Monitoring and Analysis of Hemodynamic Response to Cold Noxious Stimuli

Using Functional Near Infrared Spectroscopy

A Thesis

Submitted to the Faculty

of

Drexel University

by

Zeinab Barati

In partial fulfillment of the

requirements for the degree

of

Doctor of Philosophy

July 2013

© Copyright 2013

Zeinab Barati. All Rights Reserved.

i

Dedications

To my beloved parents for their unconditional love and support

And to my husband for his love, patience, and encouragement

ii

Acknowledgements

I would like to first thank my supervisor Dr. Kambiz Pourrezaei for his guidance and support in the course of completing this research. It was a great honor to work under his supervision. His support was far beyond supervising a research project. I learned life- lasting lessons through his thoughtful critiques and recommendations that will undoubtedly help me in my future career.

I would like to extend my gratitude to my co-adviser Dr. Issa Zakeri who generously devoted his time educating me on advanced statistical techniques and supervised the data analyses.

I am also thankful to Dr. Patricia Shewokis for her guidance in the statistical analysis during the first couple of years of this PhD research. I also thank the committee members

Dr. Arye Rosen, Dr. Joshua Jacobs, and Dr. John Grothusen for their time and effort to serve on my committee and review my dissertation.

I am very thankful to our lab‟s faculty members Dr. Meltem Izzetoglu and Dr. Hasan

Ayaz for their help and guidance with signal processing and hardware.

I also would like to thank my lab colleagues Rutvi Vyas, Kang Hee Lee, Daryl Omire-

Mayor, and Esther Lim for their help in hardware and experimentation and more importantly, making the long hours of lab work much easier and enjoyable. iii

I would like to appreciate all of my friends who supported me emotionally during my graduate studies in Philadelphia. My very special thanks go to Marjan Saboktakin, Saba

Borujerdi, Zahra Javaheri, Sedigheh Banihashemi, Somayeh Yousefi, Shokouh Pourarian and all my friends at Mahdieh community.

I would like to appreciate the financial support from the School of Biomedical

Engineering, Science, and Health Systems and especially, I would like to thank Dr. Banu

Onaral for her support and guidance.

And at the end, I am very thankful to my family and specially my husband for their encouragement and love.

iv

Table of Contents

List of Tables ______vii

List of Figures ______viii

ABSTRACT ______xiii

CHAPTER 1: INTRODUCTION ______1

1.1. Background ______1

1.2. Problem Statement ______5

1.3. Objectives ______6

1.4. Specific Aims and Research Hypotheses ______6

1.5. Significance ______7

1.6. Outline ______8

CHAPTER 2: HEMODYNAMIC RESPONSE TO REPEATED NOXIOUS COLD PRESSOR TESTS MEASURED BY FUNCTIONAL NEAR INFRARED SPECTROSCOPY ON FOREHEAD ______9

2.1. Abstract ______9

2.2. Introduction ______10

2.3. Materials and Methods ______13

2.3.1 fNIRS Principles and Instrumentation ______13

2.3.2 Subjects ______15

2.3.3 Protocol ______16

2.3.4 fNIR Data Processing ______18

2.4. Results ______19

2.4.1. The effect of a CPT on the THb concentration ______21

2.4.2. Adaptation to repeated CPTs observable in subjective ratings and THb ___ 24 v

2.4.3. Correlation analysis between ΔTHb and subjective pain scores ______30

2.4.4. Comparing ΔTHb response at the „far channels‟ with the „near channels‟ across left and right sides of the forehead ______30

2.5. Discussion ______33

2.6. Acknowledgement ______37

CHAPTER 3: FUNCTIONAL DATA ANALYSIS VIEW OF FNIRS DATA ______38

3.1. Abstract ______38

3.2. Introduction ______39

3.3. Materials and Methods ______43

3.3.1. Participants ______43

3.3.2. Protocol ______43

3.3.3. Measurements ______44

3.3.4. Functional Data Analysis ______45

3.4. Results ______50

3.4.1. fNIR Data Pre-processing ______50

3.4.2. Results of fPCA ______54

3.4.3. Rotating PCs ______58

3.4.4. Results of fCCA ______62

3.5. Discussion ______64

3.6. Acknowledgment ______69

CHAPTER 4: HEMODYNAMIC RESPONSE TO COLD TOLERANCE TESTS AT DIFFERENT TEMPERATURES ______70

4.1. Introduction ______70

4.2. Methods ______71

4.2.1. Subjects ______71

4.2.2. Instrument ______72 vi

4.2.3. Protocol ______72

4.2.4. Features ______73

4.2.5. Statistical Data Analysis ______74

4.3. Results ______80

4.3.1. Analysis of maximum reported pain, pain threshold and ______80

4.3.2. Statistical Analysis of Hb and HbO2 ______84

4.4. Discussion ______103

4.5. Conclusion ______106

CHAPTER 5: SUMMARY, FUTURE WORK, AND CLINICAL APPLICATIONS ______108

5.1. Summary ______108

5.2. Future work ______110

5.3. Clinical Applications ______111

List of References ______113

VITA ______123

vii

List of Tables

1. Results of repeated measures ANOVA on THb during the four experimental conditions: baseline, pre-stimulus, stimulus, and post-stimulus conditions……………..22

2. Pairwise condition comparisons across the Forehead Side and Channel including 95% confidence intervals and effect sizes for THb , comparing the four conditions: the baseline, 1st pre-stimulus, 1st stimulus, and 1st post-stimulus conditions…………....…23

3. Main effect of (CPT) trials repeated measures ANOVAs on THb ……………………………………………………………………………………...26

4. Pairwise cold pressor test (CPT) trials comparisons across the Forehead Side and Channel including 95% confidence intervals and effect sizes for THb …………………27

5. Within-subjects correlation values between self-reported pain rating scores and ΔTHb……………………………………………………………………………………..30

6. Pairwise channel pair comparisons across cold pressor test (CPT) trials including 95% confidence intervals and effect sizes for ΔTHb………………………………………….32

7. Regression Coefficients……………………………………………………..………...83

8. Type III tests of model covariates for the dependent variable delta_Hb………..…….88

9. Type III tests of model covariates for the dependent variable delta_HbO2…………..93

10. Type III tests of model covariates for the dependent variable MV……………….....98

11. Type III tests of model covariates for the dependent variable MA………………...102

viii

List of Figures

1. A schematic of the fNIRS probe configuration. Photons travel from a light source to a photodetector through a banana shape pathway with a penetration depth of half the source-detector distance. Measures are approximate…………………………...………..15

2. A demonstration of the placement of fNIRS probes on a subject‟s forehead. The probes are shown reversed to illustrate the location of the light source and photodetectors……………………………………………………………………………17

3. Block diagram of the protocol…………………………………………………...……17

4A-B. The hemodynamic response to three trials of cold pressor test (CPT) averaged across 20 subjects on the right (a) and left (b) sides of the forehead. The black vertical lines represent time points at which subjects switched their hand from the tepid water to the ice water and vice versa. The pink shaded field represents the baseline period, the orange shaded fields represent the immersion in tepid water condition, and the blue shaded fields represent the CPT conditions. Hb, HbO2 and THb are abbreviations for deoxy-hemoglobin, oxy-hemoglobin and total hemoglobin, respectively. „S-D‟ symbol stands for „source-detector‟……………………………………………………………....20

5. Line graphs of THb during four experimental conditions: baseline, pre-stimulus, stimulus, and post-stimulus conditions (for the definition of THb , please see section 5.1). Error bars represent the 95% confidence intervals (CI)…………………………….……22

6. Boxplots of post-stimulus pain rating scores of 20 subjects across three trials of cold pressor test (CPT). Two outliers were identified for pain scores reported after 3rd trial of CPT and are represented as circles……………………………………………..….…….25

7. Line graphs of THb across cold pressor test (CPT) trials (for the definition of THb , please see section 5.1)…………………………………………………..………….….…25 ix

8. Line graphs of THb for separate trials of cold pressor test (CPT) (for the definition of THb , please see section 5.1). Significant interactions between the Forehead Side and Channel were found for the 1st and 2nd CPT trials…………………………………...…26

9. Three sample total hemodynamic responses to the first cold pressor test (CPT) measured at the „far channel 1‟ on the right side of the forehead. The first black vertical line represents the time point at which subjects immersed their hand into the ice water and the second black vertical line shows the time when subjects immersed their hand back into the tepid water. Dashed blue vertical lines represent the detected time points for ΔTHb calculation (see 5.3 section)………………………………………………………29

10. Line graphs of ΔTHb for separate trials of cold pressor test (CPT) (for the definition of ΔTHb, please see Figures 9A-C). For all three trials of CPT, significantly different activation on the left and right sides of the forehead was observed in ΔTHb as measured by the „far channels‟, while no laterality was observed in ΔTHb as measured by the „near channels‟……………………………………………………………………………...….32

11. Non-smooth and smooth Hb and HbO2 data for 20 subjects for a „far‟ channel located on right forehead: a) non-smooth HbO2; b) smooth HbO2; c) non-smooth Hb; d) smooth Hb. The smoothing parameter was 104. The black dashed line represents the mean response across 20 subjects. The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus)…………………………………………………………….………52

12. Non-smooth and smooth Hb and HbO2 data for 20 subjects for a „far‟ channel located on left forehead: a) non-smooth HbO2; b) smooth HbO2; c) non-smooth Hb; d) smooth Hb. The smoothing parameter was 104. The black dashed line represents the mean response across 20 subjects. The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus)……………………………………………….……………………53

13. The first four principal components (PC) of HbO2 (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) „far‟ channel. In (a) and (b), x blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post- stimulus)………………………………………………………………………………….56

14. The first four principal components (PC) of Hb (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) „far‟ channel. In (a) and (b), blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post- stimulus)………………………………………………………………………………….57

15. The first four rotated principal components (PC) of HbO2 (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) „far‟ channel. In (a) and (b), blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus)………………………………………………………….…….…...60

16. The first four rotated principal components (PC) of Hb (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) „far‟ channel. In (a) and (b), blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the xi hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus)…………………………………………………………..………...61

17. Canonical correlation analysis of Hb (blue) and HbO2 (red) data for a pair of „far‟ channels (a – right, b – left) and the pair of „near‟ channels (c – right, d – left). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus)…….....………....63

18. Block diagram of the second protocol……………………………………..…….…..73

19. Scatter plot of delta_HbO2 across temperature for 7 subjects. Temperature is considered a continuous factor…………………………………………………………...76

20a-c. Box plots of (a) pain threshold, (b) pain tolerance, and (c) maximum reported pain score during the CPT across four temperatures: 1C, 5C, 10C, and 15C for 21 subjects……………………………………………………………………………...……81

21. Sample of Hb and HbO2 time series from a far channel on the right side of forehead during a cold pressor test (CPT) for four temperatures. The black vertical lines from left to right indicate the baseline recording, hand immersion in the tepid water, hand immersion in cold water, and reaching the pain tolerance……………………………….84

22a-b. Scatter plot (a) and line plot (b) of the dependent variable delta_Hb versus temperature………………………………………………………………………………86

23a-c. Line plots of the dependent variable delta_Hb versus temperature for the fixed factors Channel (a), Side (b), and Gender (c)……………………………………...... …87

24. Line plots of the dependent variable delta_Hb versus temperature showing the interaction between Channel and Gender………………………………………………..88

25a-b. Scatter plot (a) and line plot (b) of the dependent variable delta_HbO2 versus temperature………………………………………………………………………...…….90 xii

26a-c. Line plots of the dependent variable delta_HbO2 versus temperature for the fixed factors Channel (a), Side (b), and Gender (c)………………………………...……….....91

27. Line plots of the dependent variable delta_HbO2 versus temperature showing the interaction between Channel and Side………………………………………………...…92

28. Line plots of the dependent variable delta_HbO2 versus temperature showing the interaction between Channel and Gender……………………………………………..…92

29. Line plots of the dependent variable delta_HbO2 versus temperature showing the three-way interaction between Channel, Side, and Gender………………...………...... 93

30a-b. Scatter plot (a) and line plot (b) of the dependent variable MV versus temperature………………………………………………………………………………95

31a-c. Line plots of the dependent variable MV versus temperature for the fixed factors Channel (a), Side (b), and Gender (c)………………………………………………....…96

32. Line plots of the dependent variable MV versus temperature showing the interaction between Channel and Side………………………………………………………..……...97

33. Line plots of the dependent variable MV versus temperature showing the interaction between Channel and Gender…………………………………………...... 97

34a-b. Scatter plot (a) and line plot (b) of the dependent variable MA versus temperature………………………..……………………………………………………..99

35a-c. Line plots of the dependent variable MA versus temperature for the fixed factors Channel (a), Side (b), and Gender (c)………………………….…………………….....100

36. Line plots of the dependent variable MA versus temperature showing the interaction between Channel and Side………………………………..……………..……………...101

37. Line plots of the dependent variable MA versus temperature showing the interaction between Channel and Gender………………….…………………………………...…..101 xiii

ABSTRACT Monitoring and Analysis of Hemodynamic Response to Cold Noxious Stimuli Using Functional Near Infrared Spectroscopy Zeinab Barati

The gold standard for evaluating the presence, intensity, quality and location of pain is self-reporting questionnaires, if communication is not impaired. These questionnaires are highly subjective. Thus, there remains an unmet clinical need for a practical, inexpensive tool for the reliable and objective assessment of human response to pain.

Due to inherent interaction between the autonomic nervous system (ANS) and the central pain processing system, we attempted to indirectly assess the intensity of ongoing pain by measuring the ANS response to painful stimuli using functional near infrared spectroscopy (fNIRS). fNIRS can be used for noninvasive, continuous monitoring of the hemodynamic activity on the cortex as well as skin. The objective of this PhD research is to investigate the feasibility of employing fNIRS technology for objective assessment of pain in human through the hemodynamic parameters measured by fNIRS at the skin and cortex in response to noxious stimuli.

We used cold pressor test (CPT) as an experimental model of . We used three consecutive 45-min trials of a CPT. We identified a parameter from fNIRS signal that significantly correlated with subjective pain scores. In another experiment, subjects immersed their hand into cold water of varying temperatures for as long as they could tolerate the cold pain/unpleasantness. Using multi-distance probes, we observed an asymmetrical activity in the prefrontal cortex and no laterality in the skin response. This finding suggests that while the generalized autonomic response is global, there is a xiv hemispheric asymmetry in the prefrontal cortex during cold water stimulation. Moreover, we observed sex-specific cortical activation which suggests sex differences in central processing/regulating of nociceptive information.

In this research, we showed that a cold painful stimulus evokes a reproducible and consistent hemodynamic response which can be reliably detected by fNIRS as a non- invasive, portable device. We found an association between the and the evoked hemodynamic response. Further refinement of the proposed method, including incorporating advanced signal processing techniques and employing other noxious stimuli, in order to extract pure cortical response is required to make the fNIRS technique a powerful clinical tool for pain assessment.

1

CHAPTER 1: INTRODUCTION

1.1. Background

The International Association for the Study of Pain (IASP) defines pain as “an unpleasant sensory and emotional experience associated with actual or potential tissue damage, or described in terms of such damage” (1). Pain serves as an alarm to prevent possible damage to the organism, and is often a symptom of an injury or disease. When pain persists past the normal healing period, it becomes transformed to a new disease, the so called “chronic pain” with devastating psychophysical and psychosocial consequences.

Pain is the most frequently encountered symptom in daily medical practice. In the United

States, the incidence of pain is more than diabetes, heart diseases and cancer combined

(2). According to the Institute of Medicine of the National Academies report, the prevalence of chronic pain in adult Americans was at least 100 million in 2011 (3). The total annual cost of healthcare, lost wages and other expenses associated with pain ranges from $560 billion to $635 billion in the United States (2).

Pain perception and expression is highly subjective. The gold standard for evaluating the presence, intensity, quality and location of pain is self-reporting questionnaires if the patient is capable of reliable communication. These questionnaires, which are widely used in pain clinics, are highly subjective on the part of the patient as well as on the part of the healthcare providers - especially when there is more than one provider. This 2 shortcoming of pain questionnaires may result in under-prescription or over-prescription of drugs with their associated complications.

In conventional pain practice, physiological parameters (such as , , respiratory, galvanic skin response and cutaneous blood flow) and behavioral measures (such as facial grimacing and guarding the painful area) have been used to monitor the response to a noxious stimulus. However, these observations change due to many factors other than pain such as distress, medications, and illness. Thus, the clinical experience has proven that they are not practically reliable and should be interpreted cautiously (4). As such, available measures for the assessment of pain are subjective and poorly informative for design of correlative studies or adequate diagnosis and treatment of pain disorders.

Technological advancements have offered a number of novel solutions for the assessment of pain. Neuroimaging studies using advanced imaging modalities such as PET (positron emission tomography) and fMRI (functional magnetic resonance imaging) have revealed brain regions activated during a physical or psychological experience of pain (5-9). Some research has found a relation between subjects‟ report of an ongoing pain and BOLD

(blood oxygen level dependent) signal acquired by fMRI (7). These modalities have advanced our understanding of the underlying mechanisms in and have had great impact on basic science. However, clinical practice has not accepted them for routine examinations for several reasons. Firstly, they are not cost-effective and thus, not suitable for everyday applications. Secondly, they are not user-friendly and require trained personnel to operate them and therefore, not practical for large-scale studies. 3

Finally, their poor temporal resolution and noisy and enclosed environment limits the experimental manipulation.

Due to inherent interaction between the autonomic nervous system (ANS) and the central pain processing system, researchers have attempted to indirectly assess the intensity of ongoing pain by measuring the ANS response to painful stimuli. The ANS is a subdivision of peripheral nervous system and has two branches itself: sympathetic nervous system (SNS) and parasympathetic nervous system. It generally monitors the internal environment of the body and regulates the physiological parameters in order to maintain homeostasis. It is known that the ANS is actively involved in the generation and maintenance of pain (10, 11).

Function of the ANS can be assessed through monitoring electrodermal and cardiovascular reactivity in response to an external stimulus. Several challenges have been proposed to evoke the ANS such as Valsalva maneuver, CO2 inhalation, isometric hand grip, and cold pressor test (CPT). A CPT is a conventional test performed by immersing a limb into a cold water (usually freezing water) container for a specific period of time. A CPT evokes a generalized sympathetic activation and induces a tonic pain. It was first used by Hines and Brown in 1932 to experimentally raise blood pressure for the study of hypertension (12) and since then, it has been widely used in research involving psychological, cardiovascular, and neurological disorders. The application of

CPT for inducing experimental pain in healthy adults was initially introduced by Wolf and Hardy in 1941 (13).

The hemodynamic response to the CPT may provide the means for an objective assessment of the induced tonic pain. Current methods of measuring these hemodynamic 4 modulations have not yet identified an objective biomarker of pain that can be practically applied in clinical settings. A reason for this may be the limitations of current technologies.

The effect of sympathetic activation evoked by CPT on the cerebral hemodynamics has been investigated by using transcranial laser Doppler sonography for the assessment of the cerebral blood flow velocity in the middle cerebral arteries (14, 15). Sympathetically mediated changes in capillary blood flow and skin microcirculation during local cooling were studied using laser Doppler flowmetry (16). Laser Doppler utilizes low power lasers that can interrogate the outermost 0.5 to 1 mm depth of the skin. Single point laser

Doppler systems deliver good temporal but poor spatial resolution and the measured change in blood flow greatly depends on the position of the probe relative to the location of the affected arterioles and venules (17). Another shortcoming of single point laser

Doppler is its sensitivity to motion artifacts (17, 18). Recently, laser Doppler imaging has been introduced which offers much greater spatial resolution compared to single point laser Doppler flowmetry. However, its temporal resolution is limited and the cost of the laser Doppler imaging equipment limits its routine application in clinical settings (14).

