotomy in a patient with schizoaffective

disorder. Pain 1994;59:189-199.

[218] Davis KD, Hutchison WD, Lozano AM, Tasker RR, Dostrovsky JO. Human anterior

cingulate cortex neurons modulated by attention- demanding tasks. J Neurophysiol

2000;83:3575-3577.

[219] Davis KD, Kiss ZHT, Luo L, Tasker RR, Lozano AM, Dostrovsky JO. Phantom sensations

generated by thalamic microstimulation. Nature 1998;391:385-387.

[220] Davis KD, Kwan CL, Crawley AP, Mikulis DJ. fMRI of the anterior cingulate cortex

during painful, thermal and motor tasks in individual subjects. Neuroimage 1998;7:S426.

[221] Davis KD, Kwan CL, Crawley AP, Mikulis DJ. Functional MRI study of thalamic and

cortical activations evoked by cutaneous heat, cold and tactile stimuli. J Neurophysiol

1998;80:1533-1546.

240

[222] Davis KD, Kwan CL, Crawley AP, Mikulis DJ. fMRI of cortical and thalamic activations

correlated to the magnitude of pain. In: M Devor, MC Rowbotham, Z Wiesenfeld-Hallin,

editors. Progress in pain research and management, Proceedings of the 9th World

Congress on Pain, Vol. 9. Seattle: IASP Press, 2000. pp. 497-505.

[223] Davis KD, Lozano AM, Manduch M, Tasker RR, Kiss ZHT, Dostrovsky JO. Thalamic

relay site for cold perception in humans. J Neurophysiol 1999;81:1970-1973.

[224] Davis KD, Meyer RA, Campbell JN. Chemosensitivity and sensitization of nociceptive

afferents that innervate the hairy skin of monkey. J Neurophysiol 1993;69:1071-1081.

[225] Davis KD, Pope G, Chen J, Kwan CL, Crawley AP, Diamant NE. Cortical thinning in IBS:

implications for homeostatic, attention, and pain processing. Neurology 2008;70:153-

154.

[226] Davis KD, Pope GE, Crawley AP, Mikulis DJ. Neural correlates of prickle sensation: a

percept-related fMRI study. Nat Neurosci 2002;5:1121-1122.

[227] Davis KD, Pope GE, Crawley AP, Mikulis DJ. Perceptual illusion of "paradoxical heat"

engages the insular cortex. J Neurophysiol 2004;92:1248-1251.

[228] Davis KD, Taub E, Duffner F, Lozano AM, Tasker RR, Houle S, Dostrovsky JO.

Activation of the anterior cingulate cortex by thalamic stimulation in patients with

chronic pain: a positron emission tomography study. J Neurosurg 2000;92:64-69.

[229] Davis KD, Taylor KS, Hutchison WD, Dostrovsky JO, McAndrews MP, Richter EO,

Lozano AM. Human anterior cingulate cortex neurons encode cognitive and emotional

demands. J Neurosci 2005;25:8402-8406.

[230] Davis KD, Taylor SJ, Crawley AP, Wood ML, Mikulis DJ. Functional MRI of pain- and

attention-related activations in the human cingulate cortex. J Neurophysiol 1997;77:3370-

3380.

[231] Davis KD, Wood ML, Crawley AP, Mikulis DJ. fMRI of human somatosensory and

cingulate cortex during painful electrical nerve stimulation. Neuroreport 1995;7.

[232] Davis PJM, Wright EA. A new method for measuring cranial cavity volume and its

application to assessment of cerebral atrophy at autopsy. Neuropathol Appl Neurobiol

1977;3:341-358.

241

[233] Davison C, Schick W. Spontaneous pain and other subjective sensory disturbances. Arch

Neurol Psychiatry 1935;34:1204–1237.

[234] de Leeuw R, Davis CE, Albuquerque R, Carlson CR, Andersen AH. Brain activity during

stimulation of the trigeminal nerve with noxious heat. Oral Surg Oral Med Oral Pathol

Oral Radiol Endod 2006;102:750-757.

[235] Deco G, Corbetta M. The dynamical balance of the brain at rest. Neuroscientist

2011;17:107-123.

[236] Deco G, Jirsa VK, McIntosh AR. Emerging concepts for the dynamical organization of

resting-state activity in the brain. Nat Rev Neurosci 2011;12:43-56.

[237] Dejerine J. Anatomie des centres nerveux. Paris,: Reuff, 1895.

[238] DeLeo JA, Tanga FY, Tawfik VL. Neuroimmune Activation and Neuroinflammation in

Chronic Pain and Opioid Tolerance/Hyperalgesia. Neuroscientist 2004;10:40-52.

[239] DeLong MR. Activity of pallidal neurons during movement. J Neurophysiol 1971:414-

427.

[240] Derbyshire SWG, Jones AKP, Gyulai F, Clark S, Townsend D, Firestone LL. Pain

processing during three levels of noxious stimulation produces differential patterns of

central activity. Pain 1997;73:431-445.

[241] Descartes R. De Homine - Figuris et latinitate donatus Florentio Schuyl. Leiden: Leffer &

Moyardus, 1662.

[242] Descartes R, Clerselier C, La Forge Ld, Schuyl F. L'homme ... et un traitté De la formation

du foetus du mesme autheur. Paris: Charles Angot, 1664.

[243] Destrieux C, Fischl B, Dale A, Halgren E. Automatic parcellation of human cortical gyri

and sulci using standard anatomical nomenclature. Neuroimage 2010;53:1-15.

[244] Dettmers C, Adler T, Rzanny R, Van Schayck R, Gaser C, Weiss T, Miltner WH, Bruckner

L, Weiller C. Increased excitability in the primary motor cortex and supplementary motor

area in patients with phantom limb pain after upper limb amputation. Neurosci Lett

2001;307:109-112.

242

[245] DeYoung CG, Hirsh JB, Shane MS, Papademetris X, Rajeevan N, Gray JR. Testing

predictions from personality neuroscience: brain structure and the big five. Psychol Sci

2010;21:820-828.

[246] Diamant NE. Overview of functional gut disorders: a challenge. Scandinavian journal of

gastroenterology Supplement 1995;213:1-6.

[247] Diatchenko L, Anderson AD, Slade GD, Fillingim RB, Shabalina SA, Higgins TJ, Sama S,

Belfer I, Goldman D, Max MB, Weir BS, Maixner W. Three major haplotypes of the

beta2 adrenergic receptor define psychological profile, blood pressure, and the risk for

development of a common musculoskeletal pain disorder. Am J Med Genet B

Neuropsychiatr Genet 2006;141B:449-462.

[248] Diatchenko L, Nackley AG, Slade GD, Fillingim RB, Maixner W. Idiopathic pain

disorders--pathways of vulnerability. Pain 2006;123:226-230.

[249] Diatchenko L, Slade GD, Nackley AG, Bhalang K, Sigurdsson A, Belfer I, Goldman D,

Xu K, Shabalina SA, Shagin D, Max MB, Makarov SS, Maixner W. Genetic basis for

individual variations in pain perception and the development of a chronic pain condition.

Hum Mol Genet 2005;14:135-143.

[250] Dick B, Eccleston C, Crombez G. Attentional functioning in fibromyalgia, rheumatoid

arthritis, and musculoskeletal pain patients. Arthritis Rheum 2002;47:639-644.

[251] Dickenson AH, Le Bars D. Diffuse noxious inhibitory controls (DNIC) involve

trigeminothalamic and spinothalamic neurones in the rat. Exp Brain Res 1983;49:174-

180.

[252] Dickenson AH, Le Bars D. Morphine microinjections into periaqueductal grey matter of

the rat: effects on dorsal horn neuronal responses to C-fibre activity and diffuse noxious

inhibitory controls. Life Sci 1983;33 Suppl 1:549-552.

[253] Dickenson AH, Oliveras JL, Besson JM. Role of the nucleus raphe magnus in opiate

analgesia as studied by the microinjection technique in the rat. Brain Res 1979;170:95-

111.

[254] Dierssen G, Odoriz B, Hernando C. Sensory and motor responses to stimulation of the

posterior cingulate cortex in man. J Neurosurg 1969;31:435-440.

243

[255] Djouhri L, Lawson SN. Abeta-fiber nociceptive primary afferent neurons: a review of

incidence and properties in relation to other afferent A-fiber neurons in mammals. Brain

Res Brain Res Rev 2004;46:131-145.

[256] Dodd J, Kelly J. Trigeminal system. In: ER Kandel, A Schwartz, TM Jessell, editors.

Principles of Neural Science. New York: McGraw-Hill, 2000. pp. 701-710.

[257] Dostrovsky J, Craig AD. The thalamus and nociceptive processing. In: A Basbaum, MC

Bushnell, editors. Science of Pain. San Diego: Elsevier, 2009. pp. 635-654.

[258] Dostrovsky JO, Craig AD. Ascending projection systems. In: A McMahon, M

Koltzenburg, editors. Wall and Melzack's Textbook of Pain. Edinburgh, UK: Churchill

Livingstone, 2006. pp. 187-204.

[259] Dostrovsky JO, Hutchison WD, Davis KD, Lozano A. Potential role of orbital and

cingulate cortices in nociception. In: JM Besson, G Guilbaud, H Ollat, editors. Forebrain

areas involved in pain processing. Paris: John Libbey Eurotext, 1995. pp. 171-181.

[260] Dostrovsky JO, Shah Y, Gray BG. Descending inhibitory influences from periaqueductal

gray, nucleus raphe magnus, and adjacent reticular formation. II. Effects on medullary

dorsal horn nociceptive and nonnociceptive neurons. J Neurophysiol 1983;49:948-960.

[261] Double KL, Halliday GM, Kril JJ, Harasty JA, Cullen K, Brooks WS, Creasey H, Broe

GA. Topography of brain atrophy during normal aging and Alzheimer's disease.

Neurobiol Aging 1996;17:513-521.

[262] Downar J, Crawley AP, Mikulis DJ, Davis KD. A multimodal cortical network for the

detection of changes in the sensory environment. Nat Neurosci 2000;3:277-283.

[263] Downar J, Crawley AP, Mikulis DJ, Davis KD. A cortical network for the detection of

novel events across multiple sensory modalities. Neuroimage 2001;13:S310.

[264] Downar J, Crawley AP, Mikulis DJ, Davis KD. A cortical network sensitive to stimulus

salience in a neutral behavioral context across multiple sensory modalities. J

Neurophysiol 2002;87:615-620.

[265] Downar J, Mikulis DJ, Davis KD. Neural correlates of the prolonged salience of painful

stimulation. Neuroimage 2003;20:1540-1551.

244

[266] Draganski B, Gaser C, Busch V, Schuierer G, Bogdahn U, May A. Neuroplasticity:

changes in grey matter induced by training. Nature 2004;427:311-312.

[267] Draganski B, Gaser C, Kempermann G, Kuhn HG, Winkler J, Buchel C, May A. Temporal

and spatial dynamics of brain structure changes during extensive learning. J Neurosci

2006;26:6314-6317.

[268] Draganski B, Moser T, Lummel N, Ganssbauer S, Bogdahn U, Haas F, May A. Decrease

of thalamic gray matter following limb amputation. Neuroimage 2006;31:951-957.

[269] Drangsholt MT, LeResche L. Temporomandibular disorder pain. In: IK Crombie, PR

Croft, SJ Linton, L LeResche, M Von Korff, editors. Epidemiology of pain. Seattle: IASP

Press, 1999. pp. 202-233.

[270] Dreyer DA, Metz CB, Schneider RJ, Whitsel BL. Differential contributions of spinal

pathways to body representation in postcentral gyrus of Macaca mulatta. J Neurophysiol

1974;37:119- 145.

[271] Driemeyer J, Boyke J, Gaser C, Buchel C, May A. Changes in gray matter induced by

learning--revisited. PLoS One 2008;3:e2669.

[272] Dubner R, Bennett GJ. Spinal and trigeminal mechanisms of nociception. Annu Rev

Neurosci 1983;6:381-418.

[273] Dubner R, Sessle BJ, Storey AT. The neural basis of oral and facial function. New York:

Plenum Press, 1978.

[274] Duerden EG, Albanese MC. Localization of pain-related brain activation: A meta-analysis

of neuroimaging data. Hum Brain Mapp 2011.

[275] Dufton BD. Cognitive failure and chronic pain. Int J Psychiatry Med 1989;19:291-297.

[276] Dum R, Levinthal D, Strick P. The Spinothalamic System Targets Motor and Sensory

Areas in the Cerebral Cortex of Monkeys. J Neurosci 2009;29:14223-14235.

[277] Dum RP, Levinthal DJ, Strick PL. The Spinothalamic System Targets Motor and Sensory

Areas in the Cerebral Cortex of Monkeys. Journal of Neuroscience 2009;29:14223-

14235.

245

[278] Dumitriu D, Hao J, Hara Y, Kaufmann J, Janssen WGM, Lou W, Rapp PR, Morrison JH.

Selective changes in thin spine density and morphology in monkey prefrontal cortex

correlated with age-related cognitive impairment. J Neurosci 2010;30:7507-7515.

[279] Duncan GH, Bushnell MC, Lavigne GJ. Comparison of verbal and visual analogue scales

for measuring the intensity and unpleasantness of experimental pain. Pain 1989;37:295-

303.

[280] Dusser de Barenne JG. Experimental researches on sensory localisations in the cerebral

cortex. Q J Exp Physiol Cogn Med Sci 1916;9:355-390.

[281] Dusser de Barenne JG, McCulloch WS. Functional interdependence of sensory cortex and

thalamus. J Neurophysiol 1941;4:304-310.

[282] Duyn JH. The Future of Ultra-High Field MRI and fMRI for study of the human brain.

Neuroimage 2011;in press.

[283] Dworkin SF. Perspectives on the interaction of biological, psychological and social factors

in TMD. J Am Dent Assoc 1994;125:856-863.

[284] Dworkin SF, Huggins KH, LeResche L, Von Korff M, Howard J, Truelove E, Sommers E.

Epidemiology of signs and symptoms in temporomandibular disorders: clinical signs in

cases and controls. J Am Dent Assoc 1990;120:273-281.

[285] Dworkin SF, Leresche L. Research diagnostic criteria for temporomandibular disorders:

review, criteria, examinations and specifications, critique. J Craniomandib Disord

1992;6:301-355.

[286] Dworkin SF, Massoth DL. Temporomandibular disorders and chronic pain: disease or

illness? J Prosthet Dent 1994;72:29-38.

[287] Dworkin SF, Turner JA, Wilson L, Massoth D, Whitney C, Huggins KH, Burgess J,

Sommers E, Truelove E. Brief group cognitive-behavioral intervention for

temporomandibular disorders. Pain 1994;59:175-187.

[288] Dynes JB, Poppen JL. Lobotomy for intractable pain. J Am Med Assoc 1949;140:15-19.

[289] Eccleston C. Chronic pain and attention: a cognitive approach. Br J Clin Psychol 1994;33 (

Pt 4):535-547.

246

[290] Eccleston C. The attentional control of pain: Methodological and theoretical concerns. Pain

1995;63:3-10.

[291] Eccleston C, Crombez G. Pain demands attention: a cognitive-affective model of the

interruptive function of pain. Psychol Bull 1999;125:356-366.

[292] Eccleston C, Crombez G, Aldrich S, Stannard C. Attention and somatic awareness in

chronic pain. Pain 1997;72:209-215.

[293] Echols DH, Colclough JA. Abolition of painful phantom foot by resection of the sensory

cortex. J Am Med Assoc 1947;134:1476-1477.

[294] Eickhoff SB, Jbabdi S, Caspers S, Laird AR, Fox PT, Zilles K, Behrens TE. Anatomical

and functional connectivity of cytoarchitectonic areas within the human parietal

operculum. J Neurosci 2010;30:6409-6421.

[295] Eippert F, Bingel U, Schoell ED, Yacubian J, Klinger R, Lorenz J, Buchel C. Activation of

the opioidergic descending pain control system underlies placebo analgesia. Neuron

2009;63:533-543.

[296] Erasmus JFP. Studies on pain - a review of theories and observations. S Afr Med J

1948;22:683-691.

[297] Erlanger J, Gasser HS. The action potential in fibres of slow con- duction in spinal roots

and somatic nerves. A J Physiol 1930;92:43-82.

[298] Erpelding N, Moayedi M, Davis KD. Cotical correlates of temperature and pain sensitivity.

Book Cotical correlates of temperature and pain sensitivity. City, 2011.

[299] Etkin A, Egner T, Kalisch R. Emotional processing in anterior cingulate and medial

prefrontal cortex. Trends Cogn Sci 2011;15:85-93.

[300] Ettlin DA, Brugger M, Keller T, Luechinger R, Jancke L, Palla S, Barlow A, Gallo LM,

Lutz K. Interindividual differences in the perception of dental stimulation and related

brain activity. Eur J Oral Sci 2009;117:27-33.

[301] Farina S, Tinazzi M, Le Pera D, Valeriani M. Pain-related modulation of the human motor

cortex. Neurol Res 2003;25:130-142.

247

[302] Farina S, Valeriani M, Rosso T, Aglioti S, Tamburin S, Fiaschi A, Tinazzi M. Transient

inhibition of the human motor cortex by capsaicin-induced pain. A study with

transcranial magnetic stimulation. Neurosci Lett 2001;314:97-101.

[303] Farmer MA, Chanda ML, Parks EL, Baliki MN, Apkarian AV, Schaeffer AJ. Brain

functional and anatomical changes in chronic prostatitis/chronic pelvic pain syndrome. J

Urol 2011;186:117-124.

[304] Farrell MJ, Laird AR, Egan GF. Brain activity associated with painfully hot stimuli applied

to the upper limb: a meta-analysis. Hum Brain Mapp 2005;25:129-139.

[305] Fernandez-de-las-Penas C, Galan-del-Rio F, Fernandez-Carnero J, Pesquera J, rendt-

Nielsen L, Svensson P. Bilateral widespread mechanical pain sensitivity in women with

myofascial temporomandibular disorder: evidence of impairment in central nociceptive

processing. J Pain 2009;10:1170-1178.

[306] ffytche DH, Catani M. Beyond localization: from hodology to function. Phil Trans R Soc

B 2005;360:767-779.

[307] Fields HL. Understanding how opioids contribute to reward and analgesia. Reg Anesth

Pain Med 2007;32:242-246.

[308] Fields HL, Anderson SD. Evidence that raphe-spinal neurons mediate opiate and midbrain

stimulation-produced analgesias. Pain 1978;5:333-349.

[309] Fields HL, Basbaum AI, Heinricher MM. Central nervous system mechanisms of pain

modulation. In: S McMahon, M Koltzenburg, editors. Wall and Melzack's Textbook of

Pain. Edinburgh, UK: Churchill Livingstone, 2006. pp. 125-142.

[310] Fierro B, De Tommaso M, Giglia F, Giglia G, Palermo A, Brighina F. Repetitive

transcranial magnetic stimulation (rTMS) of the dorsolateral prefrontal cortex (DLPFC)

during capsaicin-induced pain: modulatory effects on motor cortex excitability. Exp

Brain Res 2010;203:31-38.

[311] Filley CM. Neurobiology of white matter disorders. In: DK Jones, editor. Diffusion MRI:

Theory, Methods and Applications. Oxford: Oxford University Press, 2011. pp. 19-30.

[312] Fillingim RB, Maixner W. Sex-related factors in temporomandibular disorders. In: RB

Fillingim, editor. Sex, Gender, and Pain, Vol. 17. Seattle: IASP Press, 2000. pp. 309-325.

248

[313] Fillingim RB, Maixner W, Kincaid S, Sigurdsson A, Harris MB. Pain sensitivity in

patients with temporomandibular disorders: relationship to clinical and psychosocial

factors. Clin J Pain 1996;12:260-269.

[314] Fillingim RB, Maixner W, Kincaid S, Silva S. Sex differences in temporal summation but

not sensory-discriminative processing of thermal pain. Pain 1998;75:121-127.

[315] Fillingim RB, Ohrbach R, Greenspan JD, Knott C, Dubner R, Bair E, Baraian C, Slade

GD, Maixner W. Potential Psychosocial Risk Factors for Chronic TMD: Descriptive Data

and Empirically Identified Domains from the OPPERA Case-Control Study. J Pain

2011;12:T46-60.

[316] Fischl B, Dale AM. Measuring the thickness of the human cerebral cortex from magnetic

resonance images. Proc Natl Acad Sci U S A 2000;97:11050-11055.

[317] Fischl B, Liu A, Dale AM. Automated manifold surgery: constructing geometrically

accurate and topologically correct models of the human cerebral cortex. IEEE Trans Med

Imaging 2001;20:70-80.

[318] Fischl B, Salat DH, Busa E, Albert M, Dieterich M, Haselgrove C, Van der Kouwe A,

Killiany R, Kennedy D, Klaveness S, Montillo A, Makris N, Rosen B, Dale AM. Whole

brain segmentation: Automated labeling of neuroanatomical structures in the human

brain. Neuron 2002;33:341-355.

[319] Fischl B, Sereno MI, Dale AM. Cortical surface-based analysis. II: Inflation, flattening,

and a surface-based coordinate system. Neuroimage 1999;9:195-207.

[320] Fischl B, Sereno MI, Tootell RB, Dale AM. High-resolution intersubject averaging and a

coordinate system for the cortical surface. Hum Brain Mapp 1999;8:272-284.

[321] Fischl B, van der KA, Destrieux C, Halgren E, Segonne F, Salat DH, Busa E, Seidman LJ,

Goldstein J, Kennedy D, Caviness V, Makris N, Rosen B, Dale AM. Automatically

parcellating the human cerebral cortex. Cereb Cortex 2004;14:11-22.

[322] Fjell AM, Westlye LT, Amlien I, Espeseth T, Reinvang I, Raz N, Agartz I, Salat DH,

Greve DN, Fischl B, Dale AM, Walhovd KB. High consistency of regional cortical

thinning in aging across multiple samples. Cereb Cortex 2009;19:2001-2012.

249

[323] Flor H, Nikolajsen L, Staehelin Jensen T. Phantom limb pain: a case of maladaptive CNS

plasticity? Nat Rev Neurosci 2006;7:873-881.

[324] Florence SL, Kaas JH. Large-scale reorganization at multiple levels of the somatosensory

pathway follows therapeutic amputation of the hand in monkeys. J Neurosci

1995;15:8083-8095.

[325] Florence SL, Taub HB, Kaas JH. Large-scale sprouting of cortical connections after

peripheral injury in adult macaque monkeys. Science 1998;282:1117-1121.

[326] Foltz EL, White LE. The role of rostral cingulumotomy in "pain" relief. Int J Neurol

1968;6:353-373.

[327] Foltz EL, White LE, Jr. Pain "relief" by frontal cingulumotomy. J Neurosurg 1962;19:89-

100.

[328] Franchi G. Persistence of vibrissal motor representation following vibrissal pad

deafferentation in adult rats. Exp Brain Res 2001;137:180-189.

[329] Franchi G, Veronesi C. Short-term reorganization of input-deprived motor vibrissae

representation following motor disconnection in adult rats. J Physiol 2006;574:457-476.

[330] Freeman W, Watts JW. Pain of organic disease relieved by prefrontal lobotomy. Lancet

1946;1:953-955.

[331] Fregni F, Freedman S, Pascual-Leone A. Recent advances in the treatment of chronic pain

with non-invasive brain stimulation techniques. Lancet Neurol 2007;6:188-191.

[332] Frisardi G, Chessa G, Sau G, Frisardi F. Trigeminal electrophysiology: a 2 x 2 matrix

model for differential diagnosis between temporomandibular disorders and orofacial pain.

BMC musculoskeletal disorders 2010;11:141.

[333] Friston KJ. Functional and effective connectivity in neuroimaging: a synthesis. Hum Brain

Mapp 1994;2:56-78.

[334] Frokjaer JB, Olesen SS, Gram M, Yavarian Y, Bouwense SA, Wilder-Smith OH, Drewes

AM. Altered brain microstructure assessed by diffusion tensor imaging in patients with

chronic pancreatitis. Gut 2011;60:1554-1562.

250

[335] Frot M, Magnin M, Mauguiere F, Garcia-Larrea L. Human SII and posterior insula

differently encode thermal laser stimuli. Cereb Cortex 2007;17:610-620.

[336] Frot M, Mauguiere F. [Operculo-insular responses to nociceptive skin stimulation in

humans. A review of the literature]. Neurophysiol Clin 1999;29:401-410.

[337] Frot M, Mauguiere F. Dual representation of pain in the operculo-insular cortex in humans.

Brain 2003;126:438-450.

[338] Frot M, Rambaud L, Guenot M, Mauguiere F. Intracortical recordings of early pain-related

CO2-laser evoked potentials in the human second somatosensory (SII) area. Clin

Neurophysiol 1999;110:133-145.

[339] Fujiwara S, Sasaki M, Wada T, Kudo K, Hirooka R, Ishigaki D, Nishikawa Y, Ono A,

Yamaguchi M, Ogasawara K. High-resolution diffusion tensor imaging for the detection

of diffusion abnormalities in the trigeminal nerves of patients with trigeminal neuralgia

caused by neurovascular compression. J Neuroimaging 2011;21:e102-108.

[340] Fukushima T, Kerr FW. Organization of trigeminothalamic tracts and other thalamic

afferent systems of the brainstem in the rat: presence of gelatinosa neurons with thalamic

connections. J Comp Neurol 1979;183:169-184.

[341] Ganchrow D. Intratrigeminal and thalamic projections of nucleus caudalis in the squirrel

monkey (Saimiri sciureus): A degeneration and autoradiographic study. J Comp Neurol

1978;178:281-311.

[342] Garcia-Larrea L, Maarrawi J, Peyron R. Thalamocingulate Mechanisms of Precentral

Cortex Stimulation for Central Pain. In: B Vogt, editor. Cingulate Neurobiology and

Disease. Oxford: Oxford University Press, 2009. pp. 438-449.

[343] Garcia-Larrea L, Perchet C, Creac'h C, Convers P, Peyron R, Laurent B, Mauguiere F,

Magnin M. Operculo-insular pain (parasylvian pain): a distinct central pain syndrome.

Brain 2010;133:2528-2539.

[344] Garcia-Larrea L, Peyron R, Mertens P, Grégoire MC, Lavenne F, Bonnefoi F, Mauguiäre

F, Laurent B, Sindou M. Positron emission tomography during motor cortex stimulation

for pain control. Stereotact Funct Neurosurg 1997;68:141-148.

251

[345] Garcia-Larrea L, Peyron R, Mertens P, Gregoire MC, Lavenne F, Le Bars D, Convers P,

Mauguiere F, Sindou M, Laurent B. Electrical stimulation of motor cortex for pain

control: a combined PET-scan and electrophysiological study. Pain 1999;83:259-273.

[346] Gardner EP, Martin JH, Jessell TM. The bodily senses. In: ER Kandel, JH Schwartz, TM

Jessell, editors. Principles of Neural Science. New York: McGraw-Hill, 2000. pp. 432-

450.

[347] Garraghty PE, Kaas JH. Functional reorganization in adult monkey thalamus after

peripheral nerve injury. Neuroreport 1991;2:747-750.

[348] Garraghty PE, Muja N. Possible use-dependent changes in adult primate somatosensory

cortex. Brain Res 1995;686:119-121.

[349] Gatchel RJ, Garafalo JP, Ellis E, Holt C. Major psychological disorders in acute and

chronic TMD: an initial examination. J Am Dent Assoc 1996;127:1356-1374.

[350] Gatchel RJ, Stowell AW, Wildenstein L, Riggs R, Ellis E, 3rd. Efficacy of an early

intervention for patients with acute temporomandibular disorder-related pain: a one-year

outcome study. J Am Dent Assoc 2006;137:339-347.

[351] Gaze RM, Gordon G. The representation of cutaneous sense in the thalamus of the cat and

monkey. Q J Exp Physiol Cogn Med Sci 1954;39:279-304.

[352] Gear RW, Levine JD. Nucleus accumbens facilitates nociception. Exp Neurol

2011;229:502-506.

[353] Gebhart GF. Descending modulation of pain. Neurosci Biobehav Rev 2004;27:729-737.

[354] Gebhart GF, Jones SL. Effects of morphine given in the brain stem on the activity of dorsal

horn nociceptive neurons. Prog Brain Res 1988;77:229-243.

[355] Gebhart GF, Sandkuhler J, Thalhammer JG, Zimmermann M. Quantitative comparison of

inhibition in spinal cord of nociceptive information by stimulation in periaqueductal gray

or nucleus raphe magnus of the cat. J Neurophysiol 1983;50:1433-1445.

[356] Geha PY, Baliki MN, Harden RN, Bauer WR, Parrish TB, Apkarian AV. The brain in

chronic CRPS pain: abnormal gray-white matter interactions in emotional and autonomic

regions. Neuron 2008;60:570-581.

252

[357] Genovese CR, Lazar NA, Nichols T. Thresholding of statistical maps in functional

neuroimaging using the false discovery rate. Neuroimage 2002;15:870-878.

[358] Gerstner G, Ichesco E, Quintero A, Schmidt-Wilcke T. Changes in regional gray matter

and white matter volume in patients with myofascial-type temporomandibular disorders:

a voxel-based morphometry study. J Orofac Pain 2011;25:99-106.

[359] Geschwind N. Disconnexion syndromes in animals and man. I. Brain 1965;88:237-294.

[360] Geschwind N. Disconnexion syndromes in animals and man. II. Brain 1965;88:585-644.

[361] Ghez C. Motor systems in the brain. In: ER Kandel, A Schwartz, TM Jessell, editors.

Principles of Neural Science. New York: McGraw-Hill, 2000. pp. 533-537.

[362] Giesler GJ, Jr. , Spiel HR, Willis WD. Organization of spinothalamic tract axons within the

rat spinal cord. J Comp Neurol 1981;195:243-252.

[363] Giesler GJ, Jr., Katter JT, Dado RJ. Direct spinal pathways to the limbic system for

nociceptive information. Trends Neurosci 1994;17:244-250.

[364] Gitelman DR, Alpert NM, Kosslyn S, Daffner K, Scinto L, Thompson W, Mesulam MM.

Functional imaging of human right hemispheric activation for exploratory movements.

Ann Neurol 1996;39:174-179.

[365] Glass JM, Park DC. Cognitive dysfunction in fibromyalgia. Curr Rheumatol Rep

2001;3:123-127.

[366] Gobel S, Hockfield S, Ruda MA. An anatomical analysis of the similarities between

medullary and spinal dorsal horns. In: Y Kawamura, R Dubner, editors. Oral-Facial

Sensory and Motor Functions. Tokyo: Quintessence, 1981. pp. 211-223.

[367] Gobel S, Purvis MB. Anatomical studies of the organization of the spinal V nucleus: the

deep bundles and the spinal V tract. Brain Res 1972;48:27-44.

[368] Godefroy JN, Thiesson D, Pollin B, Rokyta R, Azerad J. Reciprocal connections between

the red nucleus and the trigmeinal nuclei: a retrograde and anterograde tracing study.

Physiol Res 1998;47:489-500.

253

[369] Goldberg M, Mock D, Ichise M, Proulx G, Gordon A, Shandling M, Tsai S, Tenenbaum

HC. Neuropsychological Deficit in and Clinical Features of Post-Traumatic TMD. J

Orofac Pain 1996;10.

[370] Goldberg MB, Mock D, Ichise M, Proulx G, Gordon A, Shandling M, Tsai S, Tenenbaum

HC. Neuropsychologic deficits and clinical features of posttraumatic temporomandibular

disorders. J Orofac Pain 1996;10:126-140.

[371] Goldszal AF, Davatzikos C, Pham DL, Yan MX, Bryan RN, Resnick SM. An image-

processing system for qualitative and quantitative volumetric analysis of brain images. J

Comput Assist Tomogr 1998;22:827-837.

[372] Good CD, Johnsrude IS, Ashburner J, Henson RN, Friston KJ, Frackowiak RS. A voxel-

based morphometric study of ageing in 465 normal adult human brains. Neuroimage

2001;14:21-36.

[373] Goubert L, Crombez G, Van Damme S. The role of neuroticism, pain catastrophizing and

pain-related fear in vigilance to pain: a structural equations approach. Pain 2004;107:234-

241.

[374] Graven-Nielsen T, Arendt-Nielsen L. Human models and cliical manifestations of

musculoskeletal pain and pain-motor interactions. In: T Graven-Nielsen, L Arendt-

Nielsen, S Mense, editors. Fundamentals of Muskuloskeletal Pain. Seattle: IASP Press,

2008. pp. 155-188.

[375] Graven-Nielsen T, Jansson Y, Segerdahl M, Kristensen JD, Mense S, Arendt-Nielsen L,

Sollevi A. Experimental pain by ischaemic contractions compared with pain by

intramuscular infusions of adenosine and hypertonic saline. Eur J Pain 2003;7:93-102.

[376] Graven-Nielsen T, Mense S. The peripheral apparatus of muscle pain: evidence from

animal and human studies. Clin J Pain 2001;17:2-10.

[377] Greenspan JD, Lee RR, Lenz FA. Pain sensitivity alterations as a function of lesion

location in the parasylvian cortex. Pain 1999;81:273-282.

[378] Greenspan JD, Winfield JA. Reversible pain and tactile deficits associated with a cerebral

tumor compressing the posterior insula and parietal operculum. Pain 1992;50:29-39.

254

[379] Grossi ML, Goldberg MB, Locker D, Tenenbaum HC. Reduced neuropsychologic

measures as predictors of treatment outcome in patients with temporomandibular

disorders. J Orofac Pain 2001;15:329-339.

[380] Grossi ML, Goldberg MB, Locker D, Tenenbaum HC. Irritable bowel syndrome patients

versus responding and nonresponding temporomandibular disorder patients: a

neuropsychologic profile comparative study. Int J Prosthodont 2008;21:201-209.

[381] Guo LH, Schluesener HJ. The innate immunity of the central nervous system in chronic

pain: the role of Toll-like receptors. Cell Mol Life Sci 2007;64:1128-1136.

[382] Gustin SM, Peck CC, Wilcox SL, Nash PG, Murray GM, Henderson LA. Different Pain,

Different Brain: Thalamic Anatomy in Neuropathic and Non-Neuropathic Chronic Pain

Syndromes. J Neurosci 2011;31:5956-5964.

[383] Gutierrez-Mahoney CGd. The treatment of painful phantom limb by removal of post-

central cortex. J Neurosurg 1944;1:156-162.

[384] Gwilym SE, Filippini N, Douaud G, Carr AJ, Tracey I. Thalamic atrophy associated with

painful osteoarthritis of the hip is reversible after arthroplasty: a longitudinal voxel-based

morphometric study. Arthritis Rheum 2010;62:2930-2940.

[385] Ha AD, Jankovic J. Pain in Parkinson's disease. Mov Disord 2011.

[386] Haas BW, Constable RT, Canli T. Stop the sadness: Neuroticism is associated with

sustained medial prefrontal cortex response to emotional facial expressions. Neuroimage

2008;42:385-392.

[387] Haaxma R, Kuypers HG. Intrahemispheric cortical connexions and visual guidance of

hand and finger movements in the rhusus monkey. Brain 1975;98:239-260.

[388] Haber SN, Knutson B. The Reward Circuit: Linking Primate Anatomy and Human

Imaging. Neuropsychopharmacology 2009;35:4-26.

[389] Hagbarth KE, Kerr DI. Central influences on spinal afferent conduction. J Neurophysiol

1954;17:295-307.

[390] Hagberg C, Hagberg M, Kopp S. Musculoskeletal symptoms and psychosocial factors

among patients with craniomandibular disorders. Acta Odontol Scand 1994;52:170-177.

255

[391] Hagberg C, Hellsing G, Hagberg M. Perception of cutaneous electrical stimulation in

patients with craniomandibular disorders. J Craniomandib Disord 1990;4:120-125.

[392] Hagmann P, Thiran JP, Jonasson L, Vandergheynst P, Clarke S, Maeder P, Meuli R. DTI

mapping of human brain connectivity: statistical fibre tracking and virtual dissection.

Neuroimage 2003;19:545-554.

[393] Halata Z. The ultrastructure of the sensory nerve endings in the articular capsule of the

knee joint of the domestic cat (Ruffini corpuscles and Pacinian corpuscles). J Anat

1977;124:717-729.

[394] Halata Z, Grim M, Bauman KI. Friedrich Sigmund Merkel and his "Merkel cell",

morphology, development, and physiology: review and new results. Anat Rec A Discov

Mol Cell Evol Biol 2003;271:225-239.

[395] Halkjaer L, Melsen B, McMillan AS, Svensson P. Influence of sensory deprivation and

perturbation of trigeminal afferent fibers on corticomotor control of human tongue

musculature. Exp Brain Res 2006;170:199-205.

[396] Halliday AM, Logue V. Painful sensations evoked by electrical stimulation in the

thalamus. In: GG Somjen, editor. Neurophysiology studied in man. Amsterdam: Excerpta

Medica, 1972. pp. 221-230.

[397] Han ZS, Zhang ET, Craig AD. Nociceptive and thermoreceptive lamina I neurons are

anatomically distinct. Nat Neurosci 1998;1:218-225.

[398] Hanagasi HA, Akat S, Gurvit H, Yazici J, Emre M. Pain is common in Parkinson's disease.

Clin Neurol Neurosurg 2011;113:11-13.

[399] Hannam AG, Matthews B. Reflex jaw opening in response to stimulation of periodontal

mechanoreceptors in the cat. Arch Oral Biol 1969;14:415-419.

[400] Hannam AG, Matthews B, Yemm R. Changes in the activity of the masseter muscle

following tooth contact in man. Arch Oral Biol 1969;14:1401-1406.

[401] Hannam AG, Matthews B, Yemm R. The response of the masseter muscle following tooth

contact in man. J Physiol 1969;203:25P-26P.

256

[402] Hapak L, Gordon A, Locker D, Shandling M, Mock D, Tenenbaum HC. Differentiation

between musculoligamentous, dentoalveolar, and neurologically based craniofacial pain

with a diagnostic questionnaire. J Orofac Pain 1994;8:357-368.

[403] Hardy SG. Analgesia elicited by prefrontal stimulation. Brain Res 1985;339:281-284.

[404] Hardy SG, Haigler HJ. Prefrontal influences upon the midbrain: a possible route for pain

modulation. Brain Res 1985;339:285-293.

[405] Hardy SG, Leichnetz GR. Cortical projections to the periaqueductal gray in the monkey: a

retrograde and orthograde horseradish peroxidase study. Neurosci Lett 1981;22:97-101.

[406] Hardy SG, Leichnetz GR. Frontal cortical projections to the periaqueductal gray in the rat:

a retrograde and orthograde horseradish peroxidase study. Neurosci Lett 1981;23:13-17.

[407] Hari R, Karhu J, Hamalainen M, Knuutila J, Salonen O, Sams M, Vilkman V. Functional

organization of the human first and second somatosensory cortices: A neuromagnetic

study. Eur J Neurosci 1993;5:724-734.

[408] Harkins SW, Price DD, Braith J. Effects of ex- traversion and neuroticism on experimental

pain, clinical pain, and illness behavior. Pain 1989;36:209-218.

[409] Harman K, Ruyak P. Working through the pain: a controlled study of the impact of

persistent pain on performing a computer task. Clin J Pain 2005;21:216-222.

[410] Harmann PA, Carlton SM, Willis WD. Collaterals of the spinothalamic tract cells to the

periaqueductal gray: A fluorescent double-labeling study in the rat. Brain Res

1988;441:87-97.

[411] Harrison F, Corbin KB. The central pathway for the jaw-jerk. Am J Physiol 1941;135:439-

445.

[412] Hart RP, Wade JB, Martelli MF. Cognitive impairment in patients with chronic pain: the

significance of stress. Curr Pain Headache Rep 2003;7:116-126.

[413] Harting JK, Updyke BV, Van Lieshout DP. Corticotectal projections in the cat:

anterograde transport studies of twenty-five cortical areas. J Comp Neurol 1992;324:379-

414.

257

[414] Hassenbusch SJ, Pillay PK, Barnett GH. Radiofrequency cingulotomy for intractable

cancer pain using stereotaxis guided by magnetic resonance imaging. Neurosurgery

1990;27:220-223.