Recent developments of functional near infrared spectroscopy (fNIRS) suggest that it can be used for noninvasive, continuous monitoring of tissue oxygenation and regional blood flow (19). Using fNIRS technology, changes in the hemodynamics can be measured simultaneously across different sites of sympathetic innervations such as hand, forearm, and face. In addition, the fNIRS technique allows noninvasive monitoring of cerebral hemodynamic changes induced by SNS and/or nociceptive activity. Recently, a few research studies suggested the use of fNIRS for monitoring cortical activation in response 5 to noxious stimuli in new-born infants (20, 21), healthy adults (22, 23), patients undergoing cardiac surgery (24), and individuals from migraines (25). Since an fNIRS system with multi-distance probes can penetrate as far as 3cm into the tissue, it enables the simultaneous signal acquisition from multiple layers, thus effectively eliminating the LDF shortfall. In fact, the fNIR technique provides a platform with the ability to study nociceptive responses of the skin, muscle, and cerebral cortex. Using the commonly used source-detector (S-D) separation distance of 2.5 to 4 cm, fNIRS can measure changes in blood oxygenation parameters within different layers up to a depth of

1.25 to 2 cm, respectively (26). fNIRS can be portable and has low equipment and maintenance costs. It is relatively robust to motion artifacts and therefore, no immobilization is required during measurement, unlike movement constraints imposed by other functional imaging techniques.

1.2. Problem Statement

In spite of advances in medical imaging technology that significantly help basic science, there remains an unmet clinical need for a practical, inexpensive tool for the reliable and objective assessment of human response to pain. Advanced functional imaging modalities such as functional MRI and PET scans deliver superior spatial information which comes at high equipment and maintenance price. Therefore, they are not readily accessible for routine clinical use. They are also very sensitive to motion artifact and require subjects/patients proper immobilization throughout the procedure which may be very difficult to attain in specific populations such as the youth and elderly. Other modalities for the assessment of ANS function that is recognized to be related to nociception have poor sensitivity and specificity. 6

Currently, there is no practical method available for an objective assessment of pain and clinicians are basically relying on subjective self-report measures using limited scales.

Beside the subjective nature of pain scales, their applicability for explaining different types and origins of pain is questionable. Also, in certain patients populations self-report cannot be obtained due to impaired or primitive communication abilities such as the elderly and infants.

1.3. Objectives

The primary objective of this PhD research is to investigate the feasibility of employing fNIRS technology for objective assessment of pain in human. The ultimate goal of this project is to identify a robust biomarker of ongoing pain through hemodynamic parameters measured by fNIRS at the skin, muscle, and cortex in response to noxious stimuli.

1.4. Specific Aims and Research Hypotheses

The specific aims of this research project were as follows:

Specific aim 1: To determine the efficacy of fNIRS in assessing the blood flow changes at the skin, muscle, and cortex in healthy subjects in response to noxious stimuli.

Hypothesis 1: fNIRS as a powerful tool for measuring regional blood flow

changes is capable of effective assessment of SNS-mediated hemodynamic

reactivity. Using multi-distance probes, fNIRS can monitor hemodynamic

changes at various depths.

Specific aim 2: To objectively investigate pain index for different intensities of painful stimuli. 7

Hypothesis 2: Due to interaction between the SNS and pain processing system,

quantitative assessment of sympathetic nervous response to different intensities of

pain may help develop a pain index.

Specific aim 3: To study individuals‟ longitudinal response to a tonic pain.

Hypothesis 3: Pain response including subjective rating scores and objective

measures acquired with fNIRS change over the course of a prolonged noxious

stimulus. Longitudinal studies can adequately evaluate an individual‟s response

trajectory.

Specific aim 4: To investigate the SNS-driven cerebral hemodynamic response and the skin‟s sympathetic nerve activity during nociception.

Hypothesis 4: The intense generalized sympathetic activation in response to an

acute pain is a global reaction while the cortical processing of the stimulus is

hypothesized to be a secondary effect. Evaluation of the skin, muscle, and cortex

response may help investigate the share of these two systems‟ involvement.

1.5. Significance

This research will provide a new approach in pain assessment. fNIRS signal collected simultaneously from multiple regions and tissue layers will be incorporated for objective assessment of ongoing pain. Such system will minimize other imaging modalities limitations such as low penetration depth, low sensitivity, ionizing radiation, and rare accessibility. With further refinement of this technology, the proposed technique would become an indispensable adjunct in all pain treatment facilities for routine diagnostic work ups and treatment efficacy assessments, including clinical trials of new 8 medications. Therefore, we envision for fNIRS technology to have a potentially decisive impact upon pain research and acceleration of new pain medication development.

1.6. Outline

Second chapter of this dissertation reports on the results of using three consecutive trials of CPT in control subjects. We examined the correlation between individuals‟ subjective reports and the evoked hemodynamic responses to noxious stimuli. We identified a parameter from fNIRS signal that significantly correlated with subjective pain scores.

Third chapter explored variability of fNIRS signals during a CPT using a novel data analysis technique, i.e. functional data analysis (fDA). fDA provides a new platform for the analysis of fNIRS data and gives qualitative and quantitative information of the shape and weight of main components of variability of fNIRS signals. Results of this investigation helped us in study of response dynamics and characterization of the evoked response in the next set of experiments.

Fourth chapter reports on the pain tolerance experiments that were optimized by changing the setting of the cold water stimulus including the intensity and duration. We employed four different temperatures of cold water to induce mild, moderate, and intense pain in order to study the hemodynamic reactivity to different levels of a noxious stimulus.

Fifth chapter concludes the dissertation by summarizing the results and giving some suggestions for future work.

9

CHAPTER 2: HEMODYNAMIC RESPONSE TO REPEATED NOXIOUS COLD PRESSOR TESTS MEASURED BY FUNCTIONAL NEAR INFRARED SPECTROSCOPY ON FOREHEAD

The content of this chapter is a published manuscript in the journal of Annals of Biomedical Engineering.

2.1. Abstract

The objective of this research was to assess the utility of a simple near infrared spectroscopy (NIRS) technology for objective assessment of the hemodynamic response to acute pain. For this exploration, we used functional near infrared spectroscopy (fNIRS) to measure the hemodynamic response on the forehead during three trials of a cold pressor test (CPT) in 20 adults. To measure hemodynamic changes at the superficial tissues as well as the intracranial tissues, two configurations of „far‟ and „near‟ source- detector separations were used. We identified two features that were found to be fairly consistent across all subjects. The first feature was the change of total hemoglobin (THb) concentration in a given condition divided by the duration of that condition (THb ).

Statistical analyses revealed that during the first CPT trial THb significantly changed from its baseline value in all channels. Also, adaptation to repeated CPTs was observed in both THb parameter and the reported post-stimulus pain rating scores. The second feature was the difference between the maximum and the minimum of the evoked changes in the THb concentration (ΔTHb). A significant correlation was observed between the post-stimulus pain rating score and ΔTHb at all channels. An asymmetrical 10 activity was observed only at the „far‟ channels. These results suggest that fNIRS can potentially be used as a reliable technique for the assessment of the hemodynamic response to tonic pain induced by the CPT.

Key terms: numerical rating scale, pain, sympathetic nervous system

Abbreviations: CI: Confidence Interval; CPT: Cold pressor test; CSF: Cerebrospinal fluid; FDR: False Discovery Rate; LED: Light emitting diode; fDA: functional Data

Analysis; fNIRS: functional near infrared spectroscopy; Hb: deoxy-hemoglobin; HbO2: oxy-hemoglobin; LQ: Laterality Quotient; NRS: Numerical rating scale; THb: Total hemoglobin

2.2. Introduction

Cold pressor test (CPT) is a conventional test widely used in research involving psychological, cardiovascular, and neurological disorders. It was first used by Hines and

Brown in 1932 to experimentally raise blood pressure for the study of hypertension (12).

The application of CPT for inducing experimental pain in healthy adults was initially introduced by Wolf and Hardy in 1941 (27). Since then, a large number of research studies have employed CPT for two main purposes: to evoke generalized sympathetic activation and to induce tonic pain.

The effect of sympathetic activation evoked by CPT on the cerebral hemodynamics has been investigated by using transcranial laser Doppler sonography for the assessment of the cerebral blood flow velocity in the middle cerebral arteries (14, 15). Sympathetically mediated changes in capillary blood flow and skin microcirculation during local cooling were studied using laser Doppler flowmetry (16). Laser Doppler utilizes low power lasers 11 that can interrogate the outermost 0.5 to 1 mm depth of the skin. Single point laser

Doppler systems deliver good temporal but poor spatial resolution and the measured change in blood flow greatly depends on the position of the probe relative to the location of the affected arterioles and venules (17). Another shortcoming of single point laser

Doppler is its sensitivity to motion artifacts (17, 18). Recently, laser Doppler imaging has been introduced which offers much greater spatial resolution compared to single point laser Doppler flowmetry. However, its temporal resolution is limited and the cost of the laser Doppler imaging equipment limits its routine application in clinical settings (28).

The cortical processing of the tonic pain induced by CPT has been investigated in several neuroimaging studies. Some brain regions have been identified to be involved in the processing of noxious cold stimuli. Di Piero et al. used a Xenon-133 inhalation single- photon emission tomography (SPET) to assess the cerebral blood flow in response to

CPT performed on the left hand (29). They observed increased activation in regional blood flow in the contralateral frontal lobe and bilateral temporal regions as well as in the contralateral primary sensorimotor cortex in the cortical region representing the hand. In a positron emission tomography (PET) neuroimaging study, Casey et al. found increased regional cerebral blood flow (rCBF) in response to CPT in the lateral prefrontal, anterior cingulate and insular/precentral opercular cortices ipsilaterally and in the sensorimotor cortex contralaterally (30).

Despite advances in imaging technology that significantly help basic science, there remains an unmet clinical need for a practical, inexpensive tool for the reliable and objective assessment of human response to pain. Recently, a few research studies suggested the use of functional near-infrared spectroscopy (fNIRS) for monitoring 12 cortical activation in response to noxious stimuli in new-born infants (31), healthy adults

(22, 32), patients undergoing cardiac surgery (33), and individuals suffering from migraines (34). fNIRS is an emerging technology which enables real time measurement of tissue oxygenation and hemodynamics noninvasively (19). fNIRS can be portable and has low equipment and maintenance costs. It is relatively robust to motion artifacts and therefore, no movement restriction is required during measurement, unlike the constraints imposed by other functional imaging techniques. Using the commonly used source- detector (S-D) separation distance of 2.5 to 4 cm, fNIRS can measure changes in blood oxygenation parameters within different layers of the head, i.e. the scalp and grey matter up to a depth of 1.25 to 2 cm, respectively (26).

The main goal of the present research was to explore the potential of fNIRS for objective assessment of pain. Currently, there is no practical method available for an objective assessment of pain and clinicians are basically relying on the subjective self-report measures using limited scales. Beside the subjective nature of pain scales, their applicability for explaining different types and origins of pain is questionable. We aimed to investigate whether the hemodynamic parameters measured by fNIRS can be used as a biomarker of the tonic pain induced by CPT in healthy adults. We used a CPT as an experimental model of tonic pain because it is a conventional test that is easy to implement in experimental and clinical settings. Furthermore, healthy adults are typically familiar with the induced stimulus and the evoked response to the CPT in humans is well documented (35-37). 13

2.3. Materials and Methods

2.3.1 fNIRS Principles and Instrumentation

The fNIRS exploits the fact that in the near infrared range (700 nm – 900 nm), water, the main ingredient of tissues in vivo, has the lowest light absorption, whereas deoxy- hemoglobin (Hb) and oxy-hemoglobin (HbO2) chromophores are the main absorbers with distinctive absorption characteristics (38-40). By choosing two wavelengths in the near infrared spectrum and measuring the attenuation change at two different time points, the relative change in the concentration of Hb and HbO2 molecules can be calculated using the modified Beer-Lambert law (41):

Ib Hb Hb HbO2 HbO2 ΔOD log ( ) . Δc . d . DPF + . Δc . d . DPF It where, ΔOD is termed optical density and is the change in optical intensity for the wavelength , Ib is the light intensity measured during baseline, It is the light intensity

Hb HbO2 detected during or after a given task, and are the absorption coefficients of Hb and HbO2 molecules at the wavelength , ΔcHb and ΔcHbO2 are the concentration changes of Hb and HbO2 molecules due to the task, d is the physical distance between the light source and the photodetector, and DPF is the differential pathlength factor adjusted for the increased pathlength between the light source and the photodetector due to scattering at the wavelength . When measured at two wavelengths 1 and 2, this equation can be solved for the change in the concentration of Hb and HbO2 molecules.

In this study, fNIRS data were collected using the continuous wave fNIRS system first described by Chance et al. (38) and further developed in our laboratory at Drexel

University. The fNIRS system is composed of three subsystems: 1) fNIRS sensors that 14 consist of one light source and three photodetectors. The light source is a multi- wavelength light emitting diode (LED) manufactured by Epitex Inc. type L4*730/4*850 -

40Q96-I. The LED comes in a STEM TO-5 package at 730 nm and 850 nm wavelengths with an output power of 5 to 15 mW. The photodetectors are manufactured by Burr-

Brown Corporation type OPT101 and come in an 8-pin DIP package. 2) A control box for operating the LEDs and photodetectors. 3) A computer running the COBI Studio software (42) developed in our laboratory for data acquisition and real-time data visualization. The fNIRS system was calibrated in our laboratory using solid and liquid phantoms with known optical absorption and scattering parameters. The detailed specification of Drexel‟s fNIRS system including safety assessment and signal to noise estimation are described elsewhere (43-45).

The fNIRS probes used in this study utilized two configurations of S-D separation in order to test the specificity of the hemodynamic response to a CPT. Using a multi- distance probe, while the „far‟ detectors sampled a superimposed hemodynamic change over a larger banana shape pathway reaching deeper layers within the head, the „near‟ detector monitored the absorption changes in a shorter pathway through superficial layers, including the skin (Figure 1). There have been several theoretical and experimental studies to detect depth-dependent changes in absorption using different S-D separations (26, 46, 47).

The choice of the S-D distance in our research was made based on previous phantom experiments in our laboratory and Monte Carlo simulations by other groups. Okada et al.

(26) reported that for an S-D separation of 15 mm and less, the mean optical path length at the deep layers is small and thus, the tissue volume being interrogated is confined to 15 the surface layer. They also described that for an S-D spacing of 30 mm, the near infrared light penetrates into the grey matter. In our approach in using two S-D separations to investigate the hemodynamic response at multiple layers, two detectors were placed at

2.8 cm distance from the light source (far channels), and one detector was located at 1 cm from the light source (near channel). This selection leads to a nominal penetration depth of up to 0.5 cm at the „near channel‟ and up to 1.4 cm at the „far channels‟ for measuring the hemodynamic changes within superficial extracranial tissues and deep intracranial layers, respectively (48).

Figure 1. A schematic of the fNIRS probe configuration. Photons travel from a light source to a photodetector through a banana shape pathway with a penetration depth of half the source-detector distance. Measures are approximate.

2.3.2 Subjects

Twenty healthy, right-handed, as judged by the Edinburgh Handedness Inventory (49)

(Laterality Quotient (LQ): 83.03 ± 22.03), individuals (10 females) with no history of neurological, psychological, or psychiatric disorders who were analgesic-free were 16 recruited from the Drexel University community. All participants signed the informed consent form approved by the Institutional Review Board (IRB) at Drexel University. At least one day prior to any experimental session, subjects were invited to participate in an orientation session in which they would perform a CPT in the real experimental setting to demonstrate the characteristics of the stimulus and to minimize anxiety associated with the initial exposure to cold water. Subjects were instructed to refrain from smoking and drinking any caffeinated or alcoholic beverages for at least 3 hours before the experiment.

2.3.3 Protocol

Two fNIRS sensors of the same configuration as previously described were positioned symmetrically on the left and right sides of a subject‟s forehead proximate to the anterior median line (Figure 2) and were secured using a medical band aid and a Velcro strap.

Raw optical intensity measurements were collected at a sampling frequency of 2Hz in a dimly lit room with an ambient temperature of ~23oC. The effect of background light on fNIRS data was negligible and in fact, since the fNIRS measures the relative changes in

Hb and HbO2 concentrations with respect to a baseline condition, this effect was washed out in the calculations. Subjects were seated comfortably in an armchair, facing away from the experimenter to minimize any distraction.

Each experiment consisted of a baseline recording at rest followed by the immersion of the right hand up to the wrist into a bucket of circulating tepid water kept at room temperature (~23oC) for 2 minutes for adaptation. Then, subjects performed three serial trials of a CPT in the ice water (~0oC), each trial lasting 45 seconds followed by 2 minutes post-stimulus hand immersion in the tepid water for the hemodynamic recovery.

A block diagram of the protocol is shown in Figure 3. 17

Figure 2. A demonstration of the placement of fNIRS probes on a subject’s forehead. The probes are shown reversed to illustrate the location of the light source and photodetectors.

Figure 3. Block diagram of the protocol

Both water containers were equipped with commercial aquarium pumps for water circulation to minimize heat buildup around the immersed hand (50). The cold water container had a separate compartment for ice cubes in order to prevent any direct contact of the subject‟s skin with ice. Subjects received an auditory command from the experimenter when to switch their hand from the tepid water into the cold water and vice versa. At the end of each CPT and upon immersion of the hand back into the tepid water, subjects were requested to report the maximum intensity of the pain experienced during 18 the CPT on a numerical rating scale from 0 to 10 (NRS-11), where „0‟ indicates no pain and „10‟ indicates the worst imaginable pain (51).

2.3.4 fNIR Data Processing

To eliminate high frequency noise, respiration and heart pulsation artifacts, raw intensity measurements were first filtered by a finite impulse response low pass filter with a cut-off frequency set to 0.14Hz. The cut-off frequency of 0.14Hz was determined based on our previous NIRS studies (52).

Changes in the concentrations of Hb and HbO2 were calculated relative to the mean value of the optical intensity during the first 15 second of the pre-stimulus baseline recording. The Hb and HbO2 data were then smoothed using a bspline basis expansion by imposing a penalty on the roughness of the second derivative of the data with a lambda of 300 (53). Total hemoglobin (THb) concentration was obtained from:

THb = Hb + HbO2.

All signal processing calculations were performed in MATLAB (R2011a,

MathWorks, Natwick, MA) and the smoothing was performed using the functional data analysis (fDA) package for MATLAB (53). Statistical analyses were conducted using the

IBM SPSS Statistics 19. The significance criterion was  < 0.05 for all analyses. In the case of departure from sphericity, the degree of freedom associated with the corresponding F-ratio was corrected using the Greenhouse-Geisser (G-G) correction value. For subjective pain ratings, a non-parametric Friedman‟s ANOVA by ranks with

Wilcoxon Signed Ranks post hoc analyses were calculated. To control the type I error introduced by simultaneous testing of the experimental-wise error rate, we applied the 19

Benjamini and Hochberg False Discovery Rate (FDR) procedure (54) to the omnibus

ANOVAs and the post hoc multiple comparison tests.

2.4. Results

The average hemodynamic response across 20 subjects is shown in Figure 4a-b. Here, we report the results of the analysis of THb concentration calculated as the sum of Hb and

HbO2 concentrations.