[415] Hassler R. Dichotomy of facial pain conduction in the diencephalon. In: R Hassler, AE

Walker, editors. Trigeminal neuralgia. Philadelphia: Saunders, 1970. pp. 123-138.

[416] Hatanaka N, Tokuno H, Hamada I, Inase M, Ito Y, Imanishi M, Hasegawa N, Akazawa T,

Nambu A, Takada M. Thalamocortical and intracortical connections of monkey cingulate

motor areas. Journal of Comparative Neurology 2003;462:121-138.

[417] Head H, Holmes G. Sensory disturbances from cerebral lesions. Brain 1911;34:102-254.

[418] Health NIo. Management of temporomandibular disorders. National Institutes of Health

Technology Assessment Conference Statement. J Am Dent Assoc 1996;127:1595-1606.

[419] Hebb DO. Spontaneous neurosis in chimpanzees; theoretical relations with clinical and

experimental phenomena. Psychosom Med 1947;9:3-19.

[420] Heinricher MM. The brainstem and nociceptive modulation. In: AI Basbaum, MC

Bushnell, editors. Science of Pain. San Diego: Academic Press, 2009. pp. 593-626.

[421] Hernandez-Peon R, Hagbarth KE. Interaction between afferent and cortically induced

reticular responses. J Neurophysiol 1955;18:44-55.

[422] Hihara S, Notoya T, Tanaka M, Ichinose S, Ojima H, Obayashi S, Fujii N, Iriki A.

Extension of corticocortical afferents into the anterior bank of the intraparietal sulcus by

tool-use training in adult monkeys. Neuropsychologia 2006;44:2636-2646.

[423] Hippocrates. Instruments of reduction. Book Instruments of reduction. City: D.C.

Stevenson, Web Atomics, 400 B.C.E.

[424] Hippocrates. The genuine works of Hippocrates. Book The genuine works of Hippocrates.

City: The Sydenham Society, 1849.

[425] Hirata H, Okamoto K, Bereiter DA. GABA(A) receptor activation modulates corneal unit

activity in rostral and caudal portions of trigeminal subnucleus caudalis. J Neurophysiol

2003;90:2837-2849.

258

[426] Hitzig E. Hughlings Jackson and the Cortical Motor Centres in the Ligh of Physiological

Research. Brain 1900;23:545-581.

[427] Ho KC, Roessmann U, Straumfjord JV, Monroe G. Analysis of brain weight. II. Adult

brain weight in relation to body height, weight, and surface area. Arch Pathol Lab Med

1980;104:640-645.

[428] Hoagland H. Specific afferent impulses and cutaneous sensitivity. J Gen Psychol

1932;6:276-295.

[429] Hodaie M, Quan J, Chen DQ. In vivo visualization of cranial nerve pathways in humans

using diffusion-based tractography. Neurosurgery 2010;66:788-796.

[430] Hodes RL, Howland EW, Lightfoot N, Cleeland CS. The effects of distraction on

responses to cold pressor pain. Pain 1990;41:109-114.

[431] Hodges PW, Tucker K. Moving differently in pain: a new theory to explain the adaptation

to pain. Pain 2011;152:S90-98.

[432] Hoheisel U, Reinohl J, Unger T, Mense S. Acidic pH and capsaicin activate

mechanosensitive group IV muscle receptors in the rat. Pain 2004;110:149-157.

[433] Holden JE, Van Poppel AY, Thomas S. Antinociception from lateral hypothalamic

stimulation may be mediated by NK(1) receptors in the A7 catecholamine cell group in

rat. Brain Res 2002;953:195-204.

[434] Holle D, Naegel S, Krebs S, Gaul C, Gizewski E, Diener HC, Katsarava Z, Obermann M.

Hypothalamic gray matter volume loss in hypnic headache. Ann Neurol 2011;69:533-

539.

[435] Hollins M, Harper D, Gallagher S, Owings EW, Lim PF, Miller V, Siddiqi MQ, Maixner

W. Perceived Intensity and Unpleasantness of Cutaneous and Auditory Stimuli: An

Evaluation of the Generalized Hypervigilance Hypothesis. Pain 2009;141:215-221.

[436] Holtmaat A, Wilbrecht L, Knott GW, Welker E, Svoboda K. Experience-dependent and

cell-type-specific spine growth in the neocortex. Nature 2006;441:979-983.

[437] Hosobuchi Y. Combined electrical stimulation of the periaqueductal gray matter and

sensory thalamus. Applied Neurophysiology 1983;46:112-115.

259

[438] Huettel SA, McCarthy G, Song AW. Functional magnetic resonance imaging. Sunderland,

Mass.: Sinauer Associates, 2009.

[439] Hurt RW, Ballantine HT. Stereotactic anterior cingulate lesions for persistent pain: a report

of 68 cases. Clin Neurosurg 1973;21:334-351.

[440] Hutchison WD, Davis KD, Lozano AM, Tasker RR, Dostrovsky JO. Pain-related neurons

in the human cingulate cortex. Nat Neurosci 1999;2:403-405.

[441] Hylden JL, Hayashi H, Bennett GJ. Lamina I spinomesencephalic neurons in the cat

ascend via the dorsolateral funiculi. Somatosens Res 1986;4:31-41.

[442] Iadarola JM, Caudle RM. Good pain, bad pain. Science 1997;278:239-240.

[443] Iadarola MJ, Berman KF, Zeffiro TA, Byas-Smith MG, Gracely RH, Max MB, Bennett

GJ. Neural activation during acute capsaicin-evoked pain and allodynia assessed with

PET. Brain 1998;121:931-947.

[444] Iannetti GD, Leandri M, Truini A, Zambreanu L, Cruccu G, Tracey I. Adelta nociceptor

response to laser stimuli: selective effect of stimulus duration on skin temperature, brain

potentials and pain perception. Clin Neurophysiol 2004;115:2629-2637.

[445] Iannetti GD, Mouraux A. From the neuromatrix to the pain matrix (and back). Exp Brain

Res 2010;205:1-12.

[446] Iannetti GD, Porro CA, Pantano P, Romanelli PL, Galeotti F, Cruccu G. Representation of

different trigeminal divisions within the primary and secondary human somatosensory

cortex. Neuroimage 2003;19:906-912.

[447] Iannetti GD, Zambreanu L, Cruccu G, Tracey I. Operculoinsular cortex encodes pain

intensity at the earliest stages of cortical processing as indicated by amplitude of laser-

evoked potentials in humans. Neuroscience 2005;131:199-208.

[448] Ichesco E, Quintero A, Clauw DJ, Peltier S, Sundgren PM, Gerstner GE, Schmidt-Wilcke

T. Altered Functional Connectivity Between the Insula and the Cingulate Cortex in

Patients With Temporomandibular Disorder: A Pilot Study. Headache 2011;doi:

10.1111/j.1526-4610.2011.01998.x. [Epub ahead of print].

[449] Iggo A, Muir AR. The structure and function of a slowly adapting touch corpuscle in hairy

skin. J Physiol 1969;200:763-796.

260

[450] Ikeda M, Tanami T, Matsushita M. Ascending and descending internuclear connections of

the trigeminal sensory nuclei in the cat. A study with the retrograde and anterograde

horseradish peroxidase technique. Neuroscience 1984;12:1243-1260.

[451] Iwata K, Kamo H, Ogawa A, Tsuboi Y, Noma N, Mitsuhashi Y, Taira M, Koshikawa N,

Kitagawa J. Anterior cingulate cortical neuronal activity during perception of noxious

thermal stimuli in monkeys. J Neurophysiol 2005;94:1980-1991.

[452] Jacquin MF, Chiaia NL, Haring JH, Rhoades RW. Intersubnuclear connections within the

rat trigeminal brainstem complex. Somatosens Mot Res 1990;7:399-420.

[453] Jacquin MF, Renehan WE, Mooney RD, Rhoades RW. Structure-function relationships in

rat medullary and cervical dorsal horns. I. Trigeminal primary afferents. J Neurophysiol

1986;55:1153-1186.

[454] Jain N, Florence SL, Qi HX, Kaas JH. Growth of new brainstem connections in adult

monkeys with massive sensory loss. Proc Natl Acad Sci U S A 2000;97:5546-5550.

[455] Jamison RN, Sbrocco T, Parris WC. The influence of problems with concentration and

memory on emotional distress and daily activities in chronic pain patients. Int J

Psychiatry Med 1988;18:183-191.

[456] Jansons KM, Alexander DC. Persistent Angular Structure: new insights from diffusion

MRI data. Dummy version. Inf Process Med Imaging 2003;18:672-683.

[457] Jenkins IH, Brooks DJ, Nixon PD, Frackowiak RSJ, Passingham RE. Motor sequence

learning: A study with positron emission tomography. J Neurosci 1994;14:3775-3790.

[458] Jenkins WM, Merzenich MM. Reorganization of neocortical representations after brain

injury: a neurophysiological model of the bases of recovery from stroke. Prog Brain Res

1987;71:249-266.

[459] Jenkinson M, Bannister P, Brady M, Smith S. Improved optimization for the robust and

accurate linear registration and motion correction of brain images. Neuroimage

2002;17:825-841.

[460] Ji G, Neugebauer V. Pain-related deactivation of medial prefrontal cortical neurons

involves mGluR1 and GABAA receptors. J Neurophysiol 2011.

261

[461] Ji G, Sun H, Fu Y, Li Z, Pais-Vieira M, Galhardo V, Neugebauer V. Cognitive impairment

in pain through amygdala-driven prefrontal cortical deactivation. J Neurosci

2010;30:5451-5464.

[462] Johnson KO, Darian-Smith I, LaMotte C, Johnson B, Oldfield S. Coding of incremental

changes in skin temperature by a population of warm fibers in the monkey: correlation

with intensity discrimination in man. J Neurophysiol 1979;42:1332-1353.

[463] Johnson S, Summers J, Pridmore S. Changes to somatosensory detection and pain

thresholds following high frequency repetitive TMS of the motor cortex in individuals

suffering from chronic pain. Pain 2006;123:187-192.

[464] Jones AKP, Brown WD, Friston KJ, Qi LY, Frackowiak RSJ. Cortical and subcortical

localization of response to pain in man using positron emission tomography. Proc R Soc

Lond B: Biol Sci 1991;244:39-44.

[465] Jones DK. Studying connections in the living human brain with diffusion MRI. Cortex

2008;44:936-952.

[466] Jones DK, Symms MR, Cercignani M, Howard RJ. The effect of filter size on VBM

analyses of DT-MRI data. Neuroimage 2005;26:546-554.

[467] Jones EG, Pons TP. Thalamic and brainstem contributions to large-scale plasticity of

primate somatosensory cortex. Science 1998;282:1121-1125.

[468] Jones SL. Descending noradrenergic influences on pain. Prog Brain Res 1991;88:381-394.

[469] Jones SL, Sedivec MJ, Light AR. Effects of iontophoresed opioids on physiologically

characterized laminae I and II dorsal horn neurons in the cat spinal cord. Brain Res

1990;532:160-174.

[470] Julius D, Basbaum AI. Molecular mechanisms of nociception. Nature 2001;413:203-210.

[471] Juottonen K, Gockel M, SilÇn T, Hurri H, Hari R, Forss N. Altered central sensorimotor

processing in patients with complex regional pain syndrome. Pain 2002;98:315-323.

[472] Kaas JH. Sensory loss and cortical reorganization in mature primates. Prog Brain Res

2002;138:167-176.

262

[473] Kaas JH, Florence SL. Mechanisms of reorganization in sensory systems of primates after

peripheral nerve injury. Adv Neurol 1997;73:147-158.

[474] Kaas JH, Florence SL, Jain N. Subcortical contributions to massive cortical

reorganizations. Neuron 1999;22:657-660.

[475] Kaas JH, Qi HX, Burish MJ, Gharbawie OA, Onifer SM, Massey JM. Cortical and

subcortical plasticity in the brains of humans, primates, and rats after damage to sensory

afferents in the dorsal columns of the spinal cord. Exp Neurol 2008;209:407-416.

[476] Kanda M, Mima T, Oga T, Matsuhashi M, Toma K, Hara H, Satow T, Nagamine T,

Rothwell JC, Shibasaki H. Transcranial magnetic stimulation (TMS) of the sensorimotor

cortex and medial frontal cortex modifies human pain perception. Clin Neurophysiol

2003;114:860-866.

[477] Kandel ER, Spencer WA. Cellular neurophysiological approaches in the study of learning.

Physiol Rev 1968;48:65-134.

[478] Karni A, Meyer G, Jezzard P, Adams MM, Turner R, Ungerleider LG. Functional MRI

evidence for adult motor cortex plasticity during motor skill learning. Nature

1995;377:155-158.

[479] Kassubek J, Juengling FD, Els T, Spreer J, Herpers M, Krause T, Moser E, Lucking CH.

Activation of a residual cortical network during painful stimulation in long-term

postanoxic vegetative state: a 15O-H2O PET study. Journal of the Neurological Sciences

2003;212:85-91.

[480] Kaufman MP, Rybicki KJ, Waldrop TG. Effect of ischemia on responses of group III and

IV afferents. J Appl Physiol 1984;57:644-650.

[481] Keay KA, Feil K, Gordon BD, Herbert H, Bandler R. Spinal afferents to functionally

distinct periaqueductal gray columns in the rat: an anterograde and retrograde tracing

study. J Comp Neurol 1997;385:207-229.

[482] Kellgren JH. Referred Pains from Muscle. Br Med J 1938;1:325-327.

[483] Kenins P. The functional anatomy of the receptive fields of rabbit C polymodal

nociceptors. J Neurophysiol 1988;59:1098-1115.

263

[484] Kenshalo DR, Iwata K, Sholas M, Thomas DA. Response properties and organization of

nociceptive neurons in area 1 of monkey primary somatosensory cortex. J Neurophysiol

2000;84:719-729.

[485] Kenshalo DR, Jr., Giesler GJ, Jr., Leonard RB, Willis WD. Responses of neurons in

primate ventral posterior lateral nucleus to noxious stimuli. J Neurophysiol

1980;43:1594-1614.

[486] Kenshalo DR, Jr., Isensee O. Responses of primate SI cortical neurons to noxious stimuli. J

Neurophysiol 1983;50:1479-1496.

[487] Kerr FW. The Divisional Organization of Afferent Fibres of the Trigeminal Nerve. Brain

1963;86:721-732.

[488] Kerr FWL, Lippman HH. The primate spinothalamic tract as demonstrated by anterolateral

cordotomy and commissural myelotomy. Adv Neurol 1974;4:147-156.

[489] Kewman DG, Vaishampayan N, Zald D, Han B. Cognitive impairment in musculoskeletal

pain patients. Int J Psychiatry Med 1991;21:253-262.

[490] Kim JH, Suh SI, Seol HY, Oh K, Seo WK, Yu SW, Park KW, Koh SB. Regional grey

matter changes in patients with migraine: a voxel-based morphometry study. Cephalalgia

2008;28:598-604.

[491] King CD, Wong F, Currie T, Mauderli AP, Fillingim RB, Riley JL, III. Deficiency in

endogenous modulation of prolonged heat pain in patients with Irritable Bowel Syndrome

and Temporomandibular Disorder. Pain 2009;143:172-178.

[492] Kirveskari E, Vartiainen NV, Gockel M, Forss N. Motor cortex dysfunction in complex

regional pain syndrome. Clin Neurophysiol 2010;121:1085-1092.

[493] Knott GW, Holtmaat A, Wilbrecht L, Welker E, Svoboda K. Spine growth precedes

synapse formation in the adult neocortex in vivo. Nat Neurosci 2006;9:1117-1124.

[494] Koay CG, Nevo U, Chang LC, Pierpaoli C, Basser PJ. The elliptical cone of uncertainty

and its normalized measures in diffusion tensor imaging. IEEE transactions on medical

imaging 2008;27:834-846.

[495] Koechlin E. Frontal pole function: what is specifically human? Trends Cogn Sci

2011;15:241; author reply 243.

264

[496] Koechlin E, Corrado G, Pietrini P, Grafman J. Dissociating the role of the medial and

lateral anterior prefrontal cortex in human planning. Proc Natl Acad Sci U S A

2000;97:7651-7656.

[497] Koechlin E, Hyafil A. Anterior Prefrontal Function and the Limits of Human Decision-

Making. Science 2007;318:594-600.

[498] Koechlin E, Ody C, Kouneiher F. The architecture of cognitive control in the human

prefrontal cortex. Science 2003;302:1181-1185.

[499] Koyama T, Kato K, Tanaka YZ, Mikami A. Anterior cingulate activity during pain-

avoidance and reward tasks in monkeys. Neurosci Res 2001;39:421-430.

[500] Kropotov JD, Crawford HJ, Polyakov YI. Somatosensory event-related potential changes

to painful stimuli during hypnotic analgesia: Anterior cingulate cortex and anterior

temporal cortex intracranial recordings. Int J Psychophysiol 1997;27:1-8.

[501] Kruger L, Saporta S, Feldman SG. Axonal transport studies of the sensory trigeminal

complex. In: DJ Anderson, B Matthews, editors. Pain in the Trigeminal Region. New

York: Elsevier, 1977. pp. 191-201.

[502] Kuchinad A, Schweinhardt P, Seminowicz DA, Wood PB, Chizh BA, Bushnell MC.

Accelerated brain gray matter loss in fibromyalgia patients: premature aging of the brain?

J Neurosci 2007;27:4004-4007.

[503] Kulkarni B, Bentley DE, Elliott R, Youell P, Watson A, Derbyshire SW, Frackowiak RS,

Friston KJ, Jones AK. Attention to pain localization and unpleasantness discriminates the

functions of the medial and lateral pain systems. Eur J Neurosci 2005;21:3133-3142.

[504] Kumazawa T, Perl ER. Primate cutaneous receptors with unmyelinated (C) fibres and their

projection to the substantia gelatinosa. J Physiol 1977;73:287-304.

[505] Kupers RC, Svensson P, Jensen TS. Central representation of muscle pain and mechanical

hyperesthesia in the orofacial region: a positron emission tomography study. Pain

2004;108:284-293.

[506] Kurth F, Zilles K, Fox PT, Laird AR, Eickhoff S. A link between the systems: functional

differentiation and integration within the human insula revealed by meta-analysis. Brain

Struct Funct 2010;214:519-534.

265

[507] Kuypers HG. The Descending Pathways to the Spinal Cord, Their Anatomy and Function.

Prog Brain Res 1964;11:178-202.

[508] Kwan CL, Crawley AP, Mikulis DJ, Davis KD. An fMRI study of the anterior cingulate

cortex and surrounding medial wall activations evoked by noxious cutaneous heat and

cold stimuli. Pain 2000;85:359-374.

[509] Kwan CL, Diamant NE, Pope G, Mikula K, Mikulis DJ, Davis KD. Abnormal forebrain

activity in functional bowel disorder patients with chronic pain. Neurology

2005;65:1268-1277.

[510] Lacadie CM, Fulbright RK, Rajeevan N, Constable RT, Papademetris X. More accurate

Talairach coordinates for neuroimaging using non-linear registration. Neuroimage

2008;42:717-725.

[511] Lamm C, Decety J, Singer T. Meta-analytic evidence for common and distinct neural

networks associated with directly experienced pain and empathy for pain. Neuroimage

2011;54:2492-2502.

[512] LaMotte RH, Shain CN, Simone DA, Tsai EFP. Neurogenic hyperalgesia: Psychophysical

studies of underlying mechanisms. J Neurophysiol 1991;66:190-211.

[513] Lancaster JL, Tordesillas-Gutierrez D, Martinez M, Salinas F, Evans A, Zilles K,

Mazziotta JC, Fox PT. Bias between MNI and Talairach coordinates analyzed using the

ICBM-152 brain template. Hum Brain Mapp 2007;28:1194-1205.

[514] Langworthy OR, Fox HM. Thalamic Syndrome: Syndrom of the posterior cerebral artery;

a review. Arch Intern Med 1937;60:203-224.

[515] Latremoliere A, Woolf CJ. Central sensitization: a generator of pain hypersensitivity by

central neural plasticity. J Pain 2009;10:895-927.

[516] Laureys S, Faymonville ME, Peigneux P, Damas P, Lambermont B, Del Fiore G,

Degueldre C, Aerts J, Luxen A, Franck G, Lamy M, Moonen G, Maquet P. Cortical

processing of noxious somatosensory stimuli in the persistent vegetative state.

Neuroimage 2002;17:732-741.

[517] Laureys S, Owen AM, Schiff ND. Brain function in coma, vegetative state, and related

disorders. Lancet Neurol 2004;3:537-546.

266

[518] Lazar M, Alexander AL. An error analysis of white matter tractography methods: synthetic

diffusion tensor field simulations. Neuroimage 2003;20:1140-1153.

[519] Le Bars D. The whole body receptive field of dorsal horn multireceptive neurones. Brain

Res Brain Res Rev 2002;40:29-44.

[520] Le Bars D, Dickenson AH, Besson JM. Diffuse noxious inhibitory controls (DNIC). I.

Effects on dorsal horn convergent neurones in the rat. Pain 1979;6:283-304.

[521] Le Bars D, Willer JC. Diffuse noxious inhibitory controls (DNIC). In: AI Basbaum, MC

Bushnell, editors. Science of Pain. San Diego: Academic Press, 2009. pp. 763-773.

[522] Le Bihan D. Magnetic resonance diffusion imaging: introduction and concepts. In: DK

Jones, editor. Diffusion MRI. New York: Oxford University Press, 2011. pp. 57-78.

[523] Le Bihan D, Breton E, Lallemand D, Grenier P, Cabanis E, Laval-Jeantet M. MR imaging

of intravoxel incoherent motions: application to diffusion and perfusion in neurologic

disorders. Radiology 1986;161:401-407.

[524] Le Bihan D, Mangin JF, Poupon C, Clark CA, Pappata S, Molko N, Chabriat H. Diffusion

tensor imaging: concepts and applications. J Magn Reson Imaging 2001;13:534-546.

[525] Leal PR, Roch JA, Hermier M, Souza MA, Cristino-Filho G, Sindou M. Structural

abnormalities of the trigeminal root revealed by diffusion tensor imaging in patients with

trigeminal neuralgia caused by neurovascular compression: a prospective, double-blind,

controlled study. Pain 2011;152:2357-2364.

[526] Lee DM, Pendleton N, Tajar A, O'Neill TW, O'Connor DB, Bartfai G, Boonen S,

Casanueva FF, Finn JD, Forti G, Giwercman A, Han TS, Huhtaniemi IT, Kula K, Lean

ME, Punab M, Silman AJ, Vanderschueren D, Moseley CM, Wu FC, McBeth J. Chronic

widespread pain is associated with slower cognitive processing speed in middle-aged and

older European men. Pain 2010;151:30-36.

[527] Lee J, Dougherty PM, Antezana D, Lenz FA. Responses of neurons in the region of human

thalamic principal somatic sensory nucleus to mechanical and thermal stimuli graded into

the painful range. J Comp Neurol 1999;410:541-555.

[528] Lee MC, Mouraux A, Iannetti GD. Characterizing the cortical activity through which pain

emerges from nociception. J Neurosci 2009;29:7909-7916.

267

[529] Legrain V, Damme SV, Eccleston C, Davis KD, Seminowicz DA, Crombez G. A

neurocognitive model of attention to pain: behavioral and neuroimaging evidence. Pain

2009;144:230-232.

[530] Legrain V, Iannetti GD, Plaghki L, Mouraux A. The pain matrix reloaded: a salience

detection system for the body. Prog Neurobiol 2011;93:111-124.

[531] Lele PP, Sinclair DC, Weddell G. The reaction time to touch. J Physiol 1954;123:187-203.

[532] Lenz FA, Rios M, Chau D, Krauss GL, Zirh TA, Lesser RP. Painful stimuli evoke

potentials recorded from the parasylvian cortex in humans. J Neurophysiol 1998;80:2077-

2088.

[533] Lenz FA, Rios M, Zirh A, Chau D, Krauss G, Lesser RP. Painful stimuli evoke potentials

recorded over the human anterior cingulate gyrus. J Neurophysiol 1998;79:2231-2234.

[534] Lenz FA, Seike M, Lin YC, Baker FH, Rowland LH, Gracely RH, Richardson RT.

Neurons in the area of human thalamic nucleus ventralis caudalis respond to painful heat

stimuli. Brain Res 1993;623:235-240.

[535] Leo RJ, Latif T. Repetitive transcranial magnetic stimulation (rTMS) in experimentally

induced and chronic neuropathic pain: a review. J Pain 2007;8:453-459.

[536] Leonard CM. The prefrontal cortex of the rat. I. Cortical projection of the mediodorsal

nucleus. II. Efferent connections. Brain Res 1969;12:321-343.

[537] Lerch JP. Maze training in mice induces MRI-detectable brain shape changes specific to

the type of learning. Neuroimage 2011;54:2086-2095.

[538] Lerch JP, Evans AC. Cortical thickness analysis examined through power analysis and a

population simulation. Neuroimage 2005;24:163-173.

[539] Leresche L. Epidemiology of temporomandibular disorders: implications for the

investigation of etiologic factors. Crit Rev Oral Biol Med 1997;8:291-305.

[540] LeResche L, Dworkin SF, Sommers EE, Truelove EL. An epidemiologic evaluation of two

diagnostic classification schemes for temporomandibular disorders. J Prosthet Dent

1991;65:131-137.

268

[541] Lev R, Granovsky Y, Yarnitsky D. Orbitofrontal disinhibition of pain in migraine with

aura: An interictal EEG-mapping study. Cephalalgia 2010;30:910-918.

[542] Lewin W, Phillips CG. Observations on partial removal of the post-central gyrus for pain. J

Neurol Neurosurg Psychiatry 1952;15:143-147.

[543] Lewis T, Pochin E. The double pain response of the human skin to a single stimulus. Clin

Sci 1937;3:67-76.

[544] Lidsky TI, Labuszewski T, Avitable MJ, Robinson JH. The effects of stimulation of

trigeminal sensory afferents upon caudate units in cats. Brain Res Bull 1979;4:9-14.

[545] Lima D. Ascending pathways: anatomy and physiology. In: A Basbaum, MC Bushnell,

editors. Science of Pain. Slovenia: Academic Press, 2009. pp. 477-526.

[546] Lima D, Almeida A. The medullary dorsal reticular nucleus as a pronociceptive centre of

the pain control system. Prog Neurobiol 2002;66:81-108.

[547] Lima MC, Fregni F. Motor cortex stimulation for chronic pain. Neurology 2008;70:2329-

2338.

[548] Lima MC, Fregni F. Motor cortex stimulation for chronic pain: systematic review and

meta-analysis of the literature. Neurology 2008;70:2329-2337.

[549] Linden RW. Properties of intraoral mechanoreceptors represented in the mesencephalic

nucleus of the fifth nerve in the cat. J Physiol 1978;279:395-408.

[550] Lineberry CG, Vierck CJ. Attenuation of pain reactivity by caudate nucleus stimulation in

monkeys. Brain Res 1975;98:119-134.

[551] Lorenz J, Cross DJ, Minoshima S, Morrow TJ, Paulson PE, Casey KL. A unique

representation of heat allodynia in the human brain. Neuron 2002;35:383-393.

[552] Lorenz J, Minoshima S, Casey KL. Keeping pain out of mind: the role of the dorsolateral

prefrontal cortex in pain modulation. Brain 2003;126:1079-1091.

[553] Luerding R, Weigand T, Bogdahn U, Schmidt-Wilcke T. Working memory performance is

correlated with local brain morphology in the medial frontal and anterior cingulate cortex

in fibromyalgia patients: structural correlates of pain-cognition interaction. Brain

2008;131:3222-3231.

269

[554] Lugaro E. Modern problems in psychiatry. Translated by David Orr, and R. G. Rows,

Manchester: The University Press, 1909.

[555] Lund JP. Mastication and its control by the brain stem. Crit Rev Oral Biol Med 1991;2:33-

64.

[556] Lund JP, Donga R, Widmer CG, Stohler CS. The pain-adaptation model: a discussion of

the relationship between chronic musculoskeletal pain and motor activity. Can J Physiol

Pharmacol 1991;69:683-694.

[557] Luo P, Wong R, Dessem D. Ultrastructural basis for synaptic transmission between jaw-

muscle spindle afferents and trigeminothalamic neurons in the rostral trigeminal sensory

nuclei of the rat. J Comp Neurol 1995;363:109-128.

[558] Lutz J, Jager L, de Quervain D, Krauseneck T, Padberg F, Wichnalek M, Beyer A, Stahl R,

Zirngibl B, Morhard D, Reiser M, Schelling G. White and gray matter abnormalities in

the brain of patients with fibromyalgia: a diffusion-tensor and volumetric imaging study.

Arthritis Rheum 2008;58:3960-3969.

[559] MacDonald D, Kabani N, Avis D, Evans AC. Automated 3-D extraction of inner and outer

surfaces of cerebral cortex from MRI. Neuroimage 2000;12:340-356.

[560] Mainero C, Boshyan J, Hadjikhani N. Altered functional magnetic resonance imaging

resting-state connectivity in periaqueductal gray networks in migraine. Ann Neurol

2011;70:838-845.

[561] Maixner W. Biopsychosocial and Genetic Risk Factors for Temporomandibular Joint

Disorders and Related Conditions. In: T Graven-Nielsen, L Arendt-Nielsen, S Mense,

editors. Fundamentals of Muskuloskeletal Pain. Seattle: IASP Press, 2008.

[562] Maixner W. Temporomandibular Joint Disorders. In: EA Mayer, MC Bushnell, editors.

Functional Pain Syndromes. Seattle: IASP Press, 2009.

[563] Maixner W, Fillingim R, Booker D, Sigurdsson A. Sensitivity of patients with painful

temporomandibular disorders to experimentally evoked pain. Pain 1995;63:341-351.

[564] Maixner W, Fillingim R, Kincaid S, Sigurdsson A, Harris MB. Relationship between pain

sensitivity and resting arterial blood pressure in patients with painful temporomandibular

disorders. Psychosom Med 1997;59:503-511.

270

[565] Maixner W, Fillingim R, Sigurdsson A, Kincaid S, Silva S. Sensitivity of patients with

painful temporomandibular disorders to experimentally evoked pain: evidence for altered

temporal summation of pain. Pain 1998;76:71-81.

[566] Manning BH, Morgan MJ, Franklin KB. Morphine analgesia in the formalin test: evidence

for forebrain and midbrain sites of action. Neuroscience 1994;63:289-294.

[567] Mantyh PW. The ascending input to the midbrain periaqueductal gray of the primate. J

Comp Neurol 1982;211:50-64.

[568] Marchand F, Perretti M, McMahon SB. Role of the immune system in chronic pain. Nat

Rev Neurosci 2005;6:521-532.

[569] Marfurt CF. The somatotopic organization of the cat trigeminal ganglion as determined by

the horseradish peroxidase technique. Anat Rec 1981;201:105-118.

[570] Marfurt CF, Turner DF. Sensory nerve endings in the rat oro-facial region labeled by the

anterograde and transganglionic transport of horseradish peroxidase: a new method for

tracing peripheral nerve fibers. Brain Res 1983;261:1-12.

[571] Marshall J. Sensory disturbances in cortical wounds with special reference to pain. J

Neurol Neurosurg Psychiatry 1951;14:187-204.

[572] Martens MA, Moeyersoons JP. Acute and recurrent effort-related compartment syndrome

in sports. Sports Med 1990;9:62-68.

[573] Martin JH, Jessell TM. Modality coding in the somatic sensory system. In: ER Kandel, JH

Schwartz, TM Jessell, editors. Principles of Neural Science. Norwalk, Connecticut:

Appleton & Lange, 1991. pp. 341-352.

[574] Matsushita M, Ikeda M, Okado N. The cells of origin of the trigeminothalamic,

trigeminospinal and trigeminocerebellar projections in the cat. Neuroscience

1982;7:1439-1454.

[575] May A. Chronic pain may change the structure of the brain. Pain 2008;137:7-15.

[576] May A. Experience-dependent structural plasticity in the adult human brain. Trends Cogn

Sci 2011;15:475-482.

271

[577] May A. Structural brain imaging: a window into chronic pain. Neuroscientist 2011;17:209-

220.

[578] May A, Ashburner J, Buchel C, McGonigle DJ, Friston KJ, Frackowiak RS, Goadsby PJ.

Correlation between structural and functional changes in brain in an idiopathic headache

syndrome. Nat Med 1999;5:836-838.

[579] Mayberg HS, Brannan SK, Mahurin RK, Jerabek PA, Brickman JS, Tekell JL, Silva JA,

McGinnis S, Glass TG, Martin CC, Fox PT. Cingulate function in depression: A potential

predictor of treatment response. Neuroreport 1997;8:1057-1061.

[580] Mayer DJ, Price DD. Central nervous system mechanisms of analgesia. Pain 1976;2:379-

404.

[581] Mayer EA, Bushnell MC. Synthesis. In: EA Mayer, MC Bushnell, editors. Functional Pain

Syndromes. Seattle: IASP Press, 2009.

[582] Mayer EA, Labus JS, Berkley KJ. Sex differences in pain. In: JB Becker, KJ Berkley, N

Geary, E Hampson, JP Herman, EA Young, editors. Sex Differences in the Brain from

Genes to Behavior. Oxford: Oxford University Press, 2008. pp. 371-396.

[583] Mazzola L, Isnard J, Mauguiere F. Somatosensory and pain responses to stimulation of the

second somatosensory area (SII) in humans. A comparison with SI and insular responses.

Cereb Cortex 2006;16:960-968.

[584] McCaul KD, Haugtvedt C. Attention, distraction, and cold-pressor pain. Journal of

Personality and Social Psychology 1982;43:154-162.

[585] McGinnis SM, Brickhouse M, Pascual B, Dickerson BC. Age-Related Changes in the

Thickness of Cortical Zones in Humans. Brain Topogr 2011.

[586] McHaffie JG, Larson MA, Stein BE. Response properties of nociceptive and low-threshold

neurons in rat trigeminal pars caudalis. Journal of Comparative Neurology 1994;347:409-

425.

[587] McNeill C. History and evolution of TMD concepts. Oral Surg Oral Med Oral Pathol Oral

Radiol Endod 1997;83:51-60.

[588] McRae K, Hughes B, Chopra S, Gabrieli JD, Gross JJ, Ochsner KN. The neural bases of

distraction and reappraisal. J Cogn Neurosci 2010;22:248-262.

272

[589] Mehler WR. Some neurological species differences - a posteriori. Ann N Y Acad Sci

1969;167:424-468.

[590] Meier W, Klucken M, Soyka D, Bromm B. Hypnotic hypo- and hyperalgesia: Divergent

effects on pain ratings and pain-related cerebral potentials. Pain 1993;53:175-181.

[591] Melzack R. The McGill pain questionnaire: Major properties and scoring methods. Pain

1975;1:277-299.

[592] Melzack R. Pain - an overview. Acta Anaesthesiol Scand 1999;43:880-884.

[593] Melzack R, Casey KL. Sensory, motivational, and central control determinants of pain: a

new conceptual model. In: D Kenshalo, editor. The Skin Senses. Springfield: C.C.

Thomas, 1968. pp. 423-439.

[594] Melzack R, Loeser JD. Phantom body pain in paraplegics: evidence for a central "pattern

generating mechanism" for pain. Pain 1978;4:195-210.

[595] Melzack R, Wall PD. Pain mechanisms: a new theory. Science 1965;150:971-979.

[596] Menetrey D, Chaouch A, Binder D, Besson JM. The origin of the spinomesencephalic tract

in the rat: an anatomical study using the retrograde transport of horseradish peroxidase. J

Comp Neurol 1982;206:193-207.

[597] Meng ID, Hu JW, Benetti AP, Bereiter DA. Encoding of corneal input in two distinct

regions of the spinal trigeminal nucleus in the rat: cutaneous receptive field properties,

responses to thermal and chemical stimulation, modulation by diffuse noxious inhibitory

controls, and projections to the parabrachial area. J Neurophysiol 1997;77:43-56.

[598] Mense S, Hoheisel U. Morphology and functional types of muscle nociceptors. In: T

Graven-Nielsen, L Arendt-Nielsen, S Mense, editors. Fundamentals of Muskuloskeletal

Pain. Seattle: IASP Press, 2008. pp. 3-18.

[599] Mense S, Meyer H. Different types of slowly conducting afferent units in cat skeletal

muscle and tendon. J Physiol 1985;363:403-417.

[600] Merskey H. Classification of chronic pain. Descriptions of chronic pain syndromes and

definitions of pain terms. Seattle: IASP Press, 1994.

273

[601] Merzenich MM, Kaas JH, Wall J, Nelson RJ, Sur M, Fellman D. Topographic

reorganization of somatosensory cortical areas 3B and 1 in adult monkeys following

restricted deafferentation. Neuroscience 1983;8:33-55.

[602] Merzenich MM, Nelson RJ, Stryker MP, Cynader MS, Schoppmann A, Zook JM.

Somatosensory cortical map changes following digit amputation in adult monkeys. J

Comp Neurol 1984;224:591-605.

[603] Mesulam MM. Tracing neural connections of human brain with selective silver

impregnation. Observations on geniculocalcarine, spinothalamic, and entorhinal

pathways. Arch Neurol 1979;36:814-818.

[604] Mesulam MM. The functional anatomy and hemispheric specialization for directed

attention. The role of the parietal lobe and its connectivity. Trends Neurosci 1983;6:384-

387.

[605] Mesulam MM. Large-scale neurocognitive networks and distributed processing for

attention, language, and memory. Ann Neurol 1990;28:597-613.

[606] Mesulam MM. Spatial attention and neglect: parietal, frontal and cingulate contributions to

the mental representation and attentional targeting of salient extrapersonal events. Phil

Trans R Soc B 1999;354:1325-1346.

[607] Mesulam MM. Foreword. In: JD Schmahmann, DN Pandya, editors. Fiber pathways of the

brain. New York: Oxford University Press, 2006. pp. ix-x.

[608] Metz AE, Yau HJ, Centeno MV, Apkarian AV, Martina M. Morphological and functional

reorganization of rat medial prefrontal cortex in neuropathic pain. Proc Natl Acad Sci U S

A 2009;106:2423-2428.

[609] Meyer RA, Campbell JA. Myelinated nociceptive afferents account for the hyperalgesia

that follows a burn to the hand. Science 1981;213:1527-1529.

[610] Meyer RA, Campbell JN, Raja SN. Peripheral neural mechanisms of nociception. In: PD

Wall, R Melzack, editors. Textbook of pain, Vol. 3. New York: Churchill Livingstone,

1994. pp. 13-44.

[611] Meyer RA, Davis KD, Cohen RH, Treede RD, Campbell JN. Mechanically insensitive

afferents (MIAs) in cutaneous nerves of monkey. Brain Res 1991;561:252-261.

274

[612] Meyer RA, Ringkamp M, Campbell CM, Raja SN. Peripheral mechanisms of cutaneous

nociception. In: S McMahon, M Koltzenburg, editors. Wall and Melzack's Textbook of

Pain. Edinburgh, UK: Churchill Livingstone, 2006. pp. 3-35.

[613] Millan MJ. Descending control of pain. Progress in Neurobiology 2002;66:355-474.

[614] Milligan ED, Watkins LR. Pathological and protective roles of glia in chronic pain. Nat

Rev Neurosci 2009;10:23-36.

[615] Moayedi M, Desouza D, Erpelding N. Making sense of gray matter abnormalities in

chronic orofacial pain--synthesizing divergent findings. J Neurosci 2011;31:12396-

12397.

[616] Moayedi M, Weissman-Fogel I. Is the insula the "how much" intensity coder? J

Neurophysiol 2009;102:1345-1347.

[617] Moayedi M, Weissman-Fogel I, Crawley AP, Goldberg MB, Freeman BV, Tenenbaum

HC, Davis KD. Contribution of chronic pain and neuroticism to abnormal forebrain gray

matter in patients with temporomandibular disorder. Neuroimage 2011;55:277-286.

[618] Mogami H, Kuroda K, Hayakawa T, Akagi K. Ascending paths from the spinal trigeminal

nucleus and its adjacent structures. In: R Dubner, editor. Orofacial Sensory and Motor

Mechanisms. New York: Appleton-Century-Croft, 1971. pp. 205-228.