The analyses of the THb concentration were performed to assess the: 1) effect of a noxious cold water stimulus on the hemodynamic response measured by fNIRS; 2) adaptation to repeated CPTs observable in THb concentrations as well as in subjective pain scores; 3) correlation between THb concentrations and subjective pain scores within subjects; 4) specificity of the hemodynamic response to the CPTs at the superficial layers versus deep tissues; and finally 5) laterality of the hemodynamic activity measured on both sides of the forehead during the CPTs. In the following sections, we have assessed the measures and reported on our findings. 20

Figure 4A-B. The hemodynamic response to three trials of cold pressor test (CPT) averaged across 20 subjects on the right (a) and left (b) sides of the forehead. The black vertical lines represent time points at which subjects switched their hand from the tepid water to the ice water and vice versa. The pink shaded field represents the baseline period, the orange shaded fields represent the immersion in tepid water condition, and the blue shaded fields represent the CPT conditions. Hb, HbO2 and THb are abbreviations for deoxy-hemoglobin, oxy-hemoglobin and total hemoglobin, respectively. ‘S-D’ symbol stands for ‘source-detector’.

21

2.4.1. The effect of a CPT on the THb concentration

To assess the effect of a noxious cold water stimulus on the hemodynamic response, we compared four conditions: the initial baseline recording condition (30 seconds at rest), the pre-stimulus condition (the initial 2 minutes hand immersion in the tepid water), the first stimulus condition (the first CPT trial for 45 seconds) and the first post-stimulus condition (the 2 minutes hand immersion in the tepid water following the first CPT trial)

(Figure 5). We hypothesized that a noxious cold stimulus (ice water) would increase the blood flow to the head and consequently, the THb concentration would also increase.

Repeated measures ANOVAs showed that significantly changed across the four conditions in all channels (Table 1). Post hoc FDR adjusted multiple comparisons revealed that THb did not significantly change from its baseline value due to hand immersion in tepid water, except for the „near channel‟ on the right side of the forehead

(Table 2). However, hand immersion in ice water caused a significant change in THb as compared to its values at both the baseline and pre-stimulus conditions and it showed a large effect (d > 0.944). Also, THb at the pre-stimulus and stimulus conditions were significantly different from their values at the post-stimulus condition which also showed large effects (d > 1.219).

22

Figure 5. Line graphs of T b during four experimental conditions: baseline, pre-stimulus, stimulus, and post-stimulus conditions (for the definition of T b , please see section 5.1). Error bars represent the 95% confidence intervals (CI)

Table 1. Results of repeated measures ANOVA on T b during the four experimental conditions: baseline, pre-stimulus, stimulus, and post-stimulus conditions.

T b (µmoles/second) Forehead (mean ± SD+) Repeated Measures Channel * Side Baseline Pre-stimulus Stimulus Post- ANOVA stimulus 1 F(1.12,21.29) = 37.16 G-G++ 0.00 ± 0.02 -0.00 ± 0.01 0.12 ± 0.08 -0.05 ± 0.03 (2.8cm) p < 0.001 2 F(1.08,20.49) = 33.71 G-G 0.01 ± 0.02 -0.00 ± 0.01 0.13 ± 0.10 -0.05 ± 0.04 Right (2.8cm) p < 0.001 3 F(1.10,20.01) = 29.54 G-G 0.03 ± 0.03 0.02 ± 0.02 0.13 ± 0.09 -0.03 ± 0.04 (1cm) p < 0.001 1 F(1.14,21.65) = 35.16 G-G 0.00 ± 0.02 0.00 ± 0.01 0.08 ± 0.06 -0.03 ± 0.02 (2.8cm) p < 0.001 2 F(1.13,21.38) = 25.50 G-G Left 0.01 ± 0.02 -0.00 ± 0.01 0.06 ± 0.05 -0.02 ± 0.02 (2.8cm) p < 0.001 3 F(1.29,24.56) = 25.66 G-G 0.02 ± 0.04 0.02 ± 0.02 0.10 ± 0.08 -0.02 ± 0.03 (1cm) p < 0.001

*Experiment-wise error rates are False Detection Rate (FDR) adjusted (p-values). +„SD‟ stands for „standard deviation‟. ++Due to departure from sphericity, the degrees of freedom were adjusted using the Greenhouse-Geisser (G- G) correction value. 23

Table 2. Pairwise condition comparisons across the Forehead Side and Channel including 95% confidence intervals and effect sizes for T b , comparing the four conditions: the baseline, 1st pre- stimulus, 1st stimulus, and 1st post-stimulus conditions.

95% CI+ of the Forehead difference p- Cohen’s Channel Pairs of Conditions * Side Lower Upper value d Limit Limit Baseline –Pre-stimulus -0.001 0.020 0.075 ----- Baseline – Stimulus -0.150 -0.067 < 0.001 1.222 1 Baseline – Post-stimulus 0.038 0.070 < 0.001 1.563 (2.8cm) Pre-stimulus – Stimulus -0.158 -0.079 < 0.001 1.414 Pre-stimulus – Post-stimulus 0.028 0.060 < 0.001 1.276 Stimulus – Post-stimulus 0.109 0.216 < 0.001 1.429 Baseline –Pre-stimulus -0.002 0.017 0.110 ----- Baseline – Stimulus -0.178 -0.075 < 0.001 1.149 2 Baseline – Post-stimulus 0.043 0.079 < 0.001 1.584 Right (2.8cm) Pre-stimulus – Stimulus -0.182 -0.086 < 0.001 1.318 Pre-stimulus – Post-stimulus 0.035 0.072 < 0.001 1.359 Stimulus – Post-stimulus 0.123 0.252 < 0.001 1.359 Baseline –Pre-stimulus 0.008 0.026 0.001 0.871 Baseline – Stimulus -0.140 -0.053 < 0.001 1.031 3 Baseline – Post-stimulus 0.045 0.084 < 0.001 1.555 (1cm) Pre-stimulus – Stimulus -0.153 -0.068 < 0.001 1.184 Pre-stimulus – Post-stimulus 0.031 0.064 < 0.001 1.351 Stimulus – Post-stimulus 0.102 0.220 < 0.001 1.280 Baseline –Pre-stimulus -0.003 0.013 0.229 ----- Baseline – Stimulus -0.101 -0.044 < 0.001 1.183 1 Baseline – Post-stimulus 0.026 0.048 < 0.001 1.568 (2.8cm) Pre-stimulus – Stimulus -0.105 -0.049 < 0.001 1.279 Pre-stimulus – Post-stimulus 0.023 0.042 < 0.001 1.570 Stimulus – Post-stimulus 0.073 0.145 < 0.001 1.422 Baseline –Pre-stimulus -0.002 0.014 0.129 ----- Baseline – Stimulus -0.088 -0.030 < 0.001 0.944 2 Baseline – Post-stimulus 0.021 0.038 < 0.001 1.687 Left (2.8cm) Pre-stimulus – Stimulus -0.091 -0.038 < 0.001 1.136 Pre-stimulus – Post-stimulus 0.015 0.033 < 0.001 1.219 Stimulus – Post-stimulus 0.055 0.122 < 0.001 1.225 Baseline –Pre-stimulus -0.010 0.016 0.640 ----- Baseline – Stimulus -0.107 -0.037 < 0.001 0.967 3 Baseline – Post-stimulus 0.027 0.060 < 0.001 1.236 (1cm) Pre-stimulus – Stimulus -0.110 -0.040 < 0.001 0.999 Pre-stimulus – Post-stimulus 0.030 0.052 < 0.001 1.759 Stimulus – Post-stimulus 0.074 0.157 < 0.001 1.298

*Significant results (False Detection Rate (FDR) adjusted) are shown in bold. +„CI‟ stands for „confidence interval‟. 24

2.4.2. Adaptation to repeated CPTs observable in subjective pain ratings and THb

The adaptation to repeated trials of CPT observable in both subjective pain ratings

(Figure 6) and THb at the stimulus condition (Figure 7) was further explored.

Post-stimulus pain rating scores reported based on the NRS-11 (in which „0‟ means no pain and „10‟ means the worst pain imaginable) was tested with Friedman‟s ANOVA by

Ranks. There was a significant difference in the median value of the subjects‟ pain rating scores with respect to CPT trials (2 (2) = 14.70, p = 0.001). Post hoc Wilcoxon signed ranks tests yielded that the intensity of the reported pain rating scores decreased significantly across CPT trials. It was revealed that the pain score reported after the first trial of CPT was significantly higher than the pain scores given after the second and third trials (Z = 2.29, p = 0.02 and Z = 2.92, p = 0.004, respectively). Also, the pain score reported after the second trial of CPT was significantly higher than the pain score given after the third trial (Z = 2.59, p = 0.01).

25

Figure 6. Boxplots of post-stimulus pain rating scores of 20 subjects across three trials of cold pressor test (CPT). Two outliers were identified for pain scores reported after 3rd trial of CPT and are represented as circles.

Figure 7. Line graphs of T b across cold pressor test (CPT) trials (for the definition of T b , please see section 5.1)

26

Figure 8. Line graphs of T b for separate trials of cold pressor test (CPT) (for the definition of T b , please see section 5.1). Significant interactions between the Forehead Side and Channel were found for the 1st and 2nd CPT trials.

Table 3. Main effect of cold pressor test (CPT) trials repeated measures ANOVAs on T b

T b (µmoles/second) Forehead (mean ± SD+) Repeated Measures Channel * Side 1st CPT 2nd CPT 3rd CPT ANOVA

F(1.38,26.20) = 31.41 G- 1 (2.8cm) 0.12 ± 0.08 0.05 ± 0.04 0.04 ± 0.05 G++ p < 0.001 F(1.32,25.08) = 27.27 G-G Right 2 (2.8cm) 0.13 ± 0.10 0.05 ± 0.03 0.04 ± 0.05 p < 0.001 F(1.49,28,35) = 16.83 G-G 3 (1cm) 0.13 ± 0.09 0.06 ± 0.04 0.06 ± 0.06 p < 0.001 F(1.32,25.03) = 24.62 G-G 1 (2.8cm) 0.08 ± 0.06 0.03 ± 0.03 0.02 ± 0.02 p < 0.001 F(1.44,27.30) = 23.18 G-G Left 2 (2.8cm) 0.06 ± 0.05 0.02 ± 0.03 0.02 ± 0.03 p < 0.001 F(1.55,29.36) = 32.20 G-G 3 (1cm) 0.10 ± 0.08 0.05 ± 0.06 0.03 ± 0.06 p < 0.001

*Experiment-wise error rates are False Detection Rate (FDR) adjusted (p-values). +„SD‟ stands for „standard deviation‟. ++Due to departure from sphericity, the degrees of freedom were adjusted using the Greenhouse-Geisser (G- G) correction value. 27

Table 4. Pairwise cold pressor test (CPT) trials comparisons across the Forehead Side and Channel including 95% confidence intervals and effect sizes for T b

95% CI+ of the Forehead difference Cohen’s Channel Pair of CPT p-value* Side Lower Upper d Limit Limit 1st CPT – 2nd CPT 0.043 0.095 < 0.001 1.266 1 1st CPT – 3rd CPT 0.052 0.106 < 0.001 1.376 (2.8cm) 2nd CPT – 3rd CPT -0.003 0.023 0.138 ----- 1st CPT – 2nd CPT 0.051 0.125 < 0.001 1.117 2 Right 1st CPT – 3rd CPT 0.061 0.123 < 0.001 1.391 (2.8cm) 2nd CPT – 3rd CPT -0.013 0.021 0.637 ----- 1st CPT – 2nd CPT 0.033 0.107 0.001 0.886 3 1st CPT – 3rd CPT 0.045 0.103 < 0.001 1.184 (1cm) 2nd CPT – 3rd CPT -0.018 0.025 0.726 ----- 1st CPT – 2nd CPT 0.027 0.076 < 0.001 0.964 1 1st CPT – 3rd CPT 0.041 0.082 < 0.001 1.389 (2.8cm) 2nd CPT – 3rd CPT -0.002 0.021 0.088 ----- 1st CPT – 2nd CPT 0.022 0.062 < 0.001 0.971 2 Left 1st CPT – 3rd CPT 0.033 0.066 < 0.001 1.392 (2.8cm) 2nd CPT – 3rd CPT -0.004 0.18 0.174 ----- 1st CPT – 2nd CPT 0.029 0.070 < 0.001 1.131 3 1st CPT – 3rd CPT 0.045 0.083 < 0.001 1.588 (1cm) 2nd CPT – 3rd CPT 0.003 0.027 0.020 0.573

*Significant results (after controlling the False Detection Rate (FDR)) are shown in bold.

+„CI‟ stands for „confidence interval‟.

To assess the hemodynamic response across three CPT trials, repeated measures

ANOVAs were performed on THb for all channels separately on the left and right sides of the forehead (Table 3). There were significant main effects of CPT trials in THb in all channels. FDR adjusted post hoc multiple comparisons showed that THb at the first CPT trial was significantly different from the THb at the second and third CPT trials in all channels on both sides of the forehead with large effects (d > 0.886) (Table 4). However, except for the „near channel‟ on the left side, THb values at the second and third CPT 28 trials were not significantly different. This significant difference represented a moderate effect (d = 0.573).

We experimentally identified another feature which may be a better representation of the individualized hemodynamic response to CPTs. This variable, which is denoted as ΔTHb, is expected to account for the inter-subject variability in the latency of the hemodynamic response and is defined as the change in the THb concentration induced by each CPT.

ΔTHb is calculated as the difference between the minimum value of THb that is obtained within 20 seconds after hand immersion in ice water and the maximum value of THb within a 35 seconds window starting 15 seconds before removing the hand from the ice water and ending 20 seconds after it (Figure 9A-C). The minimum value of THb was searched within a region when the effect of CPT is expected to be minimal. The maximum value of THb was searched within a window in which the cumulative effect of

CPT was expected to reach its maximum with enough time for the hemodynamics to fully evolve. Although we acknowledge that within the selected time windows subjects may be cognitively involved in anticipating the next stimulus or reporting their pain, the effect of these cognitive tasks on the hemodynamics is considered to be much lower than the observed effect of a cold water stimulus and thus, it can be disregarded.

29

(A)

(B)

(C)

Figure 9. Three sample total hemodynamic responses to the first cold pressor test (CPT) measured at the ‘far channel 1’ on the right side of the forehead. The first black vertical line represents the time point at which subjects immersed their hand into the ice water and the second black vertical line shows the time when subjects immersed their hand back into the tepid water. Dashed blue vertical lines represent the detected time points for ΔT b calculation (see 5.3 section). 30

2.4.3. Correlation analysis between ΔTHb and subjective pain scores

A secondary interest was to determine whether a decrease in the reported pain scores across the three CPT trials within a subject was associated with a decrease in the measured hemodynamic parameter (i.e. ΔTHb). To answer this question, within-subjects correlation coefficients (55) were calculated. All the correlation values were significant at the 0.05 level and showed a moderate to large effects (ranging from 0.50 to 0.67), corresponding to a linear relation between the subjective pain rating scores and ΔTHb within subjects (Table 5).

Table 5. Within-subjects correlation values between self-reported pain rating scores and ΔTHb

Forehead Side Channel Within-subjects correlation* Right Far Channel 1 0.609 Far Channel 2 0.577 Near Channel 0.541 Left Far Channel 1 0.556 Far Channel 2 0.495 Near Channel 0.670

*All the correlation coefficients were significant at a 0.05 level.

2.4.4. Comparing ΔTHb response at the ‘far channels’ with the ‘near channels’ across

left and right sides of the forehead

Separate 2 × 3 (Forehead Side by Channel) repeated measures ANOVAs on ΔTHb were performed for each CPT trial to assess the laterality and specificity of the hemodynamic activity. Specificity of the hemodynamic response was measured by „far channels 1 and 31

2‟ representing the intracranial layers and a „near channel‟ representing the superficial tissues. Greenhouse-Geisser (G-G) corrections were applied for violations of the sphericity assumption on the repeated measures tests.

Results of the 2 × 3 (Forehead Side by Channel) repeated measures ANOVAs on ΔTHb showed significant interactions between the Forehead Side and Channel for every three

CPT trials (F(2.26,42.95) = 5.78 G-G, p = 0.004, F(3.05,58.03) = 5.46 G-G, p = 0.002,

F(2.18,41.33) = 7.37 G-G, p = 0.001; for the first, second and third CPT trials, respectively). There was a significant main effect of the Forehead Side for all CPT trials

(F(1,19) = 8.69, p = 0.008, F(1,19) = 8.40, p = 0.009, F(1,19) = 8.04, p = 0.01; for the first, second and third CPT trials, respectively) and a significant main effect of the

Channel for the second and third CPT trials (F(2,38) = 4.70 G-G, p = 0.02, F(2,38) =

10.75, p < 0.001) (Figure 10).

Post hoc multiple comparisons after controlling the FDR showed that there is no significant difference between the „near channel‟ and „far channels‟ on either side of the forehead.

Significant asymmetrical activity was revealed by the FDR adjusted post hoc tests on the

ΔTHb response measured by the „far channels‟, indicating moderate-to-large effects

(0.697 < d < 0.902). However, there was no laterality in the ΔTHb for the superficial tissues as measured by the „near channels‟ (Table 6).

32

Figure 10. Line graphs of ΔT b for separate trials of cold pressor test (CPT) (for the definition of ΔT b, please see Figures 9A-C). For all three trials of CPT, significantly different activation on the left and right sides of the forehead was observed in ΔT b as measured by the ‘far channels’, while no laterality was observed in ΔT b as measured by the ‘near channels’.

Table 6. Pairwise channel pair comparisons across cold pressor test (CPT) trials including 95% confidence intervals and effect sizes for ΔTHb 95% CI+ of the difference p- Cohen’s CPT Trial Pair of Channels Lower Upper value* d Limit Limit Right Far Channel 1 – Left Far Channel 1 0.817 4.151 0.006 0.697 1st Right Far Channel 2 – Left Far Channel 2 1.513 5.225 0.001 0.849 Right Near Channel – Left Near Channel -0.793 3.652 0.194 ------Right Far Channel 1 – Left Far Channel 1 0.595 2.025 0.001 0.857 2nd Right Far Channel 2 – Left Far Channel 2 0.619 2.483 0.002 0.779 Right Near Channel – Left Near Channel -0.866 1.818 0.467 ------Right Far Channel 1 – Left Far Channel 1 0.614 2.214 0.002 0.827 3rd Right Far Channel 2 – Left Far Channel 2 0.713 2.251 0.001 0.902 Right Near Channel – Left Near Channel -0.642 2.444 0.237 ------

*Significant results (after controlling the False Detection Rate (FDR)) are shown in bold. + ‘CI‟ stands for „confidence interval‟. 33

2.5. Discussion

Despite the technological advancement in medical device development, effective assessment and management of pain is poorly addressed. One possible reason could be the subjectivity of the pain experience. It is well known that the pain experience including its perception and expression is highly individualized. Therefore, there is a need for a more objective assessment of pain.

In the conventional pain practice, vital signs such as heart rate, blood pressure, respiratory rate, galvanic skin response and/or cutaneous blood flow measured by laser

Doppler flowmetry have been used to monitor the physiological parameters in response to a noxious stimulus. However, the clinical experience has proven that these physiological signs are not practically reliable and should be interpreted cautiously. For instance, physiological parameters change due to many factors other than pain such as distress, medications, and illness. Laser Doppler flowmetry data are also very noisy and the smallest motion causing any change in the contact between the probe tip and the skin creates a drastic change in the signal. Further, since the typical depth of penetration for a laser Doppler flowmeter is less than 1mm, the results are very sensitive to the type of the skin and the location where the probing is conducted (17, 18).

Over the past five years, we have used fNIRS for monitoring hemodynamic response to noxious stimuli. We aimed to find an association between a noxious stimulus and the evoked hemodynamic response and to demonstrate that induced noxious stimuli elicit a reproducible and consistent hemodynamic response which can be reliably detected by fNIRS. During this period, we have explored different protocols by changing the setting of the stimulus such as type, intensity and duration as well as fNIRS probe configurations 34 in an effort to optimize the reproducibility and strength of the response. The results presented in this manuscript demonstrate only a piece of our quest to identify reliable and robust stimulus-response parameters. The present study aimed to investigate the hemodynamic changes during repeated CPTs employed as an experimental model of the tonic pain with the capability of evoking a general acute sympathetic activation. We established that a simple NIRS device can effectively monitor the hemodynamic response to a CPT. Since the signal penetrates deeper and the light source and detector are directly in contact with the skin, there is a negligible noise due to the movement of components with respect to the skin.