[619] Monakow KH, Akert K, Kunzle H. Projections of precentral and premotor cortex to the red

nucleus and other midbrain areas in Macaca fascicularis. Exp Brain Res 1979;34:91-105.

[620] Monconduit L, Bourgeais L, Bernard JF, Le Bars D, Villanueva L. Ventromedial thalamic

neurons convey nociceptive signals from the whole body surface to the dorsolateral

neocortex. J Neurosci 1999;19:9063-9072.

[621] Monconduit L, Desbois C, Villanueva L. The integrative role of the rat medullary

subnucleus reticularis dorsalis in nociception. Eur J Neurosci 2002;16:937-944.

[622] Morecraft RJ, Tanji J. Cingulofrontal interactions and the cingulate motor areas. In: BA

Vogt, editor. Cingulate neurobiology and disease. New York: Oxford University Press,

2009. pp. 114-144.

275

[623] Morecraft RJ, Van Hoesen GW. Cingulate input to the primary and supplementary motor

cortices in the rhesus monkey: evidence for somatotopy in areas 24c and 23c. J Comp

Neurol 1992;322:471-489.

[624] Moreno CB, Hernandez-Beltran N, Munevar D, Gutierrez-Alvarez AM. Central

neuropathic pain in Parkinson's disease. Neurologia 2011.

[625] Mori S, Wakana S, van Zijl PC, Nagae-Poetscher LM. MRI Atlas of Human White Matter.

Amsterdam: Elsevier Science, 2005.

[626] Mori S, Zhang J. Principles of diffusion tensor imaging and its applications to basic

neuroscience research. Neuron 2006;51:527-539.

[627] Morrison JH, Hof PR. Life and death of neurons in the aging cerebral cortex. Int Rev

Neurobiol 2007;81:17.

[628] Moss RA, Villarosa GA, Cooley JE, Lombardo TW. Masticatory muscle activity as a

function of parafunctional, active and passive oral behavioural patterns. J Oral Rehabil

1987;14:361-370.

[629] Moulton EA, Keaser ML, Gullapalli RP, Greenspan JD. Regional intensive and temporal

patterns of functional MRI activation distinguishing noxious and innocuous contact heat.

J Neurophysiol 2005;93:2183-2193.

[630] Moulton EA, Pendse G, Morris S, Aiello-Lammens M, Becerra L, Borsook D. Segmentally

arranged somatotopy within the face representation of human primary somatosensory

cortex. Hum Brain Mapp 2009;30:757-765.

[631] Mountcastle VB, Powell TP. Neural mechanisms subserving cutaneous sensibility, with

special reference to the role of afferent inhibition in sensory perception and

discrimination. Bull Johns Hopkins Hosp 1959;105:201-232.

[632] Mouraux A, Diukova A, Lee MC, Wise RG, Iannetti GD. A multisensory investigation of

the functional significance of the "pain matrix". Neuroimage 2011;54:2237-2249.

[633] Mouraux A, Iannetti GD. Nociceptive laser-evoked brain potentials do not reflect

nociceptive-specific neural activity. J Neurophysiol 2009;101:3258-3269.

[634] Murray GM, Peck CC. Orofacial pain and jaw muscle activity: A new model. J Orofac

Pain 2007;21:263-278.

276

[635] Mylius V, Brebbermann J, Dohmann H, Engau I, Oertel WH, Moller JC. Pain sensitivity

and clinical progression in Parkinson's disease. Mov Disord 2011;26:2220-2225.

[636] Nafe JP. A quatitative theory of feeling. J Gen Psychol 1929;2:199-211.

[637] Naliboff BD, Rhudy JL. Anxiety and functional pain disorders. In: EA Mayer, MC

Bushnell, editors. Functional Pain Syndromes: Presentation and Pathophysiology. Seattle:

IASP Press, 2009. pp. 185-213.

[638] Nash PG, Macefield VG, Klineberg IJ, Gustin SM, Murray GM, Henderson LA. Changes

in human primary motor cortex activity during acute cutaneous and muscle orofacial

pain. J Orofac Pain 2010;24:379-390.

[639] Nasution ID, Shigenaga Y. Ascending and descending internuclear projections within the

trigeminal sensory nuclear complex. Brain Res 1987;425:234-247.

[640] Nauta WJ, Mehler WR. Projections of the lentiform nucleus in the monkey. Brain Res

1966;1:3-42.

[641] Nebel MB, Folger S, Tommerdahl M, Hollins M, McGlone F, Essick GK.

Temporomandibular disorder modifies cortical responses to tactile stimulation. J Pain

2010;11:1083-1094.

[642] Newman HM, Stevens RT, Apkarian AV. Direct spinal projections to limbic and striatal

areas: anterograde transport studies from the upper cervical spinal cord and the cervical

enlargement in squirrel monkey and rat. J Comp Neurol 1996;365:640-658.

[643] Nguyen JP, Nizard J, Keravel Y, Lefaucheur JP. Invasive brain stimulation for the

treatment of neuropathic pain. Nat Rev Neurol 2011.

[644] Nieuwenhuys R, Voogd J, van Huijzen C. The Human Central Nervous System. Berlin

Heidelberg: Springer, 2008.

[645] Nishida Y, Yokota T. Corneal representation within the trigeminal subnucleus caudalis and

adjacent bulbar lateral reticular formation of the cat. The Japanese journal of physiology

1991;41:551-565.

[646] Nitzan D, Benoliel R, Heir G, Dolwick F. Pain and dysfunction of the temporomandibular

joing. In: Y Sharav, R Benoliel, editors. Orofacial pain & headache. New York: Elsevier,

2008. pp. 149-192.

277

[647] Nobre AC, Sebestyen GN, Gitelman DR, Mesulam MM, Frackowiak RSJ, Frith CD.

Functional localization of the system for visuospatial attention using positron emission

tomography. Brain 1997;120:515-533.

[648] Nudo RJ, Wise BM, SiFuentes F, Milliken GW. Neural substrates for the effects of

rehabilitative training on motor recovery after ischemic infarct. Science 1996;272:1791-

1794.

[649] O'Reardon JP, Fontecha JF, Cristancho MA, Newman S. Unexpected reduction in migraine

and psychogenic headaches following rTMS treatment for major depression: a report of

two cases. CNS Spectr 2007;12:921-925.

[650] Obermann M, Nebel K, Schumann C, Holle D, Gizewski ER, Maschke M, Goadsby PJ,

Diener HC, Katsavara Z. Gray matter changes related to chronic posttraumatic headache.

Neurology 2009;73:978-993.

[651] Ochs S. A history of nerve functions : from animal spirits to molecular mechanisms.

Cambridge, UK ; New York: Cambridge University Press, 2004.

[652] Ogino Y, Nemoto H, Goto F. Somatotopy in human primary somatosensory cortex in pain

system. Anesthesiology 2005;103:821-827.

[653] Ohrbach R, Dworkin SF. Five-year outcomes in TMD: relationship of changes in pain to

changes in physical and psychological variables. Pain 1998;74:315-326.

[654] Okada-Ogawa A, Porreca F, Meng ID. Sustained morphine-induced sensitization and loss

of diffuse noxious inhibitory controls in dura-sensitive meduallry dorsal horn neurons. J

Neurosci 2009;29:15828-15835.

[655] Oldfield RC. The assessment and analysis of handedness: the Edinburgh inventory.

Neuropsychologia 1971;9:97-113.

[656] Olszewski J. On the anatomical and functional organization of the spinal trigeminal

nucleus. J Comp Neurol 1950;92:401-413.

[657] Osenbach RK. Motor cortex stimulation for intractable pain. Neurosurg Focus 2006;21:E7.

[658] Ostrowsky K, Magnin M, Ryvlin P, Isnard J, Guenot M, Mauguiere F. Representation of

pain and somatic sensation in the human insula: a study of responses to direct electrical

cortical stimulation. Cereb Cortex 2002;12:376-385.

278

[659] Owen AM, Schiff ND, Laureys S. A new era of coma and consciousness science. Prog

Brain Res 2009;177:399-411.

[660] Ozarslan E, Nevo U, Basser PJ. Anisotropy induced by macroscopic boundaries: surface-

normal mapping using diffusion-weighted imaging. Biophys J 2008;94:2809-2818.

[661] Ozarslan E, Shepherd TM, Vemuri BC, Blackband SJ, Mareci TH. Resolution of complex

tissue microarchitecture using the diffusion orientation transform (DOT). Neuroimage

2006;31:1086-1103.

[662] Pakkenberg B, Gundersen HJ. Neocortical neuron number in humans: effect of sex and

age. J Comp Neurol 1997;384:312-320.

[663] Pandya DN, Kuypers HG. Cortico-cortical connections in the rhesus monkey. Brain Res

1969;13:13-36.

[664] Papèz JW. A proposed circuit for emotion. Arch Neurol Psychiatry 1937;38:725-746.

[665] Parent A, Hazrati LN. Functional anatomy of the basal ganglia. I. The cortico- basal

ganglia-thalamo-cortical loop. Brain Res Rev 1995;20:91-127.

[666] Park DC, Glass JM, Minear M, Crofford LJ. Cognitive function in fibromyalgia patients.

Arthritis Rheum 2001;44:2125-2133.

[667] Park HJ, Westin CF, Kubicki M, Maier SE, Niznikiewicz M, Baer A, Frumin M, Kikinis

R, Jolesz FA, McCarley RW, Shenton ME. White matter hemisphere asymmetries in

healthy subjects and in schizophrenia: a diffusion tensor MRI study. Neuroimage

2004;23:213-223.

[668] Parker GJ. Analysis of MR diffusion weighted images. Br J Radiol 2004;77 Spec No

2:S176-185.

[669] Parker GJ, Alexander DC. Probabilistic anatomical connectivity derived from the

microscopic persistent angular structure of cerebral tissue. Phil Trans R Soc B

2005;360:893-902.

[670] Parker GJ, Haroon HA, Wheeler-Kingshott CA. A framework for a streamline-based

probabilistic index of connectivity (PICo) using a structural interpretation of MRI

diffusion measurements. J Magn Reson Imaging 2003;18:242-254.

279

[671] Parker GJM. Probabilistic fibre tracking. In: DK Jones, editor. Diffusion MRI: Theory,

Methods and Applications. New York: Oxford University Press, 2011. pp. 396-409.

[672] Pascual-Leone A, Amedi A, Fregni F, Merabet LB. The plastic human brain cortex. Ann

Rev Neurosci 2005;28:377-401.

[673] Pascual-Leone A, Grafman J, Hallett M. Modulation of cortical motor output maps during

development of implicit and explicit knowledge. Science 1994;263:1287-1289.

[674] Pavlides C, Miyashita E, Asanuma H. Projection from the sensory to the motor cortex is

important in learning motor skills in the monkey. J Neurophysiol 1993;70:733-741.

[675] Peck CC, Murray GM, Gerzina TM. How does pain affect jaw muscle activity? The

Integrated Pain Adaptation Model. Aust Dent J 2008;53:201-207.

[676] Peele TL. Acute and chronic parietal lobe ablations in monkeys. J Neurophysiol

1944;7:269-286.

[677] Penfield W. The Influence of the Diencephalon and Hypophysis Upon General Autonomic

Function. Can Med Assoc J 1934;30:589-598.

[678] Penfield W. Some observations on the cerebral cortex of man. Proc R Soc B

1947;134:329-347.

[679] Penfield W, Boldrey E. Somatic motor and sensory representation in the cerebral cortex of

man as studied by electrical stimulation. Brain 1937;60:389-443.

[680] Peper JS, Brouwer RM, Boomsma DI, Kahn RS, Hulshoff Pol HE. Genetic influences on

human brain structure: a review of brain imaging studies in twins. Hum Brain Mapp

2007;28:464-473.

[681] Perl ER. Myelinated afferent fibres innervating the primate skin and their response to

noxious stimuli. J Physiol 1968;197:593-615.

[682] Perl ER. Pain and the discovery of nociceptors. In: C Belmonte, F Cervero, editors.

Neurobiology of nociceptors. New York: Oxford University Press, 1996. pp. 5-36.

[683] Perl ER, Kruger L. Nociception and Pain: Evolution of concepts and observations. In: L

Kruger, editor. Pain and Touch. San Diego, CA: Academic Press, 1996. pp. 180-212.

280

[684] Perl ER, Whitlock DG. Somatic stimuli exciting spinothalamic projections to thalamic

neurons in cat and monkey. Exp Neurol 1961;3:256-296.

[685] Peters A, Morrison JH, Rosene DL, Hyman BT. Feature article: are neurons lost from the

primate cerebral cortex during normal aging? Cereb Cortex 1998;8:295-300.

[686] Petrides M, Pandya DN. Dorsolateral prefrontal cortex: comparative cytoarchitectonic

analysis in the human and the macaque brain and corticocortical connection patterns.

European Journal of Neuroscience 1999;11:1011-1036.

[687] Petrovic P, Ingvar M. Imaging cognitive modulation of pain processing. Pain 2002;95:1-5.

[688] Peyron R, Frot M, Schneider F, Garcia-Larrea L, Mertens P, Barral FG, Sindou M, Laurent

B, Mauguiere F. Role of operculoinsular cortices in human pain processing: converging

evidence from PET, fMRI, dipole modeling, and intracerebral recordings of evoked

potentials. Neuroimage 2002;17:1336-1346.

[689] Peyron R, Garcia-Larrea L, Deiber MP, Cinotti L, Convers P, Sindou M, Mauguiäre F,

Laurent B. Electrical stimulation of precentral cortical area in the treatment of central

pain: Electrophysiological and PET study. Pain 1995;62:275-286.

[690] Peyron R, Laurent B, Garcia-Larrea L. Functional imaging of brain responses to pain. A

review and meta- analysis. Neurophysiol Clin 2000;30:263-288.

[691] Picard N, Strick PL. Motor areas of the medial wall: A review of their location and

functional activation. Cereb Cortex 1996;6:342-353.

[692] Picard N, Strick PL. Imaging the premotor areas. Curr Opin Neurobiol 2001;11:663-672.

[693] Piche M, Bouin M, Arsenault M, Poitras P, Rainville P. Decreased pain inhibition in

irritable bowel syndrome depends on altered descending modulation and higher-order

brain processes. Neuroscience 2011;195C:166-175.

[694] Pierpaoli C. Artifacts in diffusion MRI. In: DK Jones, editor. Diffusion MRI : theory,

methods, and application. New York: Oxford University Press, 2011. pp. 303-318.

[695] Pierpaoli C, Jezzard P, Basser PJ, Barnett A, Di Chiro G. Diffusion tensor MR imaging of

the human brain. Radiology 1996;201:637-648.

281

[696] Pillay PK, Hassenbusch SJ. Bilateral MRI-guided stereotactic cingulotomy for intractable

pain. Stereotact Funct Neurosurg 1992;59:33-38.

[697] Pincus T, Morley S. Cognitive-processing bias in chronic pain: a review and integration.

Psychol Bull 2001;127:599-617.

[698] Plato. Timaeus. Book Timaeus. City: Project Gutenberg, 1998.

[699] Ploner M, Gross J, Timmermann L, Schnitzler A. Pain processing is faster than tactile

processing in the human brain. J Neurosci 2006;26:10879-10882.

[700] Pochon JB, Levy R, Poline JB, Crozier S, LehÇricy S, Pillon B, Deweer B, Le Bihan D,

Dubois B. The role of dorsolateral prefrontal cortex in the preparation of forthcoming

actions: an fMRI study. Cereb Cortex 2001;11:260-266.

[701] Poggio GF, Mountcastle VB. The Functional Properties of Ventrobasal Thalamic

Neuronsstudied in Unanesthetized Monkeys. J Neurophysiol 1963;26:775-806.

[702] Pons TP, Garraghty PE, Mishkin M. Serial and parallel processing of tactual information in

somatosensory cortex of rhesus monkeys. J Neurophysiol 1992;68:518-527.

[703] Pons TP, Garraghty PE, Ommaya AK, Kaas JH, Taub E, Mishkin M. Massive cortical

reorganization after sensory deafferentation in adult macaques. Science 1991;252:1857-

1860.

[704] Porreca F, Ossipov MH, Gebhart GF. Chronic pain and medullary descending facilitation.

Trends in Neurosciences 2002;25:319-325.

[705] Porro CA, Cettolo V, Francescato MP, Baraldi P. Temporal and intensity coding of pain in

human cortex. J Neurophysiol 1998;80:3312-3320.

[706] Porter LL, Sakamoto K. Organization and synaptic relationships of the projection from the

primary sensory to the primary motor cortex in the cat. J Comp Neurol 1988;271:387-

396.

[707] Potvin S, Grignon S, Marchand S. Human evidence of a supra-spinal modulating role of

dopamine on pain perception. Synapse 2009;63:390-402.

282

[708] Pozo MA, Cervero F. Neurons in the rat spinal trigeminal complex driven by corneal

nociceptors: receptive-field properties and effects of noxious stimulation of the cornea. J

Neurophysiol 1993;70:2370-2378.

[709] Price DD. Psychological and neural mechanisms of the affective dimension of pain.

Science 2000;288:1769-1772.

[710] Price DD, Dubner R. Mechanisms of first and second pain in the peripheral and central

nervous systems. J Invest Dermatol 1977;69:167-171.

[711] Price DD, Hu JW, Dubner R, Gracely RH. Peripheral suppression of first pain and central

summation of second pain evoked by noxious heat pulses. Pain 1977;3:57-68.

[712] Puri BK, Agour M, Gunatilake KD, Fernando KA, Gurusinghe AI, Treasaden IH.

Reduction in left supplementary motor area grey matter in adult female fibromyalgia

sufferers with marked fatigue and without affective disorder: a pilot controlled 3-T

magnetic resonance imaging voxel-based morphometry study. The Journal of

international medical research 2010;38:1468-1472.

[713] Quevedo AS, Coghill RC. Attentional modulation of spatial integration of pain: evidence

for dynamic spatial tuning. J Neurosci 2007;27:11635-11640.

[714] Rainville P. Brain mechanisms of pain affect and pain modulation. Curr Opin Neurobiol

2002;12:195-204.

[715] Rainville P, Carrier B, Hofbauer RK, Bushnell MC, Duncan GH. Dissociation of sensory

and affective dimensions of pain using hypnotic modulation. Pain 1999;82:159-171.

[716] Rainville P, Duncan GH, Price DD, Carrier B, Bushnell MC. Pain affect encoded in human

anterior cingulate but not somatosensory cortex. Science 1997;277:968-971.

[717] Rainville P, Feine JS, Bushnell MC, Duncan GH. A psychophysical comparison of sensory

and affective responses to four modalities of experimental pain. Somatosens Mot Res

1992;9:265-277.

[718] Raja SN, Campbell JN, Meyer RA. Evidence for different mechanisms of primary and

secondary hyperalgesia following heat injury to the glabrous skin. Brain 1984;107:1179-

1188.

283

[719] Ralston HJ, III. The synaptic organization in the dorsal horn of the spinal cord and the

ventrobasal thalamus of the cat. In: R Dubner, editor. Orofacial Sensory and Motor

Mechanisms. New York: Appleton-Century-Croft, 1971. pp. 229-250.

[720] Ralston HJ, III, Ralston DD. The distribution of dorsal root axons in laminae I, II and III of

the macaque spinal cord: a quantitative electron microscope study. J Comp Neurol

1979;184:643-684.

[721] Ramírez LM, Sandoval GP, Ballesteros LE. Temporomandibular disorders: referred

cranio-cervico-facial clinic. Medicina oral, patología oral y cirugía bucal 2005;10 Suppl

1:E18-26.

[722] Rashid W, Hadjiprocopis A, Griffin CM, Chard DT, Davies GR, Barker GJ, Tofts PS,

Thompson AJ, Miller DH. Diffusion tensor imaging of early relapsing-remitting multiple

sclerosis with histogram analysis using automated segmentation and brain volume

correction. Mult Scler 2004;10:9-15.

[723] Ray RD, Ochsner KN, Cooper JC, Robertson ER, Gabrieli JD, Gross JJ. Individual

differences in trait rumination and the neural systems supporting cognitive reappraisal.

Cogn Affect Behav Neurosci 2005;5:156-168.

[724] Ren K, Dubner R. Descending modulation in persistent pain: an update. Pain 2002;100:1-

6.