According to the literature, cold sensation, at freezing temperatures, becomes painful within the initial 10 seconds and progressively increases until it reaches its maximum at approximately 60 seconds (37). Based on our pilot experiments, we observed that 45 seconds was a reasonable duration of a CPT to induce an acute pain response in normal subjects and also to allow a detectable hemodynamic response to evolve. Moreover, the 2 minutes post-stimulus immersion in the tepid water was sufficient time for the hemodynamic response to get close to a steady state. In retrospect, we realized that we could have allowed more time between CPTs to let the hemodynamics return to its initial baseline level. However, we tried to keep the duration of the experiment as brief as possible to avoid our subjects becoming bored and other potential confounds.

It is suggested that for a stimulus that lasts for 1 second, the hemodynamic response evolves over 10 to 12 seconds with some components having longer recovery time (56).

For longer events, the hemodynamic response adds up roughly linearly over time (57).

The selected search windows for calculating the ΔTHb were determined experimentally 35 using our small sample size. We noticed that some subjects reached their maximum THb value as soon as 30 seconds after starting a CPT. Consequently, we chose to extend the search window for the maximum THb response to 15 seconds before the termination of the CPT and 20 seconds after it in all trials.

The tonic pain induced by noxious cold water is massively confounded and regulated by sympathetic activity. Thus, it is plausible to conclude that the observed hemodynamic change in response to CPT is mainly dominated by the sympathetic nervous system in terms of global increase in cerebral blood flow as a consequence of the increase in cardiac output following painful stimuli. The observed asymmetric change in ΔTHb at the intracranial tissues as measured by the „far channels‟ and no laterality in the superficial tissues as measured by the „near channels‟ could be indicators of the presence of simultaneous activations of the systemic sympathetic system as well as higher cerebral regulatory systems. While the results of this study are consistent with the previously reported cerebral asymmetry in regulations of the autonomic nervous system (58), further investigations are required prior to making any conclusion on the hemisphere specificity in response to a CPT.

Our experimental results did not provide any strong evidence to link the observed lateralized activation in the prefrontal region to any specific pain-induced cortical activity; partly due to the type of the stimulus applied and limited data analyses techniques used. However, a near infrared spectroscopy study suggests that a stress- inducing mental task induces asymmetrical activities in the prefrontal cortex (58). Thus, the lateralized activation in the prefrontal region could be related in part to the stress induced by a CPT or may reflect a general arousal due to a cold water stimulus. 36

Repeated CPTs have been previously used to investigate cardiovascular habituation and to activate endogenous analgesic systems (59, 60). In the present study, three trials of

CPT were performed to test the reliability of our measurements as well as to study the adaptation in our sample regarding subjective pain scores and the objective hemodynamic measurements. We observed that THb responses had significantly different values in the second and third CPT trials as compared to the first CPT trial in all channels with large effect sizes (d > 0.886). However, no significant difference was found between the

THb in the second and third CPT trials, except for one channel (the „near channel‟ on the left side of the forehead). This observation suggests that the hemodynamic activity adapts to repeated trials of CPT. Furthermore, the median of pain rating scores, which is a subjective measure, also decreased across the CPT trials as a manifestation of the subjects‟ adaptation to the pain. The correlation analysis showed a moderate to large effect size corresponding to a strong association between the ΔTHb and pain rating scores within subjects. However, a future study that benefits from a significantly larger sample size needs to be conducted to further validate this association. Since self-reporting is very subjective in nature, an absolute correlation between fNIRS parameters and self-reported pain scores is not expected. We suggest that a combination of our measurement and patients‟ self-report provides better information to the clinicians. In particular, our method is most useful when in various conditions subjects cannot provide reliable self- report, such as the elderly with dementia and impaired cognition, very young children, or critically ill patients (4). We hope that our objective measurement provides complimentary information to the self-report for the clinicians. 37

Undoubtedly, there is a critical need for a reproducible, brief and simple measurement that can correlate subjective pain experience to objective and quantitative parameters for both clinical and research purposes. Reliable biomarkers of pain can help in evidence- based, personalized management of pain. In the present research, we showed that the evoked hemodynamic response to noxious cold water stimuli can be measured using a non-invasive, portable device. Further refinement of the utilized method in this study may enable the measurement of the hemodynamic response at the cortex and the extracranial tissues. These refinements include incorporating advanced signal processing techniques and using other noxious stimuli that would evoke a less generalized systemic response such as hot plates and pressure algometry. Then, fNIRS can be used as a powerful clinical technique to assist clinicians in the assessment of pain with or without subjective self-reports.

2.6. Acknowledgement

We would like to thank Mr. Frank Kepic for his invaluable technical support and advice.

Authors also acknowledge Mr. Troy Carlson for designing and constructing the system for cold pressor tests.

38

CHAPTER 3: FUNCTIONAL DATA ANALYSIS VIEW OF FNIRS DATA

The content of this chapter is a submitted manuscript to the Journal of Biomedical Optics.

3.1. Abstract

Functional near infrared spectroscopy (fNIRS) is a powerful tool for the study of oxygenation and hemodynamics of living tissues. Despite continuous nature of the processes generating the data, analysis of fNIRS data have been limited to discrete-time methods. We propose a novel technique, namely functional data analysis (fDA), that converts discrete samples to continuous curves. We used fNIRS data collected on forehead during a cold pressor test (CPT) from 20 healthy subjects. Using functional principal component analysis, oxyhemoglobin (HbO2) and deoxyhemoglobin (Hb) curves were decomposed into several components based on variability across the subjects. Each component corresponded to an experimental condition and provided qualitative and quantitative information of the shape and weight of that component.

Furthermore, we applied functional canonical correlation analysis to investigate the interaction between Hb and HbO2 curves. We showed that variation of Hb and HbO2 was positively correlated during the CPT with a „far‟ channel on right forehead showing a smaller and faster HbO2 variation than Hb. This research suggests the fDA platform for the analysis of fNIRS data which solves problem of high dimensionality, enables study of response dynamics, enhances characterization of the evoked response, and may improve design of future fNIRS experiments. 39

Keywords: functional canonical correlation, cold pressor test, functional data analysis, hemodynamics, multi-distance probe, near-infrared spectroscopy, pain, functional principal component analysis.

3.2. Introduction

Functional near infrared spectroscopy (fNIRS) (61-63) is a promising imaging technology for noninvasive, continuous monitoring of regional blood flow and tissue oxygen consumption. It utilizes low intensity light at the near infrared spectrum (between

600 nm to 900 nm), within which the optical absorbance of living tissues is small and hence, light can penetrate up to a few centimeters into the tissue. The two forms of the oxygen-carrying molecule of the blood - denoted as oxyhemoglobin (HbO2) (when oxygen is bound) and deoxyhemoglobin (Hb) (when oxygen is delivered) - have distinct spectroscopic characteristic within the near infrared (NIR) spectrum. Changes in optical absorption parameters are measured at two different wavelengths within the NIR spectrum and then, are converted to changes in Hb and HbO2 concentrations using the

Beer-Lambert law.

The depth of NIR light penetration is proportional to the distance between the light source and the photodetector (64). With multiple source-detector separations (S-D), fNIRS enables concurrent monitoring of oxygenation changes at superficial and deep tissues across different body parts, such as hands, feet, and head. In a head fNIRS measurement with a typical S-D of 3 cm, NIR light penetrates sufficiently to reach the uttermost layer of the brain - the cortex. fNIRS is increasingly being used for detecting task-related oxygen consumption changes in the cortex. Mainly because of the portability 40 and low cost of the measurements, fNIRS has become considerably popular for research in neuroscience, sports medicine, psychology, psychiatry, and rehabilitation. Over the past two decades, it has been applied to several clinical conditions too, such as schizophrenia (65-72), Alzheimer‟s disease (73, 74), epilepsy (75-78), and neonatal intensive care (79-87).

The measured hemodynamic response by fNIRS is a realization of a naturally continuous physiological phenomenon. Typical fNIRS data consists of a set of discrete time points sampled every fraction of a second. In the presence of observational noise, the measured hemodynamic time series are often not smooth and fluctuating but one can assume that the true underlying trajectory is a smooth function. Moreover, due to the slow nature of the hemodynamic response and given the typical sampling rate of an fNIRS device, it is likely that the adjacent samples are correlated to some extent. It is, therefore, desirable to view an individual‟s repeated measures over a time interval as a continuous function or curve.

Traditionally, fNIRS studies involve linear regression and analysis of variance on features extracted from the recovered hemodynamic response. The time series of the evoked hemodynamic response can be directly estimated by linear deconvolution of the measured hemodynamic changes and the experimental timing function (a boxcar function that determines the timing of the experimental paradigm). An accurate estimate of the evoked response, however, requires a precise choice of the timing function (88). Another approach for describing the waveform of the evoked hemodynamic response, which is adapted from fMRI analysis, assumes a temporal shape for the so-called „hemodynamic response function‟ (HRF) (89). Given this a priori assumption, it models the HRF using a 41 set of basis functions (such as Gamma or Gaussian functions) (90). This technique has the advantage of reducing the number of unknowns by returning a smooth function with definite parameters. However, results may be biased by the choice of the canonical basis functions. The estimated evoked hemodynamic responses are averaged to enhance signal- to-noise ratio, provided the inter stimulus intervals (ISI) are sufficiently large to avoid overlap of HRFs (88).

Analysis of fNIRS data has primarily been limited to discrete-time methods. Functional data analysis (fDA) (91) is an alternative technique developed for handling large number of data sampled over a continuum which is often time. In a functional domain, we study functional objects rather than sample points; therefore, the vector observations X1, …, Xn, where n is the number of observations, are replaced by functions x1(t),…xn(t). The philosophy behind fDA is “to think of observed data functions as single entities, rather than merely as a sequence of individual observations” (91). Discrete observations can be converted to continuous functions using a linear combination of basis functions. The advantages of using such representation is twofold: 1) it provides a computational platform that allows storing and analyzing large amount of data with improved computational efficiency and flexibility as computational challenges arise due to huge number of measurements on each of small number of subjects, known as large p small n problem (see Ref. (92)); 2) smooth functions allow study of the dynamics of the underlying processes through their derivatives. fDA has been applied to neuroimaging studies (93). Viviani et al. first proposed an fDA approach for exploratory analysis of fMRI images for a single subject case (94). They showed that compared to ordinary principal component analysis (PCA), the functional 42 version of PCA could better visualize the variability of the data introduced by experimental alternations as it takes advantage of smooth functions. Later, Long et al. followed an fDA approach for dimension reduction of fMRI data for the estimation of noise covariance kernel (95).

Our primary aim in this research was to introduce fDA methodology for exploratory analysis of fNIRS data to obtain a better insight into the physiological response. Given a set of Hb and HbO2 functions collected from a number of individuals, we wish to characterize main directions of variability and investigate shared features by the two functions. We specifically would like to explore: 1) temporal variability within observations; 2) linear relationship and temporal association between Hb and HbO2 time series. We employed two novel functional techniques: functional principal component analysis (fPCA) and functional canonical correlation analysis (fCCA) (91). fPCA explores the variability between a set of observations. In a functional , it reveals when in a time-series the maximum variation between several observations occur. fPCA works without a priori knowledge of the experimental setting or any specific assumptions on the data. In other word, fPCA provides an initial assessment and helps understand variation and population structure in high dimensional data (94). We hypothesized that fPCA can discriminate events in a set of fNIRS data during a functional task. fCCA examines the modes of variation that two sets of functions share. With this method, we would like to investigate the interaction between Hb and HbO2 time series and discover how these two functions covary over time. In other words, we want to know, 43 during the course of an experiment, how variability in Hb data is related to the variability in HbO2 data and what types of and how much variation they share.

In this study, we use Hb and HbO2 parameters measured during a painful task, namely cold pressor test (CPT). A CPT is performed by immersing a limb into a cold water

(usually freezing water) container for a specific period of time. Three trials of a CPT were used to study the effect of repeated sustained noxious stimuli on the hemodynamics and the results are published elsewhere (96).

The rest of the paper is organized as follows: in the Materials and Methods, we will briefly describe the protocol and measurements and will present a description of the fDA framework; in the Results, we will illustrate the application of fPCA and fCCA on fNIRS data collected during a CPT; lastly, in the Discussion, we will review and discuss the results.

3.3. Materials and Methods

3.3.1. Participants

Twenty healthy, right-handed individuals from the Drexel University community participated in this study after giving the informed consent form approved by the

Institutional Review Board (IRB). We instructed subjects to avoid smoking and drinking any caffeinated or alcoholic beverages for at least 3 hours prior to the experiment.

3.3.2. Protocol

Subjects performed three successive trials of a CPT. Only data from the first CPT was used for the present study. The experiment block consisted of four conditions: 1) a 30 s baseline when the subject relaxed; 2) a 2 min immersion of the right hand up to the wrist 44 into a bucket of circulating tepid water (~23oC) for adaptation; 3) a 45 s immersion of the same hand into a bucket of circulating ice water (~0oC); and 4) a 2 min post-stimulus hand immersion in the tepid water for the hemodynamic recovery.

3.3.3. Measurements

We used a continuous wave fNIRS system designed and developed at Drexel University.

The principle and instrumentation of fNIRS are described elsewhere (96). The fNIRS sensors consisted of one light source with two LEDs at 730 nm and 850 nm wavelengths and three photodetectors. Two detectors were placed at 2.8 cm from the LED making the

„far‟ channels to investigate the hemodynamic response at intracranial layers and one detector was located at 1 cm from the LED making the „near‟ channel to measure the hemodynamic changes within the superficial extracranial tissues. Two fNIRS sensors with the same configuration were positioned symmetrically on the left and right sides of a subject‟s forehead proximate to the anterior median line and were secured using a medical band aid and a Velcro strap. The sampling rate of raw optical intensity measurements was 2Hz. The optical density parameters for 730 nm and 850 nm wavelengths were calculated by taking the logarithm of the ratio of the detected light intensity during baseline to the detected light intensity during the task. The optical density time series were converted to changes in oxyhemoglobin (HbO2) and deoxyhemoglobin (Hb) concentrations using the modified Beer-Lambert law (97). 45

3.3.4. Functional Data Analysis

3.3.4.1. Building continuous, smooth curves from discrete data

Modern signal or image acquisition machines sample data points from a smooth, continuous process subject to observational noise. This can be expressed as:

yj x(tj) + ej where y is the observed data, x is the assumed noise-free latent data at time t, and e is the observational error that explains the roughness of the observed data.

Prior to any functional analysis, the discrete data need to be converted to a continuous functional object. Then, a flexible method that gives control over the level of smoothing can be applied. Spline smoothing is the most popular and powerful approach with well- defined properties (98-100). Spline smoothing or regularization approach controls the degree of smoothness by a single parameter that penalizes the function‟s roughness. This parameter, denoted as the smoothing parameter ( ), compromises between the goodness of fit and the smoothness. To formulate this method, we first need to define „goodness of fit‟ and „roughness‟ criteria.

The „goodness of fit‟ is measured by the least squares criterion:

2 SSE ∑ |yj - x(tj)| j where x is the fitted data and y is the observed data.

„Roughness‟ of a function can be quantified by the total curvature criterion: 46

2 2 PEN2 ∫|D x(t)| dt

2 where D denotes the second derivative operator. A smaller PEN2 implies a less variable function, whereas a larger PEN2 indicates a rougher curve. The roughness criterion can be generalized to any arbitrary order of a function‟s derivative.

A penalized residual sum of squares is formed by putting the two opposing criteria SSE and PEN2 together:

PENSSE SSE + * PEN2 where is the smoothing parameter that controls the weight of data fit and smoothness.

As mentioned earlier, the smoothing parameter ( ) controls the trade-off between the closeness to the observed values and the smoothness. The amount of imposed smoothness, however, cannot be arbitrary. If is close to zero, we obtain an estimate close to the data and if is too large, we obtain an estimate equivalent to the linear regression estimate of the data and consequently, the details of the signal of interest may be diminished. Therefore, it is important to select a reasonably good smoothing parameter.

An appropriate smoothing parameter may be chosen subjectively by visual judgment and prior knowledge of the process generating the data. An objective, data-driven method is also developed using the generalized cross-validation (GCV) measure (101):

n SSE GCV( ) ( ) ( ) n - df( ) n - df( ) 47 where n is the number of observations, SSE is the mean square error and df( ) is the trace of the smoothing matrix (53). The optimum minimizes GCV( ) function plotted against log10 . GCV is the most popular procedures for selecting an optimal value for the smoothing parameter. However, like other cross validation methods, GCV tends to under- smooth the data.

3.3.4.2. Functional Principal Components Analysis

Principal component analysis explores main modes of variation in high dimensional data on a set of orthogonal, linearly uncorrelated directions. The functional counterpart of

PCA (fPCA) visualizes variability among a set of functional observations, which theoretically have infinite number of data values. fPCA returns a set of orthogonal weight functions or principal components (PCs) spanning the same range as the functional data.

Each PC accounts for a percentage of the total variability in data. Depending on how much variability one would like to explain, the first few PCs are retained to estimate the data.

3.3.4.3. fPCA algorithm

Like ordinary PCA, functions are first centered, i.e. the mean across all observations is removed from each observation, to avoid the first PC to mostly reflect the average response. Basically, fPCA of n observations xi, i 1,…,n finds the PC weight function ξ for which the PC score defined as:

i ∫ ξ(t) xi(t) dt

2 maximizes ∑ i constrained to: 48

∫ ξ2(t) dt ‖ξ‖2 1.

After computing the first PC weight function ξ1 and the first PC score µ1, second PC weight function ξ2 and the second PC score µ2 are calculated with an additional constraint:

∫ ξ2(t)ξ1(t) dt 0 which requires the new PC weight function to be orthogonal to that computed on the previous step. This constraint requires the new PCs to reflect new directions of variability. After each step, the amount of explained variability decreases. A sequence of descending PC scores and the corresponding weight functions are computed stepwise. By definition, µj and ξj are referred to as the eigenvalues and eigenfunctions.

To compute the functional PCs, one can formulate the fPCA as the eigenanalysis of the bivariate covariance function:

v(s,t) ∑n x (s)x (t). i i i

The functional eigenequation is then expressed as:

∫ v(s,t)ξj (t) dt jξj (s)

where µj is the j-th eigenvalue and ξj(s) is the j-th eigenfunction of the covariance function. Computer software is developed to solve the above eigenequation for pairs of ξ and µ. 49

Only the first few PCs are considered for further analysis and the remaining PCs are regarded as either irrelevant or are assumed to reflect observational error or noise. In this case, the problem is to figure out how many components are to be considered. A solution is to plot the base 10 logarithm of eigenvalues according to their size (against their indices j) and to identify the so-called “elbow” point at which the slope of the graph changes from “steep" to “flat". Then, the researcher may decide to retain only the components before the elbow. This method is called the scree or elbow test and is yet considered subjective. A traditional approach includes the components with eigenvalues greater than average. Another method is to include as many as components that account for a specific cumulative percentage of total variability.

3.3.4.4. Functional Canonical Correlation Analysis

Functional canonical correlation analysis (fCCA) is an exploratory technique that examines modes of variation that pairs of functions share (102). By fCCA, we would like to explore how variability in Hb and HbO2 data are interrelated. Given n pairs of observations (xi(t),yi(t)), i 1, …, n, fCCA estimates pairs of functions (ξj(t),휂j(t)) that most explain the interaction between x and y.