[725] Ren K, Dubner R. Pain facilitation and activity-dependent plasticity in pain modulatory

circuitry: role of BDNF-TrkB signaling and NMDA receptors. Mol Neurobiol

2007;35:224-235.

[726] Ren K, Dubner R. Descending control mechanisms. In: AI Basbaum, MC Bushnell,

editors. Science of Pain. San Diego: Academic Press, 2009. pp. 724-762.

[727] Rexed B. The cytoarchitectonic organization of the spinal cord in the cat. J Comp Neurol

1952;96:414-495.

[728] Rexed B. A cytoarchitectonic atlas of the spinal cord in the cat. J Comp Neurol

1954;100:297-379.

[729] Rey R. The history of pain. Cambridge, Mass.: Harvard University Press, 1995.

284

[730] Reynolds DV. Surgery in the rat during electrical analgesia induced by focal brain

stimulation. Science 1969;164:444-445.

[731] Rhodes DL, Liebeskind JC. Analgesia from rostral brain stem stimulation in the rat. Brain

Res 1978;143:521-532.

[732] Richards CD, Taylor DC. Electrophysiological evidence for a somatotopic sensory

projection to the striatum of the rat. Neurosci Lett 1982;30:235-240.

[733] Richter EO, Davis KD, Hamani C, Hutchison WD, Dostrovsky JO, Lozano AM.

Cingulotomy for psychiatric disease: microelectrode guidance, a callosal reference

system for documenting lesion location, and clinical results. Neurosurgery 2004;54:622-

628.

[734] Rios M, Treede R, Lee J, Lenz FA. Direct Evidence of Nociceptive Input to Human

Anterior Cingulate Gyrus and Parasylvian Cortex. Curr Rev Pain 1999;3:256-264.

[735] Robinson CJ, Burton H. Organization of somatosensory receptive fields in cortical areas

7b, retroinsula, postauditory and grnaular insula of M. fascicularis J Comp Neurol

1980;192:69-92.

[736] Robinson CJ, Burton H. Somatic submodality distribution within the second

somatosensory (SII), 7b, retroinsular, postauditory, and granular insular cortical areas of

M. fascicularis. J Comp Neurol 1980;192:93-108.

[737] Robinson CJ, Burton H. Somatotographic organization in the second somatosensory area

of M. fascicularis. J Comp Neurol 1980;192:43-67.

[738] Robinson ME, Craggs JG, Price DD, Perlstein WM, Staud R. Gray matter volumes of

pain-related brain areas are decreased in fibromyalgia syndrome. J Pain 2011;12:436-443.

[739] Roby-Brami A, Bussel B, Willer JC, Le Bars D. An electrophysiological investigation into

the pain-relieving effects of heterotopic nociceptive stimuli. Probable involvement of a

supraspinal loop. Brain 1987;110 ( Pt 6):1497-1508.

[740] Rocca MA, Ceccarelli A, Falini A, Colombo B, Tortorella P, Bernasconi L, Comi G, Scotti

G, Filippi M. Brain gray matter changes in migraine patients with T2-visible lesions: a 3-

T MRI study. Stroke 2006;37:1765-1770.

285

[741] Rodriguez-Raecke R, Niemeier A, Ihle K, Ruether W, May A. Brain gray matter decrease

in chronic pain is the consequence and not the cause of pain. J Neurosci 2009;29:13746-

13750.

[742] Rolls ET, Grabenhorst F. The orbitofrontal cortex and beyond: from affect to decision-

making. Prog Neurobiol 2008;86:216-244.

[743] Romaniello A, Cruccu G, Frisardi G, Arendt-Nielsen L, Svensson P. Assessment of

nociceptive trigeminal pathways by laser-evoked potentials and laser silent periods in

patients with painful temporomandibular disorders. Pain 2003;103:31-39.

[744] Romaniello A, Cruccu G, McMillan AS, Arendt-Nielsen L, Svensson P. Effect of

experimental pain from trigeminal muscle and skin on motor cortex excitability in

humans. Brain Res 2000;882:120-127.

[745] Rosenkranz K, Williamon A, Butler K, Cordivari C, Lees AJ, Rothwell JC.

Pathophysiological differences between musician's dystonia and writer's cramp. Brain

2005;128:918-931.

[746] Russel CK, Horsley V. Note on apparent re-representation in the cerebral cortex of the type

of sensory representation as it exists in the spinal cord. Brain 1907;29:137-151.

[747] Sakamoto T, Arissian K, Asanuma H. Functional role of the sensory cortex in learning

motor skills in cats. Brain Res 1989;503:258-264.

[748] Sakamoto T, Porter LL, Asanuma H. Long-lasting potentiation of synaptic potentials in the

motor cortex produced by stimulation of the sensory cortex in the cat: a basis of motor

learning. Brain Res 1987;413:360-364.

[749] Salat DH, Buckner RL, Snyder AZ, Greve DN, Desikan RS, Busa E, Morris JC, Dale AM,

Fischl B. Thinning of the cerebral cortex in aging. Cereb Cortex 2004;14:721-730.

[750] Salat DH, Lee SY, van der Kouwe AJ, Greve DN, Fischl B, Rosas HD. Age-associated

alterations in cortical gray and white matter signal intensity and gray to white matter

contrast. Neuroimage 2009;48:21-28.

[751] Salat DH, Tuch DS, Greve DN, van der Kouwe AJ, Hevelone ND, Zaleta AK, Rosen BR,

Fischl B, Corkin S, Rosas HD, Dale AM. Age-related alterations in white matter

286

microstructure measured by diffusion tensor imaging. Neurobiol Aging 2005;26:1215-

1227.

[752] Salomons TV, Johnstone T, Backonja M, Shackman AJ, Davidson RJ. Individual

differences in the effects of perceived controllability on pain perception: critical role of

the prefrontal cortex. J Cogn Neurosci 2007;19:993-1004.

[753] Salomons TV, Johnstone T, Backonja MM, Davidson RJ. Perceived controllability

modulates the neural response to pain. J Neurosci 2004;24:7199-7203.

[754] Sampson SM, Rome JD, Rummans TA. Slow-frequency rTMS reduces fibromyalgia pain.

Pain Med 2006;7:115-118.

[755] Sarlani E, Garrett PH, Grace EG, Greenspan JD. Temporal summation of pain

characterizes women but not men with termporomandibular disorders. J Orofac Pain

2007;21:309-317.

[756] Sarlani E, Greenspan JD. Evidence for generalized hyperalgesia in temporomandibular

disorders patients. Pain 2003;102:221-226.

[757] Sarlani E, Greenspan JD. Why look in the brain for answers to temporomandibular

disorder pain? Cells Tissues Organs 2005;180:69-75.

[758] Schallert T, Kozlowski DA, Humm JL, Cocke RR. Use-dependent structural events in

recovery of function. Adv Neurol 1997;73:229-238.

[759] Scherder E, Wolters E, Polman C, Sergeant J, Swaab D. Pain in Parkinson's disease and

multiple sclerosis: Its relation to the medial and lateral pain systems. Neurosci Biobehav

Rev 2005;29:1047-1056.

[760] Schmahmann JD, Pandya DN. Fiber pathways of the brain. New York: Oxford University

Press, 2006.

[761] Schmahmann JD, Pandya DN. Preface. In: JD Schmahmann, DN Pandya, editors. Fiber

pathways of the brain. New York: Oxford University Press, 2006. pp. xi-xii.

[762] Schmidt-Wilcke T, Ganssbauer S, Neuner T, Bogdahn U, May A. Subtle grey matter

changes between migraine patients and healthy controls. Cephalalgia : an international

journal of headache 2008;28:1-4.

287

[763] Schmidt-Wilcke T, Hierlmeier S, E. L. Altered regional brain morphology in patients with

chronic facial pain. Headache 2010;50:1278-1212-1285.

[764] Schmidt-Wilcke T, Leinisch E, Ganssbauer S, Draganski B, Bogdahn U, Altmeppen J,

May A. Affective components and intensity of pain correlate with structural differences

in gray matter in chronic back pain patients. Pain 2006;125:89-97.

[765] Schmidt-Wilcke T, Leinisch E, Straube A, Kampfe N, Draganski B, Diener HC, Bogdahn

U, May A. Gray matter decrease in patients with chronic tension type headache.

Neurology 2005;65:1483-1486.

[766] Schmidt-Wilcke T, Luerding R, Weigand T, Jurgens T, Schuierer G, Leinisch E, Bogdahn

U. Striatal grey matter increase in patients suffering from fibromyalgia--a voxel-based

morphometry study. Pain 2007;132 Suppl 1:S109-116.

[767] Schmitter AM. 17th and 18th Century Theories of Emotions. In: EN Zalta editor. Book

17th and 18th Century Theories of Emotions. City, 2010.

[768] Schnakers C, Chatelle C, Vanhaudenhuyse A, Majerus S, Ledoux D, Boly M, Bruno MA,

Boveroux P, Demertzi A, Moonen G, Laureys S. The Nociception Coma Scale: a new

tool to assess nociception in disorders of consciousness. Pain 2010;148:215-219.

[769] Schneider JS, Lidsky TI. Processing of somatosensory information in striatum of behaving

cats. J Neurophysiol 1981;45:841-851.

[770] Schoenbaum G, Roesch MR, Stalnaker TA, Takahashi YK. A new perspective on the role

of the orbitofrontal cortex in adaptive behaviour. Nat Rev Neurosci 2009;10:885-892.

[771] Scholz J, Klein MC, Behrens TE, Johansen-Berg H. Training induces changes in white-

matter architecture. Nat Neurosci 2009;12:1370-1371.

[772] Schweinhardt P, Kuchinad A, Pukall CF, Bushnell MC. Increased gray matter density in

young women with chronic vulvar pain. Pain 2008;140:411-419.

[773] Sechzer JA. The ethical dilemma of some classical animal experiments. Ann N Y Acad Sci

1983;406:5-12.

[774] Seeley WW, Menon V, Schatzberg AF, Keller J, Glover GH, Kenna H, Reiss AL, Greicius

MD. Dissociable intrinsic connectivity networks for salience processing and executive

control. J Neurosci 2007;27:2349-2356.

288

[775] Seifert F, Maihofner C. Representation of cold allodynia in the human brain--a functional

MRI study. Neuroimage 2007;35:1168-1180.

[776] Seminowicz DA, Davis KD. Interactions of pain intensity and cognitive load: the brain

stays on task. Cereb Cortex 2007;17:1412-1422.

[777] Seminowicz DA, Davis KD. Pain enhances functional connectivity of a brain network

evoked by performance of a cognitive task. J Neurophysiol 2007;97:3651-3659.

[778] Seminowicz DA, Davis KD. A re-examination of pain-cognition interactions: implications

for neuroimaging. Pain 2007;130:8-13.

[779] Seminowicz DA, Labus JS, Bueller JA, Tillisch K, Naliboff BD, Bushnell MC, Mayer EA.

Regional gray matter density changes in Brains of Patients with Irritable Bowel

Syndrome. Gastroenterology 2010;139:48-57.

[780] Seminowicz DA, Laferriere AL, Millecamps M, Yu JSC, Coderre TJ, Bushnell MC. MRI

structural brain changes associated with sensory and emotional function in a rat model of

long-term neuropathic pain. Neuroimage 2009;47:1007-1014.

[781] Seminowicz DA, Mikulis DJ, Davis KD. Cognitive modulation of pain-related brain

responses depends on behavioral strategy. Pain 2004;112:48-58.

[782] Seminowicz DA, Wideman TH, Naso L, Hatami-Khoroushahi Z, Fallatah S, Ware MA,

Jarzem P, Bushnell MC, Shir Y, Ouellet JA, Stone LS. Effective treatment of chronic low

back pain in humans reverses abnormal brain anatomy and function. J Neurosci

2011;31:7540-7550.

[783] Sessle BJ. Recent developments in pain research: central mechanisms of orofacial pain and

its control. J Endod 1986;12:435-444.

[784] Sessle BJ. Acute and chronic craniofacial pain: brainstem mechanisms of nociceptive

transmission and neuroplasticity, and their clinical correlates. Crit Rev Oral Biol Med

2000;11:57-91.

[785] Sessle BJ. Mechanisms of oral somatosensory and motor functions and their clinical

correlates. J Oral Rehabil 2006;33:243-261.

289

[786] Sessle BJ, Dubner R, Greenwood LF, Lucier GE. Descending influences of periaqueductal

gray matter and somatosensory cerebral cortex on neurones in trigeminal brain stem

nuclei. Can J Physiol Pharmacol 1976;54:66-69.

[787] Sessle BJ, Hu JW. Mechanisms of pain arising from articular tissues. Can J Physiol

Pharmacol 1991;69:617-626.

[788] Sessle BJ, Hu JW, Amano N, Zhong G. Convergence of cutaneous, tooth pulp, visceral,

neck and muscle afferents onto nociceptive and non-nociceptive neurones in trigeminal

subnucleus caudalis (medullary dorsal horn) and its implications for referred pain. Pain

1986;27:219-235.

[789] Sessle BJ, Hu JW, Dubner R, Lucier GE. Functional properties of neurons in cat trigeminal

subnucleus caudalis (medullary dorsal horn). II. Modulation of responses to noxious and

nonnoxious stimuli by periaqueductal gray, nucleus raphe magnus, cerebral cortex, and

afferent influences, and effect of naloxone. J Neurophysiol 1981;45:193-207.

[790] Shackman AJ, Salomons TV, Slagter HA, Fox AS, Winter JJ, Davidson RJ. The

integration of negative affect, pain and cognitive control in the cingulate cortex. Nat Rev

Neurosci 2011;12:154-167.

[791] Sharav Y, Benoliel R. The diagnostic process. In: Y Sharav, R Benoliel, editors. Orofacial

pain & headache. New York: Elsevier, 2008. pp. 1-17.

[792] Sharav Y, McGrath PA, Dubner R. Masseter inhibitory periods and sensations evoked by

electrical tooth pulp stimulation in patients with oral-facial pain and mandibular

dysfunction. Arch Oral Biol 1982;27:305-310.

[793] Sharp TJ. Chronic pain: a reformulation of the cognitive-behavioural model. Behav Res

Ther 2001;39:787-800.

[794] Sherman SM, Guillery RW. Exploring the thalamus and its role in cortical function.

Cambridge, MA: MIT Press, 2001.

[795] Sherrington CS. Qualitative difference of spinal reflex corresponding with qualitative

difference of cutaneous stimulus. J Physiol 1903;30:39-46.

[796] Sherrington CS. Observations on the scratch-reflex in the spinal dog. J Physiol 1906;34:1-

50.

290

[797] Sherrington CS. Man on his nature. Cambridge, England: Cambridge University press,

1955.

[798] Shibata H, Yukie M. Thalamocingulate connections in the monkey. In: BA Vogt, editor.

Cingulate Neurobiology and Disease. New York: Oxford University Press, 2009. pp. 96-

111.

[799] Shibukawa Y, Ishikawa T, Kato Y, Shintani M, Zhang ZK, Jiang T, Tazaki M, Shimono

M, Kumai T, Suzuki T, Kato M, Nakamura Y. Cortical dysfunction in patients with

temporomandibular disorders. J Oral Biosci 2009;51:65-71.

[800] Shibukawa Y, Ishikawa T, Kato Y, Zhang Z-K, Jiang T, Shintani M, Shimono M, Kumai

T, Suzuki T, Kato M, Nakamura Y. Cerebral cortical dysfunction in patients with

temporomandibular disorders in association with jaw movement observation. Pain

2007;128:180-188.

[801] Shigenaga Y, Chen IC, Suemune S, Nishimori T, Nasution ID, Yoshida A, Sato H,

Okamoto T, Sera M, Hosoi M. Oral and facial representation within the medullary and

upper cervical dorsal horns in the cat. J Comp Neurol 1986;243:388-408.

[802] Short EB, Borckardt JJ, Anderson BS, Frohman H, Beam W, Reeves ST, George MS. Ten

sessions of adjunctive left prefrontal rTMS significantly reduces fibromyalgia pain: A

randomized, controlled pilot study. Pain 2011;152:2477-2484.

[803] Siegel A, Troiano R, Royce A. Differential projections of the anterior and posterior

cingulate gyrus to the thalamus in the cat. Exp Neurol 1973;38:192-201.

[804] Simon TJ, Ding L, Bish JP, McDonald-McGinn DM, Zackai EH, Gee J. Volumetric,

connective, and morphologic changes in the brains of children with chromosome 22q11.2

deletion syndrome: an integrative study. Neuroimage 2005;25:169-180.

[805] Simone DA, Marchettini P, Caputi G, Ochoa JL. Identification of muscle afferents

subserving sensation of deep pain in humans. J Neurophysiol 1994;72:883-889.

[806] Simone DA, Sorkin LS, Oh U, Chung JM, Owens C, LaMotte RH, Willis WD. Neurogenic

hyperalgesia: Central neural correlates in responses of spinothalamic tract neurons. J

Neurophysiol 1991;66:228-246.

291

[807] Sinclair DC. Cutaneous sensation and the doctrine of specific energy. Brain 1955;78:584-

614.

[808] Slade GD, Diatchenko L, Bhalang K, Sigurdsson A, Fillingim RB, Belfer I, Max MB,

Goldman D, Maixner W. Influence of psychological factors on risk of

temporomandibular disorders. J Dent Res 2007;86:1120-1125.

[809] Slade GD, Diatchenko L, Ohrbach R, Maixner W. Orthodontic Treatment, Genetic Factors

and Risk of Temporomandibular Disorder. Semin Orthod 2008;14:146-156.

[810] Slakey TJ. Aristotle on sense perception. The Phil Rev 1961;70:470-484.

[811] Smith RL. The ascending fiber projections from the principal sensory trigeminal nucleus in

the rat. J Comp Neurol 1973;148:423-445.

[812] Smith RL. Axonal projections and connections of the principal sensory trigeminal nucleus

in the monkey. J Comp Neurol 1975;163:347-375.

[813] Smith S, Kindlmann G. Cross-subject Comparison of local Diffusion MRI Parameters. In:

H Johansen-Berg, TE Behrens, editors. Diffusion MRI. London: Academic Press, 2009.

pp. 148-174.

[814] Smith SM. Fast robust automated brain extraction. Hum Brain Mapp 2002;17:144-158.

[815] Smith SM, Jenkinson M, Johansen-Berg H, Rueckert D, Nichols TE, Mackay CE, Watkins

KE, Ciccarelli O, Cader MZ, Matthews PM, Behrens TE. Tract-based spatial statistics:

voxelwise analysis of multi-subject diffusion data. Neuroimage 2006;31:1487-1505.

[816] Smith SM, Jenkinson M, Woolrich M, Beckmann CF, Behrens TE, Johansen-Berg H,

Bannister PR, De Luca M, Drobnjak I, Flitney DE, Niazy R, Saunders J, Vickers J,

Zhang Y, De Stefano N, Brady JM, Matthews PM. Advances in functional and structural

MR image analysis and implementation as FSL. Neuroimage 2004;23:208-220.

[817] Smith SM, Johansen-Berg H, Jenkinson M, Rueckert D, Nichols TE, Miller KL, Robson

MD, Jones DK, Klein JC, Bartsch AJ, Behrens TE. Acquisition and voxelwise analysis of

multi-subject diffusion data with tract-based spatial statistics. NatProtoc 2007;2:499-503.

[818] Smith SM, Johansen-Berg H, Jenkinson M, Rueckert D, Nichols TE, Miller KL, Robson

MD, Jones DK, Klein JC, Bartsch AJ, Behrens TEJ. Acquisition and voxelwise analysis

292

of multi-subject diffusion data with tract-based spatial statistics. Nat Protoc 2007;2:499-

503.

[819] Smith SM, Nichols TE. Threshold-free cluster enhancement: addressing problems of

smoothing, threshold dependence and localisation in cluster inference. Neuroimage

2009;44:83-98.

[820] Snook L, Plewes C, Beaulieu C. Voxel based versus region of interest analysis in diffusion

tensor imaging of neurodevelopment. Neuroimage 2007;34:243-252.

[821] Soler MD, Kumru H, Pelayo R, Vidal J, Tormos JM, Fregni F, Navarro X, Pascual-Leone

A. Effectiveness of transcranial direct current stimulation and visual illusion on

neuropathic pain in spinal cord injury. Brain 2010;133:2565-2577.