As in fPCA, data are first centered to focus on the analysis of variation from the mean

(102). ξ and 휂 are termed canonical weight functions and are estimated such that the correlation between the following integrals is maximized:

ξ ∫ ξ(t) xi(t) dt and ∫ (t) yi(t) dt for i 1, …, n. 50

( ξ , ) are called canonical variates and are calculated by maximizing the squared canonical correlation defined as:

2 *∑i ξ + R2(ξ, )

*∑i ξ + *∑i +

Similar to fPCA, one can obtain a non-increasing series of squared canonical correlations by imposing the condition that successive canonical weight functions to be orthogonal

(53).

However, imposing a strong roughness penalty on the computed weight functions ξ and 휂 is a rule to obtain meaningful results. A cross-validation criterion for objective calculation of the roughness penalty is described by Leurgans et al. as follows (102):

The roughness penalty for the smoothed weight functions is obtained by finding a value for the smoothing parameter ( ) that maximizes the squared correlation of n pairs of

(-i) (-i) scores ( ξ (t)x (t)dt , (t)y (t)dt), where ξ(-i) and (-i) are the smoothed canonical ∫ i ∫ i weight functions with the i-th curve omitted.

3.4. Results

3.4.1. fNIR Data Pre-processing

Raw optical intensity data for 730 nm and 850 nm wavelengths were first filtered using a finite impulse response low pass filter with a cut-off frequency of 0.14 Hz to remove high frequency noise, respiration and heart pulsation artifacts. The filtered raw data were converted to changes in Hb and HbO2 concentrations relative to the mean value of the 51 optical intensity during the first 20 s of the baseline. Hb and HbO2 data were registered in reference to the time when the hand was immersed in the ice water.

All functional data analyses were conducted in MATLAB (R2011a, MathWorks,

Natwick, MA) using the fDA package for MATLAB (53). Hb and HbO2 were smoothed by imposing a penalty on the roughness of the second derivative of the data with a of

104. The smoothing parameter was chosen empirically since the GCV plot resembled a sigmoid function which was not informative. Figures 11 and 12 show the non-smooth and smooth Hb and HbO2 data for two „far‟ channels located on the right and left sides of forehead, respectively.

52

Figure 11. Non-smooth and smooth b and bO2 data for 20 subjects for a ‘far’ channel located on right forehead: a) non-smooth HbO2; b) smooth HbO2; c) non-smooth Hb; d) smooth Hb. The smoothing parameter was 104. The black dashed line represents the mean response across 20 subjects. The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus).

53

Figure 12. Non-smooth and smooth b and bO2 data for 20 subjects for a ‘far’ channel located on left forehead: a) non-smooth HbO2; b) smooth HbO2; c) non-smooth Hb; d) smooth Hb. The smoothing parameter was 104. The black dashed line represents the mean response across 20 subjects. The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus).

54

3.4.2. Results of fPCA

The overall mean function of Hb and HbO2 across all subjects was first subtracted from each observation before performing the fPCA. Results of fPCA of Hb and HbO2 data were obtained separately for the six channels. No significant variation between channels was observed in the calculated PCs. Thus, in favor of saving space, we show and discuss the figures for two left and right „far‟ channels shown in Figures 11 and 12. However, overall results for other channels will be discussed in the discussion section.

We estimated the number of components based on the total explained variance. We desired to account for 95% of total variability; therefore, the first four PCs were considered.

The four considered PC weight functions of HbO2 are presented in Figures 13a and 13b for the right and left „far‟ channels, respectively. The first PC always has the highest PC score and accounts for the highest variability; the remaining PCs are associated to descending PC scores and account for less amount of total variability.

A visualization technique that may simplify interpretation of fPCA results is to plot the

PCs as perturbations of the mean function (91). In doing so, a multiple of each PC is added and subtracted to the mean function. By definition, this factor is 0.2 times the root mean square of the difference between the PC and its overall average (91). Results for

PCs perturbations of HbO2 are shown in Figures 13c and 13d for the right and left „far‟ channels, respectively.

PC1 is mostly associated with amplitude variation, particularly during the stimulus and post-stimulus conditions. PC2 accounts for a considerably smaller percentage of total 55 variability and corresponds to amplitude variation mostly during pre-stimulus condition.

PC3 and PC4 account for a very small percent of total variability and reflect some small variations during the pre-stimulus and the last minute of post-stimulus conditions.

Figure 14 shows the same information as Figure 13 for Hb data. One may identify some similarities between the PCs calculated for Hb and HbO2 data. For both right and left channels, PC1 is associated with an overall amplitude variation after the baseline. PC2 and PC3 explain amplitude variation during the pre-stimulus and post-stimulus and a phase shift in early post-stimulus condition. PC4 corresponds to some already identified variations mostly during pre-stimulus condition.

56

Figure 13. The first four principal components (PC) of HbO2 (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) ‘far’ channel. In (a) and (b), blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus).

57

Figure 14. The first four principal components (PC) of Hb (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) ‘far’ channel. In (a) and (b), blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus).

58

3.4.3. Rotating PCs

As we saw in the previous section, each considered PCs did not give substantive localized information of the variability; that is several PCs may explain variation during an experimental condition or a single PC may correspond to variation during several conditions. Thus, the interpretation of PCs is not straightforward.

Although, the original PC weight functions are computed to be optimal, it does not mean that there are no other alternatives that can do just well. One solution is to rotate the original PC weight functions by a rotation matrix. Varimax is the most popular rotation method for both multivariate and functional analyses (91). The rotated PC weight functions by varimax remain orthogonal as the original PCs but they are not necessarily uncorrelated. Moreover, the rotated PC weight functions are not associated to descending

PC scores, unlike the original PCs. However, the cumulative variability explained by the rotated PCs is the same as before the rotation.

A varimax approach finds an orthogonal rotation matrix (T) to transform the PC weight functions:

A = T.B where B is a matrix containing the considered PC weight functions. A varimax solution is

2 obtained by maximizing the variance of the vector of amj; where amj are the elements of

A. This is only achievable when the amj are relatively large or relatively small. As such, the rotated weight functions have maximized variance due to few large elements and many minuscule elements. Consequently, each PC is associated with one or a small number of modes of variation, which considerably enhances interpretability. 59

Results of the varimax rotation of HbO2 and Hb data are shown in Figures 13 and 14, respectively. The weight functions rotated by the VARIMAX rotation still account for the same cumulative percentage of total variability but in different proportions.

Each rotated PC weight function (Figures 15a-b and 16a-b) shows variation during a brief time window compared with unrotated PC weight functions (Figures 13a-b and 14a-b) that described variation over a long time span. This effect of rotation is more evident in the mean function perturbations (Figures 15c-d and 16c-d).

For HbO2 data (Figure 15), the two most important PCs that account for the largest percentage of total variability identify the evoked response during the stimulus condition and the recovery during post-stimulus condition. The PC weight function associated to the stimulus condition identifies slope and amplitude variation of the evoked response and a temporal shift in the peak of response across the 20 subjects. The PC weight function associated to the post-stimulus condition corresponds to amplitude and timing variation of the hemodynamic recovery curve. The two least important PCs are associated to variations in the pre-stimulus and end of the post-stimulus conditions.

For Hb data (Figure 16), the largest variability is observed during the post-stimulus condition by PC2 and PC3 for both channels; one associated to variation in the hemodynamic settlement at the end of the post-stimulus and the other one associated to variation of peak of the evoked response occurred with a delay after removal of the stimulus. The two least important PCs identified variation during the stimulus and pre- stimulus conditions.

60

Figure 15. The first four rotated principal components (PC) of HbO2 (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) ‘far’ channel. In (a) and (b), blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus).

61

Figure 16. The first four rotated principal components (PC) of Hb (a-b) and the corresponding mean function perturbations (c-d) for a right (a-c) and a left (b-d) ‘far’ channel. In (a) and (b), blue, green, red, and cyan curves correspond to PC1, PC2, PC3, and PC4, respectively. In (c) and (d), the blue curve represents the mean function and the red and green curves are the effects of subtracting and adding a multiple of the mean function (see Section 2 in Results). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus).

62

3.4.4. Results of fCCA fCCA was applied to Hb and HbO2 curves to explore their interaction and shared variation.

We used the cross-validation criterion proposed by Leurgans et al. to find the optimum value of the smoothing parameter ( ) for the canonical weight functions (102). Based on the cross-validation results and visual inspection of plotted canonical weight functions, a lambda of 107 for „far‟ channels and a lambda of 108 for „near‟ channels was selected.

We present and discuss the first canonical weight functions, a.k.a. the leading weight functions, for two „far‟ channels and two „near‟ channels on the right and left sides of forehead (Figure 17). Results for the other „far‟ channels were similar to the presented

„far‟ channels.

Figure 17 shows that variation from the norm for Hb and HbO2 is positively correlated throughout the experiment for all channels. For the right „far‟ channel, however, the dominant mode of covariation exists in the peak and timing of Hb and HbO2 changes, with HbO2 variation being smaller and much faster than Hb response. 63

Figure 17. Canonical correlation analysis of b (blue) and bO2 (red) data for a pair of ‘far’ channels (a – right, b – left) and the pair of ‘near’ channels (c – right, d – left). The three vertical black lines from left to right indicate the hand immersion incidents in tepid water (pre-stimulus), cold water (stimulus), and tepid water (post-stimulus).

64

3.5. Discussion

Functional data analysis (fDA) characterizes the main features shared by a set of observations from a number of individuals. In this research, we proposed the fDA as a novel tool for the analysis of fNIRS data.

We showed that fPCA is capable of decomposing the Hb and HbO2 time-series collected from a number of individuals during a CPT task into experimental conditions.

Multivariate PCA has been previously used to identify the systemic interference and remove it from fNIRS measurements. Zhang et al. used spatial eigenvector-based analysis of baseline activity in diffuse optical imaging (103). They used dozens of optodes over a large area of the cortex and showed that baseline information can be used to estimate global physiological interference during a finger tapping task and a tactile stimulus. Virtanen et al. compared PCA and ICA (independent component analysis) performance in removing superficial physiological interference during hypo-capnia and hyper-capnia maneuvers (104). Instead of using a fine grid of source-detector pairs, they used two channels with short (1cm) and far (3cm) source-detector distances and showed that both PCA and ICA are capable of effectively removing the physiological activity from the fNIRS measurements. fPCA is, too, an exploratory tool to capture the principal directions of variation in the population and dimension reduction. The fundamental assumption is that the true underlying process generating the data is, in essence, a continuous phenomenon which is measured with observational noise. As such, fPCA is performed on a space of curves collected over a continuum - which could be either time or space. fPCA characterizes the 65 geometry of curves as a set of principal directions of variation that aid a more straight feature selection.

Our study provides a good understanding of the variation in the population structure of

Hb and Hbo2 at different conditions of a CPT. Our observations are summarized as follows:

 The initial 30 s baseline condition was not identified by fPCA, probably because the

baseline hemodynamic activity is assumed to be fairly robust with no significant

variation between healthy individuals. Generally, inter-subject variability in fNIRS

measurements is explained by individuals‟ differences in anatomical factors such as

skull and cerebrospinal fluid (CSF) structure, the distribution of vessels, and the ratio

of arteries and veins. These factors did not cause a variable baseline hemodynamics in

our small sample and hence, it could not be identified by fPCA. However,

individuals‟ responses to a noxious stimulus and the following recovery are highly

subjective and variable. We observed that the hemodynamic response at the stimulus

and post-stimulus conditions was highly variable between subjects and the highest

variability was seen in the HbO2 data.

 The variation in the pre-stimulus condition accounted for a small percentage of total

variability (less than 16% for HbO2 and 16% on average across all channels for Hb).

It yet suggests that a hand immersion in the tepid water evokes a variable response

that is detectable by fPCA. 66

 The variability in the stimulus condition for HbO2 was revealed in the slope,

amplitude and timing of the peak of the response. The temporal shift was observed in

the „far‟ channels only.

 For the post-stimulus condition, an early component corresponding to variation in the

hemodynamic recovery and a late component corresponding to the hemodynamic

settlement were identified:

o The early component was the most important component for HbO2 data and

showed variation in the amplitude and timing of the peak of response. No slope

variation was found during the post-stimulus condition. This component was more

important in the right channels than left channels. For Hb, this component was

associated to variation in the amplitude of response which peaked a few seconds

after the removal of the stimulus. This component was more important in the left

„far‟ channels than the right „far‟ channels and the „near‟ channels for Hb data.

o The late post-stimulus component was more important in Hb data than HbO2 data.

It accounted for most of the variation in the „near‟ channels and a large portion of

variation in „far‟ channels for Hb. fPCA is a pure exploratory tool that requires no prior information of the experiment and makes no assumptions on the form of curves. It can be helpful when timing of an evoked response relative to a stimulus is unknown or uncertain, such as in decision making or emotional reaction experiments (93). We showed that fPCA successfully identified experimental conditions during a CPT across 20 subjects. For the post-stimulus condition, for example, fPCA identified two components in Hb and HbO2: an early 67 component associated to variation in the hemodynamic recovery and a late component associated to the hemodynamic settlement. This qualitative information could not be easily obtained directly from the raw data in Figures 11 and 12. fPCA also quantifies the weight of each component as a percentage of total variation. We showed that the components and their weight varied between Hb and HbO2 parameters and channels. This information may be helpful to decide which parameter and/or channels are most associated with the hypothesized effect, in brain mapping studies for instance. With the CPT experiment, we previously showed that the amplitude and timing of the evoked response of HbO2 was correlated with subjects‟ report of the pain intensity

(96).

After exploratory steps, one can extract more pertinent information. For instance, fPCA is a good tool for computing proximities between curves in a reduced dimensional space.

Using such a distance measure, one can investigate whether the curves are similar or can be split into several classes.

In a secondary exploration, we investigated the association between Hb and HbO2 variability using fCCA. We were interested in the dominant modes of correlation between Hb and HbO2. Results demonstrated an overall positive correlation between Hb and HbO2 variation through the CPT experiment. However, for a right „far‟ channel, there was an evident time shift in the Hb response relative to HbO2. In general, fCCA gives a broad view of the interaction between two variables. The identified patterns of covariation give important directions for further investigation of interaction between Hb and HbO2 in a particular experimental setting. 68

Due to inherent interaction between the autonomic nervous system and central pain processing system, researchers try to assess the individuals‟ perception of pain through physiological signs. Hemodynamic changes at the skin and major arteries have been investigated to study the sympathetic response to a noxious stimulus (14-16). In this exploratory study, we attempted to learn what features of response to a cold noxious stimulus is most variable among control subjects. Through the results of this research, we learned about importance of the dynamics of the evoked response to a CPT as revealed by fPCA. In another study, we used this information and considered the derivatives of

HbO2 curves during a pain tolerance test using CPT. Results confirmed that the amplitude and timing of peak of the 1st and 2nd derivative of the HbO2 response right after hand immersion in cold water is a very robust feature. In future, we will use this knowledge to better understand the physiology of pain through explaining some of the observed variations by known factors such as age, gender, or information about whether the subjects are high/low tolerance to that specific stimulus.

The observed modes of variation and covariation patterns can help improve design of fNIRS studies for pain research in particular and for other experiments in general. This study was preliminary but opens an interesting avenue for future research. Without a good understanding of population structure, it would not be feasible to formulate and study physiological phenomena and identifying variability is the first step in this direction. 69

3.6. Acknowledgment

Authors would like to acknowledge Dr. Patricia A. Shewokis to suggest the functional data analysis for the cold pressor data. The first author would like to thank Dr. Shewokis for instructing her in the topic.

70

CHAPTER 4: HEMODYNAMIC RESPONSE TO COLD TOLERANCE TESTS AT DIFFERENT TEMPERATURES

4.1. Introduction

The fact that pain perception is associated with simultaneous activation of the autonomic nervous system (ANS) and the central pain processing in the brain has been often disregarded in pain neuroimaging studies. Response to noxious stimuli has been reported to be related to increase in cardiovascular parameters such as heart rate and blood pressure (6, 59, 105). This effect is more pronounced in painful stimuli that are known to evoke a substantial autonomic response. Cold pressor test (CPT) is such a stimulus that has been used in many pain studies (106-108). The deep, aching sensation induced by prolonged immersion of a limb in cold water is assumed to best mimic the condition of chronic pain (109). However, the concomitant effect of ANS and central pain processing in response to a cold water challenge has not been studied well.

Neuroimaging studies of pain using fMRI found significant correlation between perceived pain and BOLD signal in several cortical areas including medial prefrontal cortex (7). They showed that heart rate as an index of autonomic arousal was correlated with only a small fraction of pain coding specific activated pixels.

Gender differences in the processing, perception, and reaction to pain is well recognized.

Females are at higher risk of chronic pain associated with following conditions: fibromyalgia, irritable bowel syndrome, rheumatoid arthritis, osteoarthritis, temporomandibular joint disorder, chronic and migraine (110). 71

Besides psychosocial factors that influence gender differences in pain perception (111), animal (112) and human studies (113) suggest that neurophysiological differences may contribute to gender-based differences in pain perception and processing.

This research was designed to investigate sensitivity of healthy individuals to a CPT at different temperatures of water. As a secondary goal, we aimed to study the gender differences during a sustained noxious stimulus in three aspects: 1) subjective experience of pain or pain perception, 2) the autonomic reactivity as a vegetative response to pain, and 3) higher level central systems that process pain. The results would be of help in determining vulnerable subjects to developing chronic pain. For example, it has been shown that sensitivity to cold pain is a risk factor for developing post-operative pain

(114). Gender-based differences in the autonomic response and central processing of pain are also of interest in order to unravel sex-specific biological mechanisms during nociception.

4.2. Methods

4.2.1. Subjects

Twenty one right-handed individuals (11 females) were recruited from the Drexel

University community. Subjects claimed to have no history of neurological or psychological disorders. They attended an orientation session in which they received detailed information on the experimental paradigm and signed the informed consents approved by Drexel University institutional review board. Subjects were instructed to refrain from smoking and drinking alcohol at least three hours prior to the experiment. 72

4.2.2. Instrument

Same instrumentation was used in this research as the repeated trials of a CPT. For details, please refer to Chapter 2. fNIRS data were collected continuously from two multi-distance probes located on the right and left sides of forehead. For each probe, two detectors were placed at 2.8 cm distance from the light source (far channels), and one detector was located at 1 cm from the light source (near channel). For the statistical analysis, the averaged response of the two far channels on each side was used.

4.2.3. Protocol

A tolerance test using cold water at different temperatures was employed to investigate individuals‟ pain response to different intensities of a cold stimulus. Four trials of hand immersion in cold water at 15C, 10C, 5C, and 1C were used for each subject to generate low, moderate and severe pain levels. After a baseline recording for 30 s, subjects were asked to immerse their right hand in the tepid water for 2 min for adaptation. Then, the experimenter asked them verbally to put the same hand in a constant temperature bath for as long as they could tolerate the stimulated pain, but no more than 5 min. During each experiment, in addition to subjects‟ pain threshold (when the first pain was felt) and tolerance (when the pain intensity became intolerable), numerical pain rating scores on a

0 to 10 scale, where zero means no pain and ten means an intolerable pain, were recorded every 15 s. A block diagram of the protocol is shown in Figure 18. 73

Figure 18. Block diagram of the second protocol

4.2.4. Features

Based on previous study of the hemodynamic response to CPT, we found four features of deoxy-hemoglobin (Hb) and oxyhemoglobin (HbO2) time-series that best reflect the nature of the response:

1. Delta_Hb defined as the difference between Hb value before the hand immersion

in cold water and its peak during the hand immersion in cold water or a few

seconds after it.

2. Delta_HbO2 defined as the difference between HbO2 value before the hand

immersion in cold water and its peak during the hand immersion in cold water or

a few seconds after it.

3. Maximum velocity of the evoked HbO2 (MV) defined as the peak of the 1st

derivative of HbO2 occurring a few seconds after hand immersion in cold water.

4. Maximum acceleration of the evoked HbO2 (MA) defined as the peak of the 2nd

derivative of HbO2 occurring a few seconds after hand immersion in cold water. 74

4.2.5. Statistical Data Analysis

For the analysis of hemodynamic parameters collected by fNIRS, we used linear mixed effects models (115, 116). Linear mixed models provide a flexible and powerful approach for the analysis of repeated measures. In a linear mixed model approach, the dependent variable is modeled as a combination of population characteristics, called fixed effects, and subject specific effects, called random effects that are unique to a particular individual.

Fixed effects are in fact regression coefficients that model the relationship between a dependent variable and a set of known categorical predictors, called fixed factors, and continuous covariates. In linear mixed models, fixed effects are unknown and are estimated based on the given observations.

Random effects are associated with random deviation of individuals‟ trajectory from the overall fixed effects. For instance, a random intercept effect describes the deviation from the overall fixed intercept for a specific subject and a random slope effect describes the deviation from the overall fixed slope for a certain subject. The associated random factors with random effects are known variables with levels that are randomly sampled from a larger population of levels. It means that not all the possible levels of the random factors are studied and eventually, we would like to generalize the research findings to the population.

Repeated measures data involve multiple observations on the same subject across a repeated factor which may be time or different experimental conditions. These repeated measurements are likely to be correlated and therefore, any model fit requires parameter estimation of the covariance structure. Classical repeated measures ANOVA assumes the 75 sphericity of the covariance matrix. If this assumption is violated, Geisser and

Greenhouse and Huynh and Feldt have provided adjustments to the test statistics. In contrast, linear mixed models allow a wide selection on the form of the variance- covariance matrix. This option provides efficiency and flexibility compared to classical

ANOVA.

Another remarkable advantage of linear mixed models is their capability of handling a great deal of missing data. Classical ANOVA excludes entire data for a subject if only one observation is missing. Linear mixed models allow the inclusion of subjects with missing data if the data are missed at random

To explain our approach, we use the delta_HbO2 feature. The scatter plot of delta_HbO2 across temperature for 7 subjects is shown in Figure 19. It is evident that individuals‟ regression line (or trajectory) differ in their intercept as well as slope. Therefore, it would be desirable to describe the data using a mixed effects model that considers the variable intercept and slope.

76

Figure 19. Scatter plot of delta_HbO2 across temperature for 7 subjects. Temperature is considered a continuous factor.

To formulate our linear mixed effects model, we first consider a general specification for the t-th observation for the i-th subject:

where Channelit is zero or 1 if the observation Yit belongs to the near or far channel, respectively; Sideit is zero or 1 if the observation Yit belongs to right or left side of forehead, respectively; Genderi is zero or 1 if the observation Yit belongs to a male or female subject, respectively.

In this model subjects have different intercepts and slopes for temperature variable. This model is composed of two sets of covariates associated with the population-specific fixed 77 effects β0, …, β7 and the subject-specific random effects u0 and u1. ti is the random error for the i-th subject and t-th observation. The index i takes 1 to 21 (for 21 subjects) and the index t indicates the observation index. Ideally, each individual has a sequence of measurements occurring at 4 temperatures, 2 sides of forehead, and 2 channels (a total of

16 observations per subject). However, the number of observations may be different between subjects (due to missing data), i.e. observations for the i-th subject include all or a subset of the following observations:

Yi1 : observation for temperature 1C, near channel, right side

Yi2 : observation for temperature 1C, near channel, left side

Yi3 : observation for temperature 1C, far channel, right side

Yi4 : observation for temperature 1C, far channel, left side

and Yi5 to Yi8 for temperature 5C, Yi9 to Yi12 for temperature 10C, and Yi13 to Yi16 for temperature 15C, accordingly.

In a matrix notation, the linear mixed model can be formulated as follows:

Yi = Xiβ + Ziui + εi

where the first part (Xiβ) represents the fixed effects and the second part (Ziui + εi) represents the random effects and implies a correlated error structure. This model describes the mean response trajectory for the i-th subject. The mean response profile in the population is ( ) .

The vector of observations Yi is a ni × 1 vector of the dependent variable, where ni is the number of observations for the i-th subject: 78

[ ].

The design matrix Xi has a dimension of ni × 8 for which the first column corresponds for the intercept and the other columns contain the known values of the 7 covariates

(Temperature, Channel, Side, Gender, Channel*Side, Channel*Gender, and

Channel*Side*Gender) for each of the ni observations on the i-th subject:

[ ]

where T stands for Temperature, C for Channel, S for Side, and G for Gender.

The design matrix Zi has a dimension of ni × q containing the known values of a subset of total covariates that have a random effect on the dependent variable across subjects. For our case, we assumed that both the intercept and slope of individual trajectories over the repeated measure – i.e. temperature - vary randomly between subjects. Therefore, the Zi matrix would have two columns: a column of 1‟s for the random intercept and another column for the random slope:

[ ].

The ui vector for i-th subject contains the associated random effects – i.e. random intercept and random slope for our case. The random effects are assumed to have a normal distribution with a mean of zero and a variance-covariance matrix D: 79

* + ( ).

Here, D is a 2 × 2 matrix that contains the variance of each random effect of the ui vector and the covariance between the two random effects – i.e. intercept and slope:

( ) ( ) [ ]. ( ) ( )

We imposed no constraints on the form of the D matrix. So, the D matrix was an unstructured matrix.

The error term εi is a vector of ni residuals associated with observations. We assume that the residuals for the i-th subject have a normal distribution with a mean of zero and a ni × ni variance-covariance matrix Ri:

[ ] ( ).

We also assume that residuals for different subjects are independent of each other and ε and u are mutually independent. In our case, we assumed a diagonal structure for the Ri matrix. This structure implies that residuals for observations on the i-th subject are uncorrelated and have equal variance. Therefore, Ri matrix can be simply identified by the variance of εi:

2 Ri = Var(εi) σ I, where I here is a 16 × 16 identity matrix.

Application of a linear mixed effects model requires the specification of the mean process together with a model for the covariance structure of the responses on each subject. To 80 develop an appropriate mixed effects model, we followed an approach similar to recommendation of Diggle (117). The first step is selecting an initial set of fixed effects in sufficient generality to specify the mean structure. The second step is selecting random effects and intr-individual variation to specify a model for the covariance structure.

Akaike‟s information criterion (AIC) and Bayesian information criterion (BIC) were used to compare the models with different covariance structures.

4.3. Results

4.3.1. Analysis of maximum reported pain, pain threshold and pain tolerance

The box plot of pain threshold, pain tolerance, and maximum reported pain score during the CPT across four temperatures are shown in Figure 20a-c.

Pain threshold, pain tolerance, and maximum reported pain score were tested with

Friedman‟s ANOVA by Ranks. There was a significant difference in the median value of the pain threshold, tolerance, and maximum score with respect to temperatures (2 (3) =

42.43, p-value < 0.001; 2 (3) = 23.73, p-value < 0.001; and 2 (3) = 31.97, p-value <

0.001, respectively).

81

(a)

(b)

(c) Figure 20a-c. Box plots of (a) pain threshold, (b) pain tolerance, and (c) maximum reported pain score during the CPT across four temperatures: 1C, 5C, 10C, and 15C for 21 subjects. 82

Post hoc Wilcoxon signed ranks tests adjusted by Bonferroni criterion yielded that:

1) Pain threshold increased significantly by water temperature, except for 1C to 5C.

It was revealed that the pain threshold at 1C was significantly lower than the pain

threshold at 10C and 15C (Z = -3.60, p-value < 0.001 and Z = -4.02, p-value <

0.001, respectively). Also, the pain threshold at 5C was significantly lower than

the pain threshold at 10C and 15C (Z = -3.65, p-value < 0.001 and Z = -3.91, p-

value < 0.001). Moreover, the pain threshold at 10C was significantly lower than

the pain threshold at 15C (Z = -3.01, p-value < 0.01).

2) Pain tolerance at 1C was significantly lower than tolerance at higher temperatures,

i.e. 5C, 10C, and 15C (Z = -3.05, p-value < 0.01, Z = -2.95, p-value < 0.01, and Z

= -3.17, p-value < 0.01, respectively).

3) Maximum pain score reported during the CPT decreased significantly by increase

in water temperature. It was revealed that the maximum pain score at 1C was

significantly higher than the maximum pain score at 5C, 10C, and 15C (Z = -3.48,

p-value < 0.001, Z = -3.46, p-value < 0.01, and Z = -3.46, p-value < 0.01,

respectively). Also, maximum pain score at 5C was significantly higher than the

maximum pain score at 10C and 15C (Z = -2.89, p-value < 0.01 and Z = -3.13, p-

value < 0.01)

The correlation analysis between the pain tolerance data and objective parameters measured by fNIRS was performed for 1C only because the variability of pain tolerance data in other temperatures was questionable. We found significant correlation between pain tolerance at 1C and delta_Hb in all four channels: the right near channel (R = -0.54, 83 p-value = 0.02), right far channel (R = -0.51, p-value = 0.02), left near channel (R = -

0.55, p-value = 0.02), and left far channel (R = -0.59, p-value = 0.006).

Furthermore, linear regression analysis yielded that pain tolerance at 1C can be predicted by delta_HbO2 at the right near channel, delta_Hb at the left near channel, MV at the right near channel, and MV at the right far channel (Table 7).

Table 7. Regression Coefficients

Predictors Unstandardized Coefficients Standardized t Sig. Coefficients

B Std. Error Beta

(Constant) 107.94 42.24 2.56 .029

delta_N_L_Hb_0 -62.93 16.06 -.58 -3.92 .003

delta_N_R_HbO2_0 57.04 23.79 .99 2.40 .037

mv_N_R_0 -7499.54 1743.22 -1.80 -4.30 .002

mv_F_R_0 2397.13 909.69 1.02 2.64 .025

R = 0.89, R2 = 0.79

84

4.3.2. Statistical Analysis of Hb and HbO2

Samples of Hb and HbO2 time series for four different temperatures collected at a far channel located on the right side of the forehead are shown in Figure 21. This typical subject could complete the CPT at 5C, 10C, and 15C. At 0C, he/she removed his/her hand from the cold water before the 5 min cut-off duration.

Figure 21. Sample of Hb and HbO2 time series from a far channel on the right side of forehead during a cold pressor test (CPT) for four temperatures. The black vertical lines from left to right indicate the baseline recording, hand immersion in the tepid water, hand immersion in cold water, and reaching the pain tolerance.

85

We tested four dependent variables: delta_Hb, delta_HbO2, MV, and MA. In summary, there was a significant main effect of Channel and Side and a significant interaction between Channel and Gender in all the considered features. For the common significant interaction, post hoc multiple comparisons yielded that responses at the near and far channels were significantly different in males and not in females. Also, a significant difference between males and females was revealed at the far channels and not the near channels.

For HbO2 features, i.e. delta_HbO2, MV and MA, a significant interaction between

Channel and Side was found. Post hoc analyses showed that the right and left responses were significantly different for the far channels only and not the near channels.

Moreover, the near and far channels were significantly different for the right side of forehead and not the left side.

Furthermore, an interesting three-way interaction was found between Channel, Side, and

Gender in delta_HbO2. Post hoc t-tests yielded that the right and left responses were significantly different at the far channels for both males and females, whereas no such laterality was observed at the near channels neither for males nor for females. Also, the near and far channels were significantly different for the right side for males only.

Females did not show any significant difference between the near and far channels for neither sides of forehead. Males and females responses were significantly different only at the right far channel.

A detailed account of the results is followed. 86

4.3.2.1. Delta_Hb

The overall trend of delta_Hb change across temperature is shown is Figure 22. The red line in the scatter plot depicts the mean trajectory of delta_Hb. The line plot shows the mean of response and standard deviation at each temperature.

(a) (b)

Figure 22a-b. Scatter plot (a) and line plot (b) of the dependent variable delta_Hb versus

temperature.

The dependent variable delta_Hb described by the linear mixed model in (1) yielded a significant main effect of temperature, channel, and side (Figure 23), and a significant interaction between gender and channel (Figure 24).

87

(a)

(b)

(c)

Figure 23a-c. Line plots of the dependent variable delta_Hb versus temperature for the fixed factors Channel (a), Side (b), and Gender (c).

88

Figure 24. Line plots of the dependent variable delta_Hb versus temperature showing the interaction between Channel and Gender

Table 8. Type III tests of model covariates for the dependent variable delta_Hb

Source Numerator df Denominator df F Sig.

Intercept 1 19.176 93.236 .000

temp 1 17.608 12.661 .002

chan 1 246.810 18.817 .000

side 1 244.228 6.620 .011

gender 1 18.735 1.398 .252

chan * gender 1 246.964 8.258 .004

89

Delta_Hbit = -1.36 + 3.85E-2*Temperatureit + 1.10E-1*Channelit – 1.88E-1*Sideit –

4.40E-1*Genderi + 4.31E-1*Channelit*Genderi (1)

Post hoc pairwise comparisons on the significant interaction between Gender and

Channel using paired t-test adjusted by Sidak criterion revealed that:

- There is no significant difference between males and females responses at near

channels whereas there is a significant difference between males and females

responses at far channels (p-value = 0.043).

- The near and far channels responses were significantly different for males (p-

value < 0.001) but not significantly different for females.

4.3.2.2. Delta_HbO2

The overall trend of delta_HbO2 change across temperature is shown is Figure 25. The red line in the scatter plot depicts the mean trajectory of delta_HbO2. The line plot shows the mean of response and standard deviation at each temperature.

90

(a) (b)

Figure 25a-b. Scatter plot (a) and line plot (b) of the dependent variable delta_HbO2 versus

temperature.

The dependent variable delta_HbO2 described by the linear mixed model in (2) yielded a significant main effect of temperature, channel, side, and gender (Figure 26), and significant interactions between side and channel (Figure 27), gender and channel (Figure

28), and gender and side and channel (Figure 29).

91

(a)

(b)

(c)

Figure 26a-c. Line plots of the dependent variable delta_HbO2 versus temperature for the fixed factors Channel (a), Side (b), and Gender (c). 92

Figure 27. Line plots of the dependent variable delta_HbO2 versus temperature showing the interaction between Channel and Side.

Figure 28. Line plots of the dependent variable delta_HbO2 versus temperature showing the interaction between Channel and Gender. 93

Figure 29. Line plots of the dependent variable delta_HbO2 versus temperature showing the three- way interaction between Channel, Side, and Gender.

Table 9. Type III tests of model covariates for the dependent variable delta_HbO2

Source Numerator df Denominator df F Sig.

Intercept 1 18.558 54.916 .000

temp 1 19.738 8.005 .010

chan 1 253.611 16.443 .000

side 1 253.409 15.164 .000

gender 1 18.910 4.280 .053

chan * side 1 255.213 27.548 .000

chan * gender 1 253.638 8.767 .003

chan * side * gender 2 254.207 4.441 .013

94

Delta_HbO2it = 3.38 – 1.12E-1*Temperatureit + 2.66E-1*Channelit + 1.30*Sideit +

1.40*Genderi – 1.01*Channelit*Sideit – 8.40E-3*Channelit*Genderi –

1.18*Channelit*Sideit*Genderi (2)

Post hoc pairwise comparisons on the significant interaction between Channel and Side using paired t-test adjusted by Sidak criterion revealed that:

- There is a significant difference between right and left responses at far channels

(p-value < 0.001) whereas there is no significant difference between right and left

responses at near channels.

- The near and far channels responses were significantly different for the right side

(p-value < 0.001) but not significantly different for the left side.

Post hoc pairwise comparisons on the significant interaction between Channel and

Gender using paired t-test adjusted by Sidak criterion revealed that:

- There is a significant difference between males and females responses at far

channels (p-value = 0.009) whereas there is no significant difference between

males and females at near channels.

- The near and far channels responses were significantly different for males (p-

value < 0.001) but not significantly different for the females.

Post hoc pairwise comparisons on the significant interaction between Channel and Side and Gender using paired t-test adjusted by Sidak criterion revealed that:

- There is a significant difference between right and left responses at far channels

for both males and females (p-value < 0.001 and p-value = 0.001, respectively). 95

- The near and far channels responses were significantly different for the right side

for males only (p-value < 0.001).

- Males and females response at far channels were significantly difference for the

right side only (p-value = 0.001).

4.3.2.3. Maximum Velocity of the Evoked Response (MV)

The overall trend of MV change across temperature is shown is Figure 30. The red line in the scatter plot depicts the mean trajectory of MV. The line plot shows the mean of response and standard deviation at each temperature.

(a) (b)

Figure 30a-b. Scatter plot (a) and line plot (b) of the dependent variable MV versus temperature.

The dependent variable MV described by the linear mixed model in (3) yielded a significant main effect of temperature, channel, side, and gender (Figure 31), and significant interactions between side and channel (Figure 32) and gender and channel

(Figure 33). 96

(a)

(b)

(c)

Figure 31a-c. Line plots of the dependent variable MV versus temperature for the fixed factors Channel (a), Side (b), and Gender (c).

97

Figure 32. Line plots of the dependent variable MV versus temperature showing the interaction between Channel and Side.

Figure 33. Line plots of the dependent variable MV versus temperature showing the interaction between Channel and Gender.

98

Table 10. Type III tests of model covariates for the dependent variable MV

Source Numerator df Denominator df F Sig.

Intercept 1 17.806 83.130 .000

temp 1 20.239 22.136 .000

chan 1 255.040 20.301 .000

side 1 254.365 19.409 .000

chan * side 1 255.983 15.379 .000

gender 1 18.633 5.591 .029

chan * gender 1 255.099 17.357 .000

MV_HbO2it = 0.04 – 0.002*Temperatureit + 0.01*Channelit + 0.02*Sideit + 0.03*Genderi

–0.02*Channelit*Sideit – 0.02*Channelit*Genderi (3)

Post hoc pairwise comparisons on the significant interaction between Channel and Side using paired t-test adjusted by Sidak criterion revealed that:

- There is a significant difference between right and left responses at the far

channels only (p-value < 0.001).

- The near and far channels were significantly different for the right side only (p-

value < 0.001).

Post hoc pairwise comparisons on the significant interaction between Channel and

Gender using paired t-test adjusted by Sidak criterion revealed that:

- There is a significant difference between males and females responses at far

channels (p-value = 0.002) whereas there is no significant difference between

males and females at near channels. 99

- The near and far channels were significantly different for males (p-value < 0.001)

but not significantly different for the females.

4.3.2.4. Maximum Acceleration of the Evoked Response (MA):

The overall trend of MA change across temperature is shown is Figure 34. The red line in the scatter plot depicts the mean trajectory of MA. The line plot shows the mean of response and standard deviation at each temperature.

(a) (b)

Figure 34a-b. Scatter plot (a) and line plot (b) of the dependent variable MA versus temperature.

The dependent variable MA described by the linear mixed model in (4) yielded a significant main effect of channel, side, and gender (Figure 35), and significant interactions between side and channel (Figure 36) and gender and channel (Figure 37).

100

(a)

(b)

(c)

Figure 35a-c. Line plots of the dependent variable MA versus temperature for the fixed factors Channel (a), Side (b), and Gender (c). 101

Figure 36. Line plots of the dependent variable MA versus temperature showing the interaction between Channel and Side.

Figure 37. Line plots of the dependent variable MA versus temperature showing the interaction between Channel and Gender.

102

Table 11. Type III tests of model covariates for the dependent variable MA

Source Numerator df Denominator df F Sig.

Intercept 1 81.583 89.693 .000

temp 1 72235.955 2.499 .114

chan 1 303.900 9.668 .002

side 1 301.442 12.755 .000

chan * side 1 303.066 7.682 .006

gender 1 67.167 23.417 .000

chan * gender 1 303.951 18.654 .000

MA_HbO2it = 9.77E-4 – 4.31E-5*Temperatureit + 2.76E-4*Channelit + 4.37E-4*Sideit +

1.14E-3 *Genderi –3.83E-4*Channelit*Sideit – 6.03E-4*Channelit*Genderi (4)

Post hoc pairwise comparisons on the significant interaction between Channel and Side using paired t-test adjusted by Sidak criterion revealed that:

- There is a significant difference between right and left responses at the far

channels only (p-value < 0.001).

- The near and far channels were significantly different for the right side only (p-

value < 0.001).

Post hoc pairwise comparisons on the significant interaction between Channel and

Gender using paired t-test adjusted by Sidak criterion revealed that:

- There is a significant difference between males and females responses at far

channels (p-value = 0.006) and near channels (p-value < 0.001). 103

- The near and far channels were significantly different for males (p-value < 0.001) but

not significantly different for the females.

4.4. Discussion

Pioneering work on using CPT for pain study by Wolf and Hardy suggested that pain threshold to cold water is at 18C (27). Any decrease in water temperature is reported to be proportional to increase in pain score. The cold water temperatures in this research were chosen accordingly to evoke different levels of pain.

In the present study, we investigated nociceptive response and cardiovascular reactivity recorded by far and near channels. We observed bilateral activation in far and near channels with very similar trend of change through the time course of the experiment.

While no significant difference in the near channels responses were found, far channels on the right side showed significantly higher activation than left side. Also, for HbO2 features, we found significant difference between far and near channels on the right side only. We speculate that this laterality reflects asymmetrical activation in the prefrontal cortex during the CPT. This finding may suggest that there is a hemispheric response to noxious cold stimuli unilateral to the stimulus.

This result is consistent with a PET study using CPT at 6C (5). They showed ipsilateral increases in rCBF in the lateral prefrontal cortex (Brodmann‟s areas 10 and 46) (left hand immersion). A 133Xe SPECT study of CPT, however, found contralateral response in the frontal lobe (left hand immersion) (118). A comparative study of common brain regions activated to hot and cold plates stimuli showed bilateral prefrontal response and ipsilateral dorsolateral prefrontal (DLPF) activation (9). They reported that cold stimulus 104 evoked larger brain regions and these two pain modalities were significantly different in the frontal lobe. The activation of DLPF cortex to cold stimulus was observed in an fMRI study of cold induced in healthy subjects (8). Another study, however, did not find a significant activation in DLPF cortex during cold water immersion (1C) (119).

Time-series analysis of fNIRS parameters throughout the tolerance test was not conducted here. Since the autonomic response fairly quickly adapts to the stimulus while the pain perception may last for the entire duration of immersion, analysis of dynamics of the hemodynamic response may unravel some interesting mechanisms.

Gender-specific activation detected by fNIRS

Gender difference in perception and expression of pain has been widely addressed in literature (see for example (120)). Several studies have shown that women are more sensitive to cold pain than men. Reports on sex differences in cold pain tolerance are more prevalent than cold pain threshold. A comprehensive survey of literature during

1998 to 2008 on sex differences in pain perception found that 80% of CPT studies showed that females tolerate significantly less pain than males (121). Another review by

Fillingim et al. showed that only one study out of 23 on cold tolerance did not show a gender difference (120). This particular study included 15 male and 19 female participants. Riley suggests that considering the moderate effect size of threshold and tolerance, a sample of 41 subjects for each gender group is needed to reach an adequate power (122).

In our study, we could not find any sex difference in neither pain tolerance, nor pain threshold. Considering the small size of our sample (10 males, 11 females), it seems that 105 our study lacked enough power to catch the sex differences repeatedly reported in cold pain tolerance. However, since males and females pain tolerance and threshold are comparable in our sample size, it is plausible to compare their fNIRS signal under the condition of equivalent pain perception (to the same level of a sustained painful stimulus).

Our results indicate that there is no gender-specific skin hemodynamic response to a cold stimulus. On the other hand, a significant gender difference in the cortical hemodynamic parameters (measured by far channels) suggests a sex difference in cortical response to the painful stimulation.

For both genders, we observed a hemispheric lateralization in response to cold stimulus.

Particularly, a marked HbO2 activation at the right far channel suggests an ipsilateral increase in cortical blood flow. This activation was not significantly different from the near channel response in females (p-value = 0.069). This may be justified by the confounding effect of skin response to the cold water stimulus given the high sensitivity of fNIRS signal to superficial layers. Another reason could be the smaller size of the

HbO2 response in females requiring larger sample size. No gender difference was observed in HbO2 at the left far channel.

Sex differences in cerebral activation in response to experimental and clinical noxious stimuli have been studied using advanced neuroimaging modalities (113, 123-129).

Despite identifying some gender differences in cerebral responses, there seems to be a large variability in sex-related activation and deactivation patterns across studies. Three following neuroimaging studies found gender-specific activation in response to noxious stimuli in the prefrontal cortex. In an fMRI study, Moulton et al. discovered sex 106 differences in BOLD signal in response to noxious contact heat stimuli (127). They observed that women show significantly more voxels with negative signal change than men in the S1 and DLPF cortices. They stated that this sex difference may be in part explained by gender differences in baseline BOLD signal. In a PET study, Derbyshire et al. found significantly larger activation in males than females in response to equalized laser stimulation in several contralateral areas including prefrontal, primary and secondary somatosensory cortices (125). Sex differences in medial prefrontal cortex activation in response to sub-threshold and intense electrical stimulation was observed in an fMRI research showing a stronger activity in women (129). These studies suggest that the prefrontal cortex may play an important role in mediating gender differences in the neurophysiological response to painful stimulation.

We did not find any sex difference or lateralization in the near channels response.

Nevertheless, we cannot conclude that skin response is unlikely to be related to perceived pain intensity.

4.5. Conclusion

Our study has revealed cutaneous and cortical hemodynamic response to noxious cold stimulus. Moreover, we observed sex-specific cortical activation. This suggests sex differences in central processing of nociceptive information and that there may be underlying gender differences in the neural mechanisms that mediate pain perception.

The results of this research may help in producing a clear and consistent pattern of sex differences in human pain sensitivity. 107

The origin of signal change in the prefrontal cortex during a CPT is still unclear since the nociceptive component and autonomic aspect of the response are superimposed.

Complete dissociation of the two elements requires careful control of the experiment to include non-noxious challenges for ANS arousal and pure nociceptive stimuli for pain system process.

We used water stimulus at close to cold pain threshold (15C) to elicit less autonomic response. However, results showed that regardless of the water temperature, skin and deep responses were significantly different. In an effort to evoke a pure autonomic response, Harper et al. used the Valsalva maneuver and found significant activation in a number of brain regions including the medial and orbital prefrontal cortex that showed response to the CPT as well (130). Their observation suggests that the evoked response to the CPT in the prefrontal cortex may demonstrate ANS activation.

Because cold noxious stimulus evokes substantial cardiovascular reaction, differentiation of nociceptive processing and autonomic reaction to the CPT challenge requires careful adjustment of the experiment‟s parameters.

108

CHAPTER 5: SUMMARY, FUTURE WORK, AND CLINICAL APPLICATIONS

5.1. Summary

This PhD research suggests utilizing fNIRS technique for objective assessment of pain.

We found an association between a noxious stimulus and the evoked hemodynamic response and demonstrated that induced cold noxious stimulus evokes a reproducible and consistent hemodynamic response which can be reliably detected by fNIRS.

In the first set of experiments, we used three trials of a CPT to test the reliability of our measurement and study the patterns of adaptation to repeated noxious stimuli. We observed that a CPT evokes a reproducible and consistent bilateral hemodynamic response at the skin and deeper cortical layers. While no laterality of the response was seen at the skin, right hemisphere prefrontal cortex showed significantly higher activation than left hemisphere.

The measured hemodynamic response to an acute painful stimulus is mainly dominated by ANS reactivity. Our results suggest that the cortex might also be involved in at least some aspect of the observed nociceptive response. We also observed adaptation to repeated exposures to a CPT in both the hemodynamic parameters and subjective pain scores. Correlation analyses on pain rating scores and total oxygenation parameter showed moderate to large effects which can be translated to a meaningful association between subjective self-reports and objective measurements by fNIRS. This effect can be explained by inherent interaction between autonomic nervous system and central pain processing system. 109

In the second set of experiments, we employed four CPTs at different temperatures of cold water (1C, 5C, 10C, and 15C) to evoke variable pain levels. We observed that as the temperature of cold water decreases, the pain threshold decreases and maximum pain score reported during the immersion increases. Also, pain tolerance at 1C was significantly lower than tolerance at higher temperatures. Significant correlation was revealed between pain tolerance at 1C and change in Hb response.

Right hemisphere dominance in the hemodynamic response to cold water stimuli was again observed in these experiments. No laterality of the response was seen at the skin.

We also observed important gender differences in the hemodynamic response recorded by channels that scan the surface of prefrontal cortex. Gender-specific response was seen at the right far channel suggesting that right prefrontal cortex may mediate gender differences in the neurophysiological response to pain. Gender difference was not found in the skin hemodynamic activity, providing further support for the specificity of fNIRS technique.

We also employed a novel data analysis technique, namely functional data analysis

(fDA), to explore the variability in our measurements. fDA converts discrete observations over a continuum to continuous curves. It provides computational efficiency and flexibility and allows study of the dynamics of processes. We used functional principal component analysis (fPCA) and functional canonical correlation analysis (fCCA) to investigate the variability between the hemodynamic curves during a CPT and examine shared modes of variation between Hb and HbO2 curves.

We showed that fPCA could effectively discriminate different experimental conditions and characterize main features shared by subjects. It also quantifies the weight of each 110 feature that could be helpful in including the most significant features for statistical analysis. This information cannot be easily obtained from the original observations. For example, we found that not only the peak of HbO2 response is an important feature but also, the slope of the evoked response is highly variable between subjects and should be included in future analyses. We also investigated the interaction between Hb and HbO2 curves by fCCA and found an overall strong correlation between Hb and HbO2 variations through a CPT.

5.2. Future work fNIRS application for the assessment of pain is very recent but literature shows a fast growing interest in such a novel solution. Increasing refinement of the techniques proposed in this dissertation would be necessary to make fNIRS a clinical tool for the assessment of chronic pain condition. These refinements include:

• Since the hemodynamic response to cold water immersion is mainly confounded by

the ANS response, using other noxious stimuli that would evoke a less generalized

systemic response such as hot plates and pressure algometry would help in

discriminating the central processing and autonomic reactivity.

• Advanced signal processing techniques needs to be incorporated to separate the

hemodynamic response of different tissue layers. In particular, isolating the cortical

response may help in developing neurofeedback settings. In such settings, pain

patients learn to regulate their brain activity to soothe their pain. This framework aims

designing personalized treatments through the use of an individual‟s own data for

self-regulation. 111

• Novel algorithms for functional data analysis (in contrast to feature analysis) of pain

data with a focus on inter-subjects variability is required to better understand the

central (central nervous system) and general (autonomic nervous system) components

of the hemodynamic response to pain.

• Other modalities, such as EEG and fMRI, should be integrated to complement fNIRS

by improving the temporal and spatial resolution of measurements.

5.3. Clinical Applications

This PhD research focused on the study of acute pain response in healthy individuals between 18 and 40 years old with no pre-existing pain condition. Our sample subjects were fully communicative and able to give their report of pain. The association between self-report of pain and the hemodynamic parameters found in this research would particularly benefit patient populations for whom communication for self-report is primitive such as infants or impaired such as elders with dementia or post-stroke complications. There have been several researches studying infant‟s hemodynamic response to noxious stimuli. Future studies should include elders with impaired cognition.

An important practical consideration for elders is the use of hot water immersion rather than CPT because most seniors suffer from arthritis in their extremities. The hot water immersion test is equally capable of triggering a pain response without the confounding effect of seen in a CPT.

Another possible application of the proposed technique is the assessment of opioid- induced (OIH), which is heightened sensitivity to pain after prolonged exposure to opioid. OIH limits the efficacy of pain medications. CPT has been long used 112 to assess the sensitivity to opioid administration. The affordability and ease of fNIRS use can provide an objective tool for longitudinal monitoring of OIH development.

Last but not least, we suggest the application of the proposed technique for the assessment of preoperative sensitivity to cold water stimulus. Several studies have shown that sensitivity to cold pain is a risk factor for developing post-operative pain. Our technique can effectively and objectively examine pre- and post-surgery sensitivity to pain.

113

List of References

1. "Classification of chronic pain. Descriptions of chronic pain syndromes and definitions of pain terms. Prepared by the International Association for the Study of Pain, Subcommittee on Taxonomy," Pain. Suppl. 3,S1-226 (1986)

2. L. S. Simon, "Relieving pain in America: A blueprint for transforming prevention, care, education, and research," Journal of pain & palliative care pharmacotherapy 26(2), 197-198 (2012)

3. "American Academy of Pain Medicine - Get the facts on pain."

4. M. Ranger and C. Gélinas, "Innovating in pain assessment of the critically ill: exploring cerebral near-infrared spectroscopy as a bedside approach," Nursing 0, (2012)

5. K. L. Casey, S. Minoshima, T. J. Morrow and R. A. Koeppe, "Comparison of human cerebral activation pattern during cutaneous warmth, heat pain, and deep cold pain," J. Neurophysiol. 76(1), 571-581 (1996)

6. E. Disbrow, M. Buonocore, J. Antognini, E. Carstens and H. A. Rowley, "Somatosensory cortex: a comparison of the response to noxious thermal, mechanical, and electrical stimuli using functional magnetic resonance imaging," Hum. Brain Mapp. 6(3), 150-159 (1998)

7. C. A. Porro, V. Cettolo, M. P. Francescato and P. Baraldi, "Temporal and intensity coding of pain in human cortex," J. Neurophysiol. 80(6), 3312-3320 (1998)

8. F. Seifert and C. Maihofner, "Representation of cold allodynia in the human brain--a functional MRI study," Neuroimage 35(3), 1168-1180 (2007)

9. I. Tracey, L. Becerra, I. Chang, H. Breiter, L. Jenkins, D. Borsook and R. G. Gonzalez, "Noxious hot and cold stimulation produce common patterns of brain activation in humans: a functional magnetic resonance imaging study," Neurosci. Lett. 288(2), 159-162 (2000)

10. G. Wasner, J. Schattschneider, A. Binder and R. Baron, "Complex regional pain syndrome--diagnostic, mechanisms, CNS involvement and therapy," Spinal Cord 41(2), 61-75 (2003)

11. W. Janig and R. Baron, "Complex regional pain syndrome: mystery explained?," Lancet Neurol 2(11), 687-697 (2003)

12. E. A. Hines and G. E. Brown, "A standard stimulus for measuring vasomotor reactions: Its application in the study of hypertension," Mayo Clin. Proc. 7,332-335 (1932)

13. S. a. J. D. H. Wolf, "Studies on pain. Observations on pain due to local cooling and on factors involved in the "cold pressor" effect," J. Clin. Invest. 20(5), 521–533 (1941) 114

14. R. B. Panerai, S. L. Dawson, P. J. Eames and J. F. Potter, "Cerebral blood flow velocity response to induced and spontaneous sudden changes in arterial blood pressure," Am J Physiol Heart Circ Physiol 280(5), H2162-2174 (2001)

15. S. Roatta, G. Micieli, D. Bosone, G. Losano, R. Bini, A. Cavallini and M. Passatore, "Effect of generalised sympathetic activation by cold pressor test on cerebral haemodynamics in healthy humans," J. Auton. Nerv. Syst. 71(2-3), 159-166 (1998)

16. M. Music, Z. Finderle and K. Cankar, "Cold perception and cutaneous microvascular response to local cooling at different cooling temperatures," Microvasc. Res. 81(3), 319-324 (2011)

17. A. M. Seifalian, G. Stansby, A. Jackson, K. Howell and G. Hamilton, "Comparison of laser Doppler perfusion imaging, laser Doppler flowmetry, and thermographic imaging for assessment of blood flow in human skin," Eur. J. Vasc. Surg. 8(1), 65-69 (1994)

18. I. D. Swain and L. J. Grant, "Methods of measuring skin blood flow," Phys. Med. Biol. 34(2), 151-175 (1989)

19. P. Rolfe, "In vivo near-infrared spectroscopy," Annual Review of Biomedical Engineering 2,715-754 (2000)

20. M. Bartocci, L. L. Bergqvist, H. Lagercrantz and K. J. S. Anand, "Pain activates cortical areas in the preterm newborn brain," PAIN 122(1–2), 109-117 (2006)

21. R. Slater, S. Boyd, J. Meek and M. Fitzgerald, "Cortical pain responses in the infant brain," PAIN 123(3), 332 (2006)

22. L. Becerra, W. Harris, D. Joseph, T. Huppert, D. A. Boas and D. Borsook, "Diffuse optical tomography of pain and tactile stimulation: activation in cortical sensory and emotional systems," Neuroimage 41(2), 252-259 (2008)

23. L. Becerra, W. Harris, M. Grant, E. George, D. Boas and D. Borsook, "Diffuse optical tomography activation in the somatosensory cortex: specific activation by painful vs. non-painful thermal stimuli," Plos One 4(11), (2009)

24. C. Gelinas, M. Choiniere, M. Ranger, A. Denault, A. Deschamps and C. Johnston, "Toward a new approach for the detection of pain in adult patients undergoing cardiac surgery: Near-infrared spectroscopy-A pilot study," Heart & Lung 39(6), 485-493 (2010)

25. Y. Watanabe, H. Tanaka, I. Dan, K. Sakurai, K. Kimoto, R. Takashima and K. Hirata, "Monitoring cortical hemodynamic changes after sumatriptan injection during migraine attack by near-infrared spectroscopy," Neuroscience Research 69(1), 60-66 (2011)

26. E. Okada, M. Firbank, M. Schweiger, S. R. Arridge, M. Cope and D. T. Delpy, "Theoretical and experimental investigation of near-infrared light propagation in a model of the adult head," Appl Opt 36(1), 21-31 (1997)

27. S. Wolf and J. D. Hardy, "Studies on pain. Observations on pain due to local cooling and on factors involved in the "cold pressor" effect," J. Clin. Invest. 20(5), 521-533 (1941) 115

28. J. R. Grothusen and R. J. Schwartzman, "Laser Doppler imaging: usefulness in chronic pain medicine," Pain Physician 15(5), 491-498 (2011)

29. V. Di Piero, S. Ferracuti, U. Sabatini, P. Pantano, G. Cruccu and G. L. Lenzi, "A cerebral blood flow study on tonic pain activation in man," Pain 56,167-173 (1994)

30. K. L. Casey, S. Minoshima, T. J. Morrow and R. A. Koeppe, "Comparison of human cerebral activation patterns during cutaneous warmth, heat pain, and deep cold pain," J. Neurophysiol. 76(1), 571-581 (1996)

31. M. Bartocci, L. L. Bergqvist, H. Lagercrantz and K. J. Anand, "Pain activates cortical areas in the preterm newborn brain," Pain 122(1-2), 109-117 (2006)

32. L. Becerra, W. Harris, M. Grant, E. George, D. Boas and D. Borsook, "Diffuse optical tomography activation in the somatosensory cortex: specific activation by painful vs. non-painful thermal stimuli," PLoS One 4(11), e8016 (2009)

33. C. Gelinas, M. Choiniere, M. Ranger, A. Denault, A. Deschamps and C. Johnston, "Toward a new approach for the detection of pain in adult patients undergoing cardiac surgery: near-infrared spectroscopy--a pilot study," Heart Lung 39(6), 485-493 (2010)