[822] Sommer M, Koch MA, Paulus W, Weiller C, Buchel C. Disconnection of speech-relevant

brain areas in persistent developmental stuttering. Lancet 2002;360:380-383.

[823] Sotnichencko TS. Convergence of the descending pathways of motor, visual and limbic

cortex in the cat di- and mesencephalon. Brain Res 1976;116:401-415.

[824] Sowell ER, Peterson BS, Thompson PM, Welcome SE, Henkenius AL, Toga AW.

Mapping cortical change across the human life span. Nat Neurosci 2003;6:309-315.

[825] Squire LR. David R. Curtis. In: LR Squire editor. Book David R. Curtis, Vol. 5. City:

Academic Press, 2006. pp. 170-225.

[826] Stacey MJ. Free nerve endings in skeletal muscle of the cat. J Anat 1969;105:231-254.

[827] Stahnisch FW. Francois Magendie (1783-1855). J Neurol 2009;256:1950-1952.

[828] Starr CJ, Sawaki L, Wittenberg GF, Burdette JH, Oshiro Y, Quevedo AS, McHaffie JG,

Coghill RC. The contribution of the putamen to sensory aspects of pain: insights from

structural connectivity and brain lesions. Brain 2011;134:1987-2004.

[829] Starr PA, Vitek JL, DeLong M, Bakay RAE. Magnetic resonance imaging-based

stereotactic localization of the globus pallidus and subthalamic nucleus. Neurosurgery

1999;44:303-313.

[830] Steno N, Andrault R. Discours sur l'anatomie du cerveau (1669). Paris: Classiques Garnier,

2009.

293

[831] Stewart WA, King RB. Fiber Projections from the Nucleus Caudalis of the Spinal

Trigeminal Nucleus. J Comp Neurol 1963;121:271-286.

[832] Stohler CS. Craniofacial pain and motor function: pathogenesis, clinical correlates, and

implications. Crit Rev Oral Biol Med 1999;10:504-518.

[833] Stone TT. Phantom limb pain and central pain; relief by ablation of a portion of posterior

central cerebral convolution. Arch Neurol Psychiatry 1950;63:739-748.

[834] Strange BA, Henson RN, Friston KJ, Dolan RJ. Anterior prefrontal cortex mediates rule

learning in humans. Cereb Cortex 2001;11:1040-1046.

[835] Sugiyo S, Takemura M, Dubner R, Ren K. Trigeminal transition zone/rostral ventromedial

medulla connections and facilitation of orofacial hyperalgesia after masseter

inflammation in rats. J Comp Neurol 2005;493:510-523.

[836] Sullivan MJ, Bishop SR, Pivik J. The pain catastrophizing scale: development and

validation. Psychol Assessment 1995;7:524-532.

[837] Svensson P. What can human experimental pain models teach us about clinical TMD?

Arch Oral Biol 2007;52:391-394.

[838] Svensson P, Beydoun A, Morrow TJ, Casey KL. Human intramuscular and cutaneous pain:

Psychophysical comparisons. Exp Brain Res 1997;114:390-392.

[839] Svensson P, Graven-Nielsen T. Craniofacial muscle pain: review of mechanisms and

clinical manifestations. J Orofac Pain 2001;15:117-145.

[840] Svensson P, Jadidi F, Arima T, Baad-Hansen L, Sessle BJ. Relationships between

craniofacial pain and bruxism. J Oral Rehabil 2008;35:524-547.

[841] Svensson P, List T, Hector G. Analysis of stimulus-evoked pain in patients with

myofascial temporomandibular pain disorders. Pain 2001;92:399-409.

[842] Svensson P, Romaniello A, Wang K, Arendt-Nielsen L, Sessle BJ. One hour of tongue-

task training is associated with plasticity in corticomotor control of the human tongue

musculature. Exp Brain Res 2006;173:165-173.

[843] Tal M, Devor M. Anatomy and neurophysiology of orofacial pain In: Y Sharav, R

Benoliel, editors. Orofacial pain & headache. New York: Elsevier, 2008. pp. 19-44.

294

[844] Talairach J, Tournoux P. Co-planar stereotaxic atlas of the human brain. New York:

Thieme Medical Publishers Inc., 1988.

[845] Talairach J, Tournoux P, Bancaud J. Traitement chirurgical central de la douleur, du

thalamus (non compris) au cortex parietal. Acta Neurochirurgica 1959;7: 48-143.

[846] Talbot JD, Marrett S, Evans AC, Meyer E, Bushnell MC, Duncan GH. Multiple

representation of pain in human cerebral cortex. Science 1991;251:1355-1358.

[847] Taylor KS, Anastakis DJ, Davis KD. Cutting your nerve changes your brain. Brain

2009;132:3122-3133.

[848] Taylor KS, Anastakis DJ, Davis KD. Chronic Pain and Sensorimotor Deficits Following

Peripheral Nerve Injury. Pain 2010;15:582-591.

[849] Taylor KS, Seminowicz DA, Davis KD. Two systems of resting state connectivity between

the insula and cingulate cortex. Hum Brain Mapp 2009;30:2731-2745.

[850] Tenenbaum HC, Mock D, Gordon AS, Goldberg MB, Grossi ML, Locker D, Davis KD.

Sensory and affective components of orofacial pain: is it all in your brain? Crit Rev Oral

Biol Med 2001;12:455-468.

[851] Terkelsen AJ, Andersen OK, Molgaard H, Hansen J, Jensen TS. Mental stress inhibits pain

perception and heart rate variability but not a nociceptive withdrawal reflex. Acta Physiol

Scand 2004;180: 405-414.

[852] Teutsch S, Herken W, Bingel U, Schoell E, May A. Changes in brain gray matter due to

repetitive painful stimulation. Neuroimage 2008;42:845-849.

[853] Thalhammer JG, LaMotte RH. Spatial properties of nociceptor sensitization following heat

injury of the skin. Brain Res 1982;231:257-265.

[854] Thompson PM, Cannon TD, Narr KL, van Erp T, Poutanen VP, Huttunen M, Lonnqvist J,

Standertskjold-Nordenstam CG, Kaprio J, Khaledy M, Dail R, Zoumalan CI, Toga AW.

Genetic influences on brain structure. Nat Neurosci 2001;4:1253-1258.

[855] Timmermann L, Ploner M, Haucke K, Schmitz F, Baltissen R, Schnitzler A. Differential

coding of pain intensity in the human primary and secondary somatosensory cortex. J

Neurophysiol 2001;86:1499-1503.

295

[856] Tinazzi M, Zarattini S, Valeriani M, Romito S, Farina S, Moretto G, Smania N, Fiaschi A,

Abbruzzese G. Long-lasting modulation of human motor cortex following prolonged

transcutaneous electrical nerve stimulation (TENS) of forearm muscles: evidence of

reciprocal inhibition and facilitation. Exp Brain Res 2005;161:457-464.

[857] Todd AJ, Koerber HR. Neuroanatomical substrates of spinal nociception. In: SB

McMahon, M Koltzenburg, editors. Wall and Melzack's Textbook of Pain. Edingburgh,

UK: Churchill Livingstone, 2006. pp. 73–90.

[858] Toga AW, Thompson PM, Sowell ER. Mapping brain maturation. Trends Neurosci

2006;29:148-159.

[859] Tokuno H, Tanji J. Input organization of distal and proximal forelimb areas in the monkey

primary motor cortex: a retrograde double labeling study. J Comp Neurol 1993;333:199-

209.

[860] Torta DM, Cauda F. Different functions in the cingulate cortex, a meta-analytic

connectivity modeling study. Neuroimage 2011;56:2157-2172.

[861] Torvik A. The ascending fibers from the main trigeminal sensory nucleus; an experimental

study in the cat. Am J Anat 1957;100:1-15.

[862] Tournier JD, Calamante F, Gadian DG, Connelly A. Direct estimation of the fiber

orientation density function from diffusion-weighted MRI data using spherical

deconvolution. Neuroimage 2004;23:1176-1185.

[863] Towell AD, Purves AM, Boyd SG. CO 2 laser activation of nociceptive and non-

nociceptive thermal afferents from hairy and glabrous skin. Pain 1996;66:79-86.

[864] Tracey I, Ploghaus A, Gati JS, Clare S, Smith S, Menon RS, Matthews PM. Imaging

attentional modulation of pain in the periaqueductal gray in humans. J Neurosci

2002;22:2748-2752.

[865] Trachtenberg JT, Chen BE, Knott GW, Feng G, Sanes JR, Welker E, Svoboda K. Long-

term in vivo imaging of experience-dependent synaptic plasticity in adult cortex. Nature

2002;420:788-794.

[866] Travell J. Pain and disability of the shoulder and arm. J Am Med Assoc 1942;120:417-422.

[867] Treede RD. Peripheral acute pain mechanisms. AnnMed 1995;27:213-216.

296

[868] Treede RD, Apkarian AV, Bromm B, Greenspan JD, Lenz FA. Cortical representation of

pain: functional characterization of nociceptive areas near the lateral sulcus. Pain

2000;87:113-119.

[869] Treede RD, Kenshalo DR, Gracely RH, Jones AKP. The cortical representation of pain.

Pain 1999;79:105-111.

[870] Treede RD, Meyer RA, Raja SN, Campbell JN. Evidence for two different heat

transduction mechanisms in nociceptive primary afferents innervating monkey skin. J

Physiol 1995;483:747-758.

[871] Treede RD, Vogel H, Rios M, Krauss G, Lesser RP, Lenz FA. Pain-related evoked

potentials from parasylvian cortex in humans. Electroencephalogr Clin Neurophysiol

Suppl 1999;49:250-254.

[872] Truelove E, Huggins KH, Mancl L, Dworkin SF. The efficacy of traditional, low-cost and

non-splint therapies for temporomandibular disorders. J Am Dent Assoc 2006;137:1099-

1107.

[873] Tsou K, Jang CS. Studies on the Site of Analgesic Action of Morphine by Intracerebral

Micro-Injection. Sci Sin 1964;13:1099-1109.

[874] Tsubokawa T, Katayama Y, Yamamoto T, Hirayama T, Koyama S. Chronic motor cortex

stimulation for the treatment of central pain. Acta Neurochir Suppl (Wien) 1991;52:137-

139.

[875] Tsujimoto S, Genovesio A, Wise SP. Evaluating self-generated decision in the frontal

polar cortex of monkeys. Nat Neurosci 2010;13:120-126.

[876] Tsujimoto T, Ogawa M, Tsukada H, Kakiuchi T, Sasaki K. Decline of the monkey's limbic

and prefrontal activity during task repetition. Neurosci Lett 2000;283:69-72.

[877] Tuch DS. Q-ball imaging. Magn Reson Med 2004;52:1358-1372.

[878] Tuch DS, Reese TG, Wiegell MR, Makris N, Belliveau JW, Wedeen VJ. High angular

resolution diffusion imaging reveals intravoxel white matter fiber heterogeneity. Magn

Reson Med 2002;48:577-582.

[879] Tuch DS, Reese TG, Wiegell MR, Wedeen VJ. Diffusion MRI of complex neural

architecture. Neuron 2003;40:885-895.

297

[880] Tucker WI, Dynes JB. Indications for lobotomy. Lahey Clin Bull 1949;6:90-96.

[881] Turner JA, Dworkin SF, Mancl L, Huggins KH, Truelove EL. The roles of beliefs,

catastrophizing, and coping in the functioning of patients with temporomandibular

disorders. Pain 2001;92:41-51.

[882] Turp JC, Komine F, Hugger A. Efficacy of stabilization splints for the management of

patients with masticatory muscle pain: a qualitative systematic review. Clin Oral Investig

2004;8:179-195.

[883] Turp JC, Kowalski CJ, O'Leary N, Stohler CS. Pain maps from facial pain patients indicate

a broad pain geography. J Dent Res 1998;77:1465-1471.

[884] Usunoff KG, Marani E, Schoen JH. The trigeminal system in man. Adv Anat Embryol Cell

Biol 1997;136:I-X, 1-126.

[885] Valentini E, Hu L, Chakrabarti B, Hu Y, Aglioti SM, Iannetti GD. The primary

somatosensory cortex largely contributes to the early part of the cortical response elicited

by nociceptive stimuli. Neuroimage 2011.

[886] Valentini E, Torta DM, Mouraux A, Iannetti GD. Dishabituation of Laser-evoked EEG

Responses: Dissecting the Effect of Certain and Uncertain Changes in Stimulus Modality.

J Cogn Neurosci 2011.

[887] Valet M, Gundel H, Sprenger T, Sorg C, Muhlau M, Zimmer C, Henningsen P, Tolle TR.

Patients with pain disorder show gray-matter loss in pain-processing structures: a voxel-

based morphometric study. Psychosom Med 2009;71:49-56.

[888] Valet M, Sprenger T, Boecker H, Willoch F, Rummeny E, Conrad B, Erhard P, Tolle TR.

Distraction modulates connectivity of the cingulo-frontal cortex and the midbrain during

pain - an fMR1 analysis. Pain 2004;109:399-408.

[889] Valfre W, Rainero I, Bergui M, Pinessi L. Voxel-based morphometry reveals gray matter

abnormalities in migraine. Headache 2008;48:109-117.

[890] van Buren JM. Sensory responses from stimulation of the inferior Rolandic and Sylvian

regions in man. Journal of Neurosurgery 1983;59:119-130.

[891] Van Damme S, Crombez G, Eccleston C. Disengagement from pain: the role of

catastrophic thinking about pain. Pain 2004;107:70-76.

298

[892] Van Damme S, Crombez G, Eccleston C. Coping with pain: a motivational perspective.

Pain 2008;139:1-4.

[893] Van Damme S, Legrain V, Vogt J, Crombez G. Keeping pain in mind: a motivational

account of attention to pain. Neuroscience and biobehavioral reviews 2010;34:204-213.

[894] Van der Glas HW, Cadden SW, Van der Bilt A. Mechanisms underlying the effects of

remote noxious stimulation and mental activities on exteroceptive jaw reflexes in man.

Pain 2000;84:193-202.

[895] van Wijk G, Veldhuijzen DS. Perspective on diffuse noxious inhibitory controls as a model

of endogenous pain modulation in clinical pain syndromes. J Pain 2010;11:408-419.

[896] Villanueva L, Bouhassira D, Le Bars D. The medullary subnucleus reticularis dorsalis

(SRD) as a key link in both the transmission and modulation of pain signals. Pain

1996;67:231-240.

[897] Vogel H, Port JD, Lenz FA, Solaiyappan M, Krauss G, Treede RD. Dipole source analysis

of laser-evoked subdural potentials recorded from parasylvian cortex in humans. J

Neurophysiol 2003;89:3051-3060.

[898] Vogt BA. Pain and emotion interactions in subregions of the cingulate gyrus. Nat Rev

Neurosci 2005;6:533-544.

[899] Vogt BA. Regions and subregions of the cingulate cortex. In: BA Vogt, editor. Cingulate

neurobiology and disease. New York: Oxford University Press, 2009. pp. 5-30.

[900] Vogt BA. Regions and subregions of the cingulate cortex. In: BA Vogt, editor. Cingulate

Neurobiology and Disease. New York: Oxford University Press, 2009. pp. 1-30.

[901] Vogt BA, Pandya DN. Cingulate cortex of the rhesus monkey: II. Cortical afferents. J

Comp Neurol 1987;262:271-289.

[902] Vogt BA, Sikes RW. The medial pain system, cingulate cortex, and parallel processing of

nociceptive information. Prog Brain Res 2000;122:223-235.

[903] Vogt BA, Sikes RW. Cingulate nociceptive circuitry and roles in pain processing: the

cingulate premotor pain model. In: BA Vogt, editor. Cingulate neurobiology and disease.

New York: Oxford University Press, 2009. pp. 312-338.

299

[904] Vogt BA, Sikes RW, Vogt LT. Anterior cingulate cortex and the medial pain system. In:

BA Vogt, M Gabriel, editors. Neurobiology of cingulate cortex and limbic thalamus: A

comprehensive handbook. Boston: Birkhauser, 1993. pp. 313-344.

[905] Vogt BA, Vogt L, Farber NB, Bush G. Architecture and neurocytology of monkey

cingulate gyrus. J Comp Neurol 2005;485:218-239.

[906] Vogt BA, Vogt L, Laureys S. Cytology and functionally correlated circuits of human

posterior cingulate areas. Neuroimage 2006;29:452-466.

[907] Wade JB, Dougherty LM, Hart RP, Rafii A, Price DD. A canonical correlation analysis of

the influence of neuroticism and extraversion on chronic pain, suffering, and pain

behavior. Pain 1992;51:67-73.

[908] Wager TD, Rilling JK, Smith EE, Sokolik A, Casey KL, Davidson RJ, Kosslyn SM, Rose

RM, Cohen JD. Placebo-induced changes in FMRI in the anticipation and experience of

pain. Science 2004;303:1162-1167.

[909] Wager TD, Scott DJ, Zubieta JK. Placebo effects on human mu-opioid activity during

pain. Proc Natl Acad Sci U S A 2007;104:11056-11061.

[910] Wakana S, Caprihan A, Panzenboeck MM, Fallon JH, Perry M, Gollub RL, Hua K, Zhang

J, Jiang H, Dubey P, Blitz A, van ZP, Mori S. Reproducibility of quantitative

tractography methods applied to cerebral white matter. Neuroimage 2007;36:630-644.

[911] Walhovd KB, Westlye LT, Amlien I, Espeseth T, Reinvang I, Raz N, Agartz I, Salat DH,

Greve DN, Fischl B, Dale AM, Fjell AM. Consistent neuroanatomical age-related volume

differences across multiple samples. Neurobiol Aging 2009.

[912] Walker HK. Cranial Nerve V: The Trigeminal Nerve. In: HK Walker, WD Hall, JW Hurst,

editors. Clinical Methods: The History, Physical, and Laboratory Examinations. Boston,

1990.

[913] Wall PD. The laminar organization of dorsal horn and effects of descending impulses. J

Physiol 1967;188:403-423.

[914] Wall PD, Egger MD. Formation of new connexions in adult rat brains after partial

deafferentation. Nature 1971;232:542-545.

300

[915] Wall PD, Woolf CJ. Muscle but not cutaneous C-afferent input produces prolonged

increases in the excitability of the flexion reflex in the rat. J Physiol 1984;356:443-458.

[916] Walther K, Bendlin BB, Glisky EL, Trouard TP, Lisse JR, Posever JO, Ryan L. Anti-

inflammatory drugs reduce age-related decreases in brain volume in cognitively normal

older adults. Neurobiol Aging 2011;32:497-505.

[917] Wang K, Svensson P, Sessle BJ, Cairns BE, Arendt-Nielsen L. Painful conditioning

stimuli of the craniofacial region evokes diffuse noxious inhibitory controls in men and

women. J Orofac Pain 2010;24:255-261.

[918] Wang N, May PJ. Peripheral muscle targets and central projections of the mesencephalic

trigeminal nucleus in macaque monkeys. Anat Rec 2008;291:974-987.

[919] Wartolowska K, Hough MG, Jenkinson M, Andersson J, Wordsworth BP, Tracey I.

Structural brain changes in rheumatoid arthritis. Arthritis Rheum 2011.

[920] Watkins LR, Maier SF, Goehler LE. Immune activation: the role of pro-inflammatory

cytokines in inflammation, illness responses and pathological pain states. Pain

1995;63:289-302.

[921] Watkins LR, Milligan ED, Maier SF. Spinal cord glia: new players in pain. Pain

2001;93:201-205.

[922] Watson A, El-Deredy W, Iannetti GD, Lloyd D, Tracey I, Vogt BA, Nadeau V, Jones AK.

Placebo conditioning and placebo analgesia modulate a common brain network during

pain anticipation and perception. Pain 2009;145:24-30.

[923] Watts JW, Freeman W. Psychosurgery for the relief of unbearable pain. J Int Coll Surg

1946;9:679-683.

[924] Weddell G. Somesthesis and the chemical senses. Annu Rev Psychol 1955;6:119-136.

[925] Wedeen RP, Udasin I, Fiedler N, D'Haese P, De Broe M, Gelpi E, Jones KW, Gochfeld M.

Urinary biomarkers as indicators of renal disease. Ren Fail 1999;21:241-249.