34. Y. Watanabe, H. Tanaka, I. Dan, K. Sakurai, K. Kimoto, R. Takashima and K. Hirata, "Monitoring cortical hemodynamic changes after sumatriptan injection during migraine attack by near-infrared spectroscopy," Neurosci. Res. 69(1), 60-66 (2011)

35. W. Lovallo, "The cold pressor test and autonomic function: a review and integration," Psychophysiology 12(3), 268-282 (1975)

36. A. Streff, L. K. Kuehl, G. Michaux and F. Anton, "Differential physiological effects during tonic painful hand immersion tests using hot and ice water," Eur. J. Pain 14(266-272 (2010)

37. N. E. Walsh, L. Schoenfeld, S. Ramamurthy and J. Hoffman, "Normative model for cold pressor test," Am. J. Phys. Med. 68(1), 6-11 (1989)

38. B. Chance, E. Anday, S. Nioka, S. Zhou, L. Hong, K. Worden, C. Li, T. Murray, Y. Ovetsky, D. Pidikiti and R. Thomas, "A novel method for fast imaging of brain function, non- invasively, with light," Opt. Express. 2(10), 411-423 (1998)

39. M. Izzetoglu, K. Izzetoglu, S. C. Bunce, H. Ayaz, A. Devaraj, B. Onaral and K. Pourrezaei, "Functional Near-Infrared Neuroimaging," IEEE Trans. Neural Syst. and Rehabil. Eng. 13(2), 153-159 (2005)

40. A. Villringer and B. Chance, "Non-invasive optical spectroscopy and imaging of human brain function," Trends Neurosci. 20,435–442 (1997)

41. M. Cope, D. T. Delpy, E. O. Reynolds, J. Wray and P. van der Zee, "Methods of quantitating cerebral near infrared spectroscopy data," Adv. Exp. Med. Biol. 222,183-189 (1988) 116

42. H. Ayaz, P. A. Shewokis, A. Curtin, M. Izzetoglu, K. Izzetoglu and B. Onaral, "Using MazeSuite and functional near infrared spectroscopy to study learning in spatial navigation," J Vis Exp 56, (2011)

43. A. Bozkurt and B. Onaral, "Safety assessment of near infrared light emitting diodes for diffuse optical measurements," Biomed. Eng. Online 3(9), (2004)

44. A. Bozkurt, A. Rosen, H. Rosen and B. Onaral, "A portable near infrared spectroscopy system for bedside monitoring of newborn brain," Biomed. Eng. Online 4(29), (2005)

45. S. C. Bunce, M. Izzetoglu, K. Izzetoglu, B. Onaral and K. Pourrezaei, "Functional near- infrared spectroscopy an emerging neuroimaging modality," IEEE Eng. Med. Biol. 25(4), 54-62 (2006)

46. H. Dehghani and D. T. Delpy, "Near-infrared spectroscopy of the adult head: effect of scattering and absorbing obstructions in the cerebrospinal fluid layer on light distribution in the tissue," Appl. Opt 39(25), 4721-4729 (2000)

47. R. Saager and A. Berger, "Measurement of layer-like hemodynamic trends in scalp and cortex: implications for physiological baseline suppression in functional near-infrared spectroscopy," J Biomed Opt 13(3), 034017 (2008)

48. E. M. Sevick, C. L. Burch and B. Chance, "Near-infrared optical imaging of tissue phantoms with measurement in the change of optical path lengths," Adv. Exp. Med. Biol. 345,815-823 (1994)

49. R. Oldfield, "The assessment and analysis of handedness: The Edinburgh inventory," Neuropsychologia 9(1), 97-113 (1971)

50. C. L. von Baeyer, T. Piira, C. T. Chambers, M. Trapanotto and L. K. Zeltzer, "Guidelines for the cold pressor task as an experimental pain stimulus for use with children," The journal of pain : official journal of the American Pain Society 6(4), 218-227 (2005)

51. W. W. Downie, P. A. Leatham, V. M. Rhind, V. Wright, J. A. Branco and J. A. Anderson, "Studies with pain rating scales," Ann. Rheum. Dis. 37,378-381 (1978)

52. A. Devaraj, "Signal Processing for Functional Near Infrared Neuroimaging," Drexel University (2005).

53. J. O. Ramsay, G. Hooker and S. Graves, Functional Data Analysis with R and MATLAB, Springer (2009).

54. Y. Benjamini and Y. Hochberg, "Controlling the false rate discovery: a practical and powerful approach to multiple testing," J. R. Stat. Soc. B. 57(1), 289-300 (1995)

55. J. M. Bland and D. G. Altman, "Calculating correlation coefficients with repeated observations: Part 1--Correlation within subjects," BMJ 310(6977), 446 (1995)

56. R. L. Buckner, "Event-related fMRI and the hemodynamic response," Hum. Brain Mapp. 6,373–377 (1998) 117

57. G. M. Boynton, S. A. Engel, G. H. Glover and D. J. Heeger, "Linear systems analysis of functional magnetic resonance imaging in human V1," J. Neurosci. 16(13), 4207-4221 (1991)

58. M. Tanida, K. Sakatani, R. Takano and K. Tagai, "Relation between asymmetry of prefrontal cortex activities and the autonomic nervous system during a mental arithmetic task: near infrared spectroscopy study," Neurosci. Lett. 369(1), 69-74 (2004)

59. A. Stancak, Jr., A. Yamamotova, I. P. Kulls and I. V. Sekyra, "Cardiovascular adjustments and pain during repeated cold pressor test," Clin. Auton. Res. 6(2), 83-89 (1996)

60. L. L. Washington, S. J. Gibson and R. D. Helme, "Age-related differences in the endogenous analgesic response to repeated cold water immersion in human volunteers," Pain 89(1), 89-96 (2000)

61. P. Rolfe, "In vivo near-infrared spectroscopy," Annu Rev Biomed Eng 2,715-754 (2000)

62. A. Villringer and B. Chance, "Non-invasive optical spectroscopy and imaging of human brain function," Trends Neurosci. 20(10), 435-442 (1997)

63. M. Ferrari and V. Quaresima, "A brief review on the history of human functional near- infrared spectroscopy (fNIRS) development and fields of application," Neuroimage 63(2), 921- 935 (2012)

64. H. Dehghani and D. T. Delpy, "Near-infrared spectroscopy of the adult head: effect of scattering and absorbing obstructions in the cerebrospinal fluid layer on light distribution in the tissue," Applied optics 39(25), 4721-4729 (2000)

65. K. Taniguchi, S. Sumitani, Y. Watanabe, M. Akiyama and T. Ohmori, "Multi-channel near-infrared spectroscopy reveals reduced prefrontal activation in schizophrenia patients during performance of the kana Stroop task," J. Med. Invest. 59(1-2), 45-52 (2012)

66. S. Shimodera, Y. Imai, N. Kamimura, I. Morokuma, H. Fujita, S. Inoue and T. A. Furukawa, "Mapping hypofrontality during letter fluency task in schizophrenia: a multi-channel near-infrared spectroscopy study," Schizophr. Res. 136(1-3), 63-69 (2012)

67. S. Koike, R. Takizawa, Y. Nishimura, K. Marumo, M. Kinou, Y. Kawakubo, M. A. Rogers and K. Kasai, "Association between severe dorsolateral prefrontal dysfunction during random number generation and earlier onset in schizophrenia," Clin. Neurophysiol. 122(8), 1533- 1540 (2011)

68. Y. Fujita, M. Takebayashi, K. Hisaoka, M. Tsuchioka, S. Morinobu and S. Yamawaki, "Asymmetric alternation of the hemodynamic response at the prefrontal cortex in patients with schizophrenia during electroconvulsive therapy: a near-infrared spectroscopy study," Brain Res. 1410,132-140 (2011)

69. M. Azechi, M. Iwase, K. Ikezawa, H. Takahashi, L. Canuet, R. Kurimoto, T. Nakahachi, R. Ishii, M. Fukumoto, K. Ohi, Y. Yasuda, H. Kazui, R. Hashimoto and M. Takeda, "Discriminant analysis in schizophrenia and healthy subjects using prefrontal activation during frontal lobe tasks: a near-infrared spectroscopy," Schizophr. Res. 117(1), 52-60 (2010) 118

70. K. Ikezawa, M. Iwase, R. Ishii, M. Azechi, L. Canuet, K. Ohi, Y. Yasuda, N. Iike, R. Kurimoto, H. Takahashi, T. Nakahachi, R. Sekiyama, T. Yoshida, H. Kazui, R. Hashimoto and M. Takeda, "Impaired regional hemodynamic response in schizophrenia during multiple prefrontal activation tasks: a two-channel near-infrared spectroscopy study," Schizophr. Res. 108(1-3), 93-103 (2009)

71. A. Watanabe and T. Kato, "Cerebrovascular response to cognitive tasks in patients with schizophrenia measured by near-infrared spectroscopy," Schizophr. Bull. 30(2), 435-444 (2004)

72. F. Okada, Y. Tokumitsu, Y. Hoshi and M. Tamura, "Impaired interhemispheric integration in brain oxygenation and hemodynamics in schizophrenia," Eur. Arch. Psychiatry Clin. Neurosci. 244(1), 17-25 (1994)

73. A. J. Fallgatter, M. Roesler, L. Sitzmann, A. Heidrich, T. J. Mueller and W. K. Strik, "Loss of functional hemispheric asymmetry in Alzheimer's dementia assessed with near-infrared spectroscopy," Brain Res. Cogn. Brain Res. 6(1), 67-72 (1997)

74. C. Hock, K. Villringer, F. Muller-Spahn, M. Hofmann, S. Schuh-Hofer, H. Heekeren, R. Wenzel, U. Dirnagl and A. Villringer, "Near infrared spectroscopy in the diagnosis of Alzheimer's disease," Ann. N. Y. Acad. Sci. 777,22-29 (1996)

75. E. Watanabe, Y. Nagahori and Y. Mayanagi, "Focus diagnosis of epilepsy using near- infrared spectroscopy," Epilepsia 43 Suppl 9,50-55 (2002)

76. K. Haginoya, M. Munakata, R. Kato, H. Yokoyama, M. Ishizuka and K. Iinuma, "Ictal cerebral haemodynamics of childhood epilepsy measured with near-infrared spectrophotometry," Brain 125(Pt 9), 1960-1971 (2002)

77. E. Watanabe, A. Maki, F. Kawaguchi, Y. Yamashita, H. Koizumi and Y. Mayanagi, "Noninvasive cerebral blood volume measurement during seizures using multichannel near infrared spectroscopic topography," J Biomed Opt 5(3), 287-290 (2000)

78. D. K. Sokol, O. N. Markand, E. C. Daly, T. G. Luerssen and M. D. Malkoff, "Near infrared spectroscopy (NIRS) distinguishes seizure types," Seizure 9(5), 323-327 (2000)

79. R. Bokiniec, A. Zbiec, J. Seliga, P. Sawosz, A. Liebert, I. Klosinska, A. Bargiel, R. Maniewski and M. Kornacka, "Assessment of brain oxygenation in term and preterm neonates using near infrared spectroscopy," Advances in medical sciences 57(2), 348-355 (2012)

80. M. Ranger, C. C. Johnston, C. Limperopoulos, J. E. Rennick and A. J. du Plessis, "Cerebral near-infrared spectroscopy as a measure of nociceptive evoked activity in critically ill infants," Pain Res Manag 16(5), 331-336 (2011)

81. P. Giliberti, V. Mondi, A. Conforti, M. Haywood Lombardi, S. Sgro, P. Bozza, S. Picardo, A. Dotta and P. Bagolan, "Near infrared spectroscopy in newborns with surgical disease," J Matern Fetal Neonatal Med 24 Suppl 1,56-58 (2011)

82. G. E. Moreno, M. L. Pilan, C. Manara, R. Magliola, J. C. Vassallo, M. Balestrini, A. M. Lenz, M. Krynski, M. Althabe and L. Landry, "Regional venous oxygen saturation versus mixed venous saturation after paediatric cardiac surgery," Acta Anaesthesiol. Scand. 57(3), 373-379 (2013) 119

83. A. Milan, F. Freato, V. Vanzo, G. Camporese, M. Baraldi, L. Chiandetti and P. Zaramella, "Near-infrared spectroscopy measure of limb peripheral perfusion in neonatal arterial thromboembolic disease," Minerva Pediatr. 64(6), 633-639 (2012)

84. R. M. Cerbo, R. Cabano, A. Di Comite, S. Longo, R. Maragliano and M. Stronati, "Cerebral and somatic rSO2 in sick preterm infants," J Matern Fetal Neonatal Med 25 Suppl 4,97-100 (2012)

85. G. Greisen, T. Leung and M. Wolf, "Has the time come to use near-infrared spectroscopy as a routine clinical tool in preterm infants undergoing intensive care?," Philos Transact A Math Phys Eng Sci 369(1955), 4440-4451 (2011)

86. W. Baerts, P. M. Lemmers and F. van Bel, "Cerebral oxygenation and oxygen extraction in the preterm infant during desaturation: effects of increasing FiO(2) to assist recovery," Neonatology 99(1), 65-72 (2011)

87. A. Petrova and R. Mehta, "Near-infrared spectroscopy in the detection of regional tissue oxygenation during hypoxic events in preterm infants undergoing critical care," Pediatr Crit Care Med 7(5), 449-454 (2006)

88. T. J. Huppert, S. G. Diamond, M. A. Franceschini and D. A. Boas, "HomER: a review of time-series analysis methods for near-infrared spectroscopy of the brain," Appl Opt 48(10), D280- 298 (2009)

89. K. J. Friston, P. Jezzard and R. Turner, "Analysis of functional MRI time-series," Hum. Brain Mapp. 1,153-171 (1994)

90. Y. Zhang, D. H. Brooks and D. A. Boas, "A haemodynamic response function model in spatio-temporal diffuse optical tomography," Phys. Med. Biol. 50(19), 4625-4644 (2005)

91. J. O. Ramsay and B. W. Silverman, Functional Data Analysis, Springer (2005).

92. I. M. Johnstone and D. M. Titterington, "Statistical challenges of high-dimensional data," Phil. Trans. R. Soc. A 367(1906), 4237-4253 (2009)

93. T. S. Tian, "Functional data analysis in brain imaging studies," Front Psychol 1(35 (2010)

94. R. Viviani, G. Gron and M. Spitzer, "Functional principal component analysis of fMRI data," Hum. Brain Mapp. 24(2), 109-129 (2005)

95. C. J. Long, E. N. Brown, C. Triantafyllou, I. Aharon, L. L. Wald and V. Solo, "Nonstationary noise estimation in functional MRI," Neuroimage 28(4), 890-903 (2005)

96. Z. Barati, P. A. Shewokis, M. Izzetoglu, R. Polikar, G. Mychaskiw and K. Pourrezaei, "Hemodynamic response to repeated noxious cold pressor tests measured by functional near infrared spectroscopy on forehead," Ann. Biomed. Eng. 41(2), 223-237 (2013)

97. B. Chance, E. Anday, S. Nioka, S. Zhou, L. Hong, K. Worden, C. Li, T. Murray, Y. Ovetsky, D. Pidikiti and R. Thomas, "A novel method for fast imaging of brain function, non- invasively, with light," Opt Express 2(10), 411-423 (1998) 120

98. G. Wahba and S. Wold, "A completely automatic french curve: fitting spline functions by cross validation," Communications in Statistics 4(1), 17 (1975)

99. B. W. Silverman, "Some Aspects of the Spline Smoothing Approach to Non-Parametric Regression Curve Fitting," J. R. Statist. Soc. B 47(1), 52 (1985)

100. R. L. Eubank, Spline smoothing and nonparametric regression, M. Dekker (1988).

101. P. Craven and G. Wahba, "Smoothing noisy data with spline functions," Numer. Math. 31(4), 27 (1979)

102. S. E. Leurgans, R. A. Moyeed and B. W. Silverman, "Canonical correlation analysis when the data are curves," J. R. Statist. Soc. B 55(3), 16 (1993)

103. Y. Zhang, D. H. Brooks, M. A. Franceschini and D. A. Boas, "Eigenvector-based spatial filtering for reduction of physiological interference in diffuse optical imaging," J Biomed Opt 10(1), 11014 (2005)

104. J. Virtanen, T. Noponen and P. Merilainen, "Comparison of principal and independent component analysis in removing extracerebral interference from near-infrared spectroscopy signals," J Biomed Opt 14(5), 054032 (2009)

105. I. Sendowski, G. Savourey, Y. Besnard and J. Bittel, "Cold induced vasodilatation and cardiovascular responses in humans during cold water immersion of various upper limb areas," Eur J Appl Physiol Occup Physiol 75(6), 471-477 (1997)

106. U. N. Frankenstein, W. Richter, M. C. McIntyre and F. Remy, "Distraction modulates anterior cingulate gyrus activations during the cold pressor test," Neuroimage 14(4), 827-836 (2001)

107. B. Hellstrom and U. Lundberg, "Pain perception to the cold pressor test during the menstrual cycle in relation to estrogen levels and a comparison with men," Integr. Physiol. Behav. Sci. 35(2), 132-141 (2000)

108. J. Tesarz, A. Gerhardt, K. Schommer, R. D. Treede and W. Eich, "Alterations in endogenous pain modulation in endurance athletes: An experimental study using quantitative sensory testing and the cold-pressor task," Pain 154(7), 1022-1029 (2013)

109. P. Rainville, J. S. Feine, M. C. Bushnell and G. H. Duncan, "A psychophysical comparison of sensory and affective responses to four modalities of experimental pain," Somatosens. Mot. Res. 9(4), 265-277 (1992)

110. A. M. Unruh, "Gender variations in clinical pain experience," Pain 65(2-3), 123-167 (1996)

111. C. D. Myers, J. L. Riley, 3rd and M. E. Robinson, "Psychosocial contributions to sex- correlated differences in pain," Clin. J. Pain 19(4), 225-232 (2003)

112. C. A. Flores, X. M. Wang, K. M. Zhang and S. S. Mokha, "Orphanin FQ produces gender-specific modulation of trigeminal nociception: behavioral and electrophysiological observations," Neuroscience 105(2), 489-498 (2001) 121

113. P. E. Paulson, S. Minoshima, T. J. Morrow and K. L. Casey, "Gender differences in pain perception and patterns of cerebral activation during noxious heat stimulation in humans," Pain 76(1-2), 223-229 (1998)

114. T. Bisgaard, B. Klarskov, J. Rosenberg and H. Kehlet, "Characteristics and prediction of early pain after laparoscopic cholecystectomy," Pain 90(3), 261-269 (2001)

115. B. T. West, K. B. Welch and A. T. Galecki, Linear Mixed Models: A Practical Guide Using Statistical Software, Chapman and Hall/CRC (2006).

116. P. Diggle, P. Heagerty, K. Liang and S. Zeger, Analysis of longitudinal data (Oxford Statistical Science Series), Oxford University Press, USA (2013).

117. P. Diggle, "An approach to the analysis of repeated measurements," Biometrics 44(959- 971 (1988)

118. V. Di Piero, S. Ferracuti, U. Sabatini, P. Pantano, G. Cruccu and G. L. Lenzi, "A cerebral blood flow study on tonic pain activation in man," Pain 56(2), 167-173 (1994)

119. K. L. Casey, P. Svensson, T. J. Morrow, J. Raz, C. Jone and S. Minoshima, "Selective opiate modulation of nociceptive processing in the human brain," J. Neurophysiol. 84(1), 525-533 (2000)

120. R. B. Fillingim, C. D. King, M. C. Ribeiro-Dasilva, B. Rahim-Williams and J. L. Riley, 3rd, "Sex, gender, and pain: a review of recent clinical and experimental findings," The journal of pain : official journal of the American Pain Society 10(5), 447-485 (2009)

121. M. Racine, Y. Tousignant-Laflamme, L. A. Kloda, D. Dion, G