[926] Weigelt A, Terekhin P, Kemppainen P, Dorfler A, Forster C. The representation of

experimental tooth pain from upper and lower jaws in the human trigeminal pathway.

Pain 2010;149:529-538.

301

[927] Weissman-Fogel I, Moayedi M, Taylor KS, Pope G, Davis KD. Cognitive and default-

mode resting state networks: do male and female brains "rest" differently? Hum Brain

Mapp 2010;31:1713-1726.

[928] Weissman-Fogel I, Moayedi M, Tenenbaum HC, Goldberg MB, Freeman BV, Davis KD.

Abnormal cortical activity in patients with temporomandibular disorder evoked by

cognitive and emotional tasks. Pain 2011;152:384-396.

[929] Whitsel BL, Favorov OV, Li Y, Quibrera M, Tommerdahl M. Area 3a neuron response to

skin nociceptor afferent drive. Cereb Cortex 2009;19:349-366.

[930] Whitsel BL, Petrucelli LM, Werner G. Symmetry and connectivity in the map of the body

surface in somatosensory area II of primates. J Neurophysiol 1969;32:170-183.

[931] Wiberg M, Westman J, Blomqvist A. The projection to the mesencephalon from the

sensory trigeminal nuclei. An anatomical study in the cat. Brain Res 1986;399:51-68.

[932] Wiech K, Kalisch R, Weiskopf N, Pleger B, Stephan KE, Dolan RJ. Anterolateral

prefrontal cortex mediates the analgesic effect of expected and perceived control over

pain. J Neurosci 2006;26:11501-11509.

[933] Wiech K, Lin CS, Brodersen KH, Bingel U, Ploner M, Tracey I. Anterior insula integrates

information about salience into perceptual decisions about pain. J Neurosci

2010;30:16324-16331.

[934] Wiech K, Ploner M, Tracey I. Neurocognitive aspects of pain perception. Trends Cogn Sci

2008;12:8.

[935] Wiech K, Seymour B, Kalisch R, Stephan KE, Koltzenburg M, Driver J, Dolan RJ.

Modulation of pain processing in hyperalgesia by cognitive demand. Neuroimage

2005;27:59-69.

[936] Wiech K, Tracey I. The influence of negative emotions on pain: Behavioural effects and

neural mechanisms. Neuroimage 2009;47:987-994.

[937] Wik G, Fischer H, Bragee B, Finer B, Fredrikson M. Functional anatomy of hypnotic

analgesia: a PET study of patients with fibromyalgia. Eur J Pain 1999;3:7-12.

[938] Wilkinson HA. Cingulotomy. J Neurosurg 2009;110:607.

302

[939] Wilkinson HA, Davidson KM, Davidson RI. Bilateral anterior cingulotomy for chronic

noncancer pain. Neurosurgery 1999;45:1129-1134.

[940] Williams ZM, Bush G, Rauch SL, Cosgrove GR, Eskandar EN. Human anterior cingulate

neurons and the integration of monetary reward with motor responses. Nat Neurosci

2004;7:1370-1375.

[941] Willis WD, Jr. The Pain System. The neural basis of nociceptive transmission in the

mammalian nervous system. In: PL Gildenberg, editor. Pain and Headache, Vol. 1. Basel:

S.Karger, 1985. pp. 1-346.

[942] Willis WD, Jr., Coggeshall RE. Sensory mechanisms of the spinal cord, Vol. 2. New York:

Plenum, 1991.

[943] Willis WD, Trevino DL, Coulter JD, Maunz RA. Responses of primate spinothalamic tract

neurons to natural stimulation of the hindlimb. J Neurophysiol 1974;37:358-372.

[944] Willis WD, Westlund KN. Neuroanatomy of the pain system and of the pathways that

modulate pain. J Clin Neurophysiol 1997;14:2-31.

[945] Wilson SAK. Croonian Lectures: On some disorders of motility and of muscle tone with

special reference to the corpus striatum. Lancet 1925;206:268-276.

[946] Witting N, Svensson P, Gottrup H, Arendt-Nielsen L, Jensen TS. Intramuscular and

intradermal injection of capsaicin: a comparison of local and referred pain. Pain

2000;84:407-412.

[947] Woda A. Pain in the trigeminal system: from orofacial nociception to neural network

modeling. J Dent Res 2003;82:764-768.

[948] Woda A, Azerad J, Guilbaud G, Besson JM. [A microphysiologic study of thalamic

projection of the cat's dental pulp (author's transl)]. Brain Res 1975;89:193-213.

[949] Wood PB, Glabus MF, Simpson R, Patterson JC, 2nd. Changes in gray matter density in

fibromyalgia: correlation with dopamine metabolism. J Pain 2009;10:609-618.

[950] Woolf CJ, Salter MW. Neuronal plasticity: increasing the gain in pain. Science

2000;288:1765-1769.

303

[951] Woolf CJ, Salter MW. Plasticity and pain: role of the dorsal horn. In: KL McMahon, M

Koltzenburg, editors. Textbook of Pain Philadelphia: Elsevier, 2006. pp. 91-106.

[952] Woolsey CN. Organization of somatic sensory and motor areas of the cerebral cortex. In:

HF Harlow, CN Woolsey, editors. Biological and biochemical basis of behavior.

Madison: University of Wisconsin Press, 1958.

[953] Woolsey CN, Erickson TC, Gilson WE. Localization in somatic sensory and motor areas

of human cerebral cortex as determined by direct recording of evoked potentials and

electrical stimulation. J Neurosurg 1979;51:476-506.

[954] Wright CI, Williams D, Feczko E, Barrett LF, Dickerson BC, Schwartz CE, Wedig MM.

Neuroanatomical correlates of extraversion and neuroticism. Cereb Cortex 2006;16:1809-

1819.

[955] Yamasaki H, Kakigi R, Watanabe S, Naka D. Effects of distraction on pain perception:

magneto- and electro-encephalographic studies. Brain Res Cogn Brain Res 1999;8:73-76.

[956] Yarkoni T, Poldrack RA, Nichols TE, Van Essen DC, Wager TD. Large-scale automated

synthesis of human functional neuroimaging data. Nature methods 2011;8:665-670.

[957] Yasui Y, Itoh K, Mizuno N, Nomura S, Takada M, Konishi A, Kudo M. The posteromedial

ventral nucleus of the thalamus (VPM) of the cat: direct ascending projections to the

cytoarchitectonic subdivisions. J Comp Neurol 1983;220:219-228.

[958] Yeung JC, Rudy TA. Sites of antinociceptive action of systemically injected morphine:

involvement of supraspinal loci as revealed by intracerebroventricular injection of

naloxone. J Pharmacol Exp Ther 1980;215:626-632.

[959] Yeung JC, Yaksh TL, Rudy TA. Effect on the nociceptive threshold and EEG activity in

the rat of morphine injected into the medial thalamus and the periaqueductal gray.

Neuropharmacology 1978;17:525-532.

[960] Yoshida A, Sessle BJ, Dostrovsky JO, Chiang CY. Trigeminal and dorsal column nuclei

projections to the anterior pretectal nucleus in the rat. Brain Res 1992;590:81-94.

[961] Young GB, Blume WT. Painful epileptic seizures. Brain 1983;106 (Pt 3):537-554.

[962] Younger JW, Chu LF, D'Arcy NT, Trott KE, Jastrzab LE, Mackey SC. Prescription opioid

analgesics rapidly change the human brain. Pain 2011;152:1803-1810.

304

[963] Younger JW, Shen YF, Goddard G, Mackey SC. Chronic myofascial temporomandibular

pain is associated with neural abnormalities in the trigeminal and limbic systems. Pain

2010;149:222-228.

[964] Yu XH, Zhang ET, Craig AD, Shigemoto R, Ribeiro-da-Silva A, De Koninck Y. NK-1

receptor immunoreactivity in distinct morphological types of lamina I neurons of the

primate spinal cord. J Neurosci 1999;19:3545-3555.

[965] Zambito Marsala S, Tinazzi M, Vitaliani R, Recchia S, Fabris F, Marchini C, Fiaschi A,

Moretto G, Giometto B, Macerollo A, Defazio G. Spontaneous pain, pain threshold, and

pain tolerance in Parkinson's disease. J Neurol 2011;258:627-633.

[966] Zarei M, Johansen-Berg H, Smith S, Ciccarelli O, Thompson AJ, Matthews PM.

Functional anatomy of interhemispheric cortical connections in the human brain. J Anat

2006;209:311-320.

[967] Zhang ET, Craig AD. Morphology and distribution of spinothalamic lamina I neurons in

the monkey. J Neurosci 1997;17:3274-3284.

[968] Ziegler DA, Piguet O, Salat DH, Prince K, Connally E, Corkin S. Cognition in healthy

aging is related to regional white matter integrity, but not cortical thickness. Neurobiol

Aging 2010;31:1912-1926.

[969] Zilles K. Neuronal plasticity as an adaptive property of the central nervous system. Ann

Anat 1992;174:383-391.

305

Appendices

306

Appendix I: Recruitment Letter

307

Appendix II: Consent forms

308

Consent Form

Title: Pain, Temperature and Tactile Perceptions in Normal Subjects and in

Patients with Neurological Disorders

Investigators: K.D. Davis, Tel.: 416-603-5662

Co-Investigators: H.C. Tenenbaum

M. Goldberg

Introduction:

You are being asked to take part in a research study. Please read this explanation about the study and its

risks and benefits before you decide if you would like to take part. You should take as much time as you

need to make your decision. You should ask the study doctor or study staff to explain anything that you

do not understand and make sure that all of your questions have been answered before signing this

consent form. Before you make your decision, feel free to talk about this study with anyone you wish.

Participation in this study is voluntary.

Purpose:

We are studying the brain mechanisms of touch, temperature, pain sensation and visual and auditory

perception. Therefore, we will assess your perception of various stimuli presented and also obtain images

of your brain with the magnetic resonance imaging (MRI) scanner during the stimulation.

Procedures:

There are two parts to the project, which may take place on separate days or a single day, and will require

approximately 2 hours of your time in total. In the first part, we will ask about your health, your pain and

also conduct some very basic personality and cognitive tests to measure your brain’s ability to process

information. This will take approximately 30-40 minutes. The second part includes some additional

309

questionnaires and a MRI (magnetic resonance imaging) scan (at the Toronto Western Hospital) to take

pictures of your brain while lying still and also while viewing words that may be related to your pain. The

MRI will take approximately 1-1.5 hours.

As part of these procedures you will be trained to rate the intensity of a stimulus and to describe its

quality. In some sessions the ratings will consist of a verbal or written indication of the stimulus. In other

sessions you may be trained to use a computer to rate the stimulus. You may also be asked to watch visual

images or listen to sounds. You may have your skin marked with water soluble ink in studies in which we

deliver mechanical, thermal and/or electrical stimuli to a small area of your skin. Following each stimulus

you will be asked for a rating. For those studies, mechanical (pressure) stimuli will consist of one or more

of several types such as slender nylon filaments, cotton swabs, tuning forks and pins. The thermal stimuli

will be delivered with a probe that is precisely controlled and may be warm, cold or painful. You may be

asked to place you hand into a cold water bath. The electrical stimuli will be delivered by electrodes

placed on your skin. The visual images will be presented on a screen and the auditory stimuli will be

heard through headphones.

To image your brain, you will be placed into the narrow tube of the MRI machine. The MRI makes loud

banging noises when it is taking pictures. At any time, if you are experiencing distress, you can freely

communicate with the MR technician through a microphone in the MRI. You are free to terminate the

procedure if you feel extreme claustrophobia.

Risks:

There are no major risks associated with the experimental protocols other than the discomfort of the

painful stimuli. At any time during these sessions, you are free to withdraw from any stimulus and are

free to terminate a session.

Benefits:

You will not receive any direct benefit from being in this study. Information learned from this study may

help other people with chronic pain in the future.

Voluntary Participation:

Your participation in this study is voluntary. You may decide not to be in this study, or to be in the study

now and then change your mind later. You may leave the study at any time without affecting your current

or future medical care.

310

Confidentiality:

If you agree to join this study, the study doctor and his/her study team will look at your personal health

information and collect only the information they need for the study. Personal health information is any

information that could be used to identify you and includes your:

• name, • address, • date of birth, • new or existing medical records, that includes types, dates and results of medical tests or

procedures.

The information that is collected for the study will be kept in a locked and secure area by the study doctor

for 7 years. Only the study team or the people or groups listed below will be allowed to look at your

records. Your participation in this study also may be recorded in your medical record at this hospital.

Representatives of the Mount Sinai Hospital Research Ethics Board may look at the study records and at

your personal health information to check that the information collected for the study is correct and to

make sure the study followed proper laws and guidelines.

All information collected during this study, including your personal health information, will be kept

confidential and will not be shared with anyone outside the study unless required by law. You will not be

named in any reports, publications, or presentations that may come from this study.

If you decide to leave the study, the information about you that was collected before you left the study

will still be used. No new information will be collected without your permission.

In Case You Are Harmed in the Study:

If you become ill, injured or harmed as a result of taking part in this study, you will receive care.

The reasonable costs of such care will be covered for any injury, illness or harm that is directly a

result of being in this study. In no way does signing this consent form waive your legal rights nor

does it relieve the investigators, sponsors or involved institutions from their legal and

professional responsibilities. You do not give up any of your legal rights by signing this consent

form.

Expenses Associated with Participating in the Study:

311

You will receive remuneration for your time at approx. $30-40/hour or $75/study.

Questions:

If you have any questions regarding these procedures please contact Dr. Karen D. Davis at 416-603-5662.

For any questions regarding your rights as a research subject please contact Dr. Heslegrave, (Chair, UHN

and Mount Sinai Hospital Research Ethics Boards) at 416-340-4557 or 416-586-4875. This person is not

involved with the research project in any way and contacting him will not affect your participation in the

study.

Consent:

This study has been explained to me and any questions I had have been answered.

I know that I may leave the study at any time. I agree to take part in this study.

Print Study Participant’s Name Signature Date

(You will be given a signed copy of this consent form)

My signature means that I have explained the study to the participant named above. I have answered all

questions.

Print Name of Person Obtaining Consent Signature Date

312

Appendix III: Edinburgh Handedness Inventory

Edinburgh Handedness Inventory Page 1 of 1

Handedness Questionnaire

Instructions

Please indicate your preferences in the use of hands in the following activities. If you are really indifferent, select “Either”. Where the preference is so strong that you would never try to use the other hand select “No”.

When: Which hand do you prefer? Do you ever use the other hand?

Writing: L R Either Yes No Drawing: L R Either Yes No

Throwing: L R Either Yes No Using scissors: L R Either Yes No

Using a toothbrush: L R Either Yes No Using a knife (without fork): L R Either Yes No

Using a spoon: L R Either Yes No Using a broom (upper hand): L R Either Yes No

Striking a match: L R Either Yes No Opening a box (lid): L R Either Yes No

Thank you for your responses.

Adapted from: Mark S. Cohen

Based on: Oldfield, R.C. 1971

313

Appendix IV: Pain Catastrophizing Scale

< Copyright � 1995 Michael JL Sullivan

PCS Client No.: __________ Age: _____ Sex: M(__) F(__) Date: _______________

Everyone experiences painful situations at some point in their lives. Such experiences may include headaches, tooth pain, joint or muscle pain. People are often exposed to situations that may cause pain such as illness, injury, dental procedures or surgery. We are interested in the types of thoughts and feelings that you have when you are in pain. Listed below are thirteen statements describing different thoughts and feelings that may be associated with pain. Using the following scale, please indicate the degree to which you have these thoughts and feelings when you are experiencing pain. 0 – not at all 1 – to a slight degree 2 – to a moderate degree 3 – to a great degree 4 – all the time When I’m in pain …

1� I worry all the time about whether the pain will end.

2� I feel I can’t go on.

3� It’s terrible and I think it’s never going to get any better.

4� It’s awful and I feel that it overwhelms me.

5� I feel I can’t stand it anymore.

6� I become afraid that the pain will get worse.

7� I keep thinking of other painful events.

8� I anxiously want the pain to go away.

9� I can’t seem to keep it out of my mind.

10� I keep thinking about how much it hurts.

11� I keep thinking about how badly I want the pain to stop.

12� There’s nothing I can do to reduce the intensity of the pain.

13� I wonder whether something serious may happen.

…Total

314

Appendix V: TMD pain assessment

Date: ……./……./…….

Patient Name:

Patient No.:

Date of Birth: ……./……./…….

What is your highest level of Education:

Phase of the menstrual cycle

What day did your last period start: ……./……./…….

Oral contraceptive Y N

Site and Radiation:

ES HS UNI M

1 2 3 4

Preauricular

Intra-auricular

Temporal

Massetric

315

1. How long have you had TMD? _________________________

2. Describe how your pain feels? _________________________

3. McGill Pain Questionnaire

4. Frequency: Constant-………….

Occasional- ………………… per week

………………… per day

Pain History:

Worse in a.m.? Y N

Worse in p.m.? Y N

Responsive to analgesics? Y N

Current medication/treatment

Other Pain Conditions?

Aggravating factors?

Relieving factors?

Y N _______________________________________

Y N arthritis, fibromyalgia, IBS _____________________

Y N eating, chewing, yawning, talking,____________________

Y N rest, warmth, cold, chewing,__________________________

316

Other symptoms?

Ear symptoms?

Sleep affected?

Clenching/Bruxism

Do you have pain in regions in other parts of your body?

Current Pain Rating: NPS 0 (no pain)-10 (most intense pain imaginable)

Intensity- _________

NPS (0 (not unpleasant)-10 (most unpleasant pain imaginable)

Unpleasantness- ________

History of symptoms: Since onset, pain has improved, stayed the same, worsened?

Has the nature of pain changed since onset?

Y N clicking, locking, limited mouth opening, _________________

Y N pop ________________________________________________

Y N

Y N Prevention, disturbance, wake early a.m.______________________

Y N Specify____________________________________________

317

Appendix VI: McGill Pain Questionnaire

318

Appendix VII: McGill Pain Questionnaire results for Patients with TMD

Table A-1: McGill Pain Questionnaire result for patients with TMD

Abbreviations for “Evaluative” words selected: A – annoying; T – troublesome; M – miserable; I – intense; U - unbearable

Sub

Number of Words Pain

Intensity Sensory

(Q.1-10)

Affective

(Q.11-15)

Evaluative

(Q.16)

Miscellaneous

(Q17-20)

Total

1 10 1 1 I 3 15 1

2 7 0 2 AT 1 10 2

3 5 0 0 1 6 2

4 7 2 3 ATM 1 13 3

5 7 1 0 2 10 2

6 10 1 3 ATI 4 18 2

7 9 1 1 A 4 15 1

8 13 1 3 TIU 4 21 2

9 12 2 4 AMIU 5 23 2

10 10 5 4 ATMI 5 24 2

11 10 0 0 1 11 1

12 7 2 2 AT 4 15 3

13 11 0 0 4 15 2

14 5 4 2 AM 1 12 2

15 11 2 1 T 2 16 2

16 11 2 1 A 5 19 2

17 10 0 3 ATI 4 17 1

319

Copyright Acknowledgements

320

We acknowledge the following copyright permissions:

The American Association for the Advancement of Science for permission to reprint the figure

“Wall and Melzack’s Gate Control Theory of Pain mechanism” (Figure 2-2) from [595].

Oxford University Press for permission to reprint the figure “Diffusion tensor imaging

reconstruction of Meynert’s classification of white matter tracts in the brain” (Figure 2-4) from

[147].

Elsevier Limited for permission to reprint 5 figures (Figures 4-1, 4-2, 4-3, 4-4 and 4-5) from

[198].

Elsevier Limited for permission to reprint the published articles: [617]:

Moayedi M, Weissman-Fogel I, Crawley AP, Goldberg MB, Freeman BV, Tenenbaum HC,

Davis KD. Contribution of chronic pain and neuroticism to abnormal forebrain gray matter in

patients with temporomandibular disorder. Neuroimage 2011; 55(1): 277-86

and

Moayedi M, Weissman-Fogel I, Salomons TV, Crawley AP, Goldberg MB, Freeman BV,

Tenenbaum HC, Davis KD. Abnormal gray matter aging in chronic pain patients. Brain

Research 2012; 1456: 82–93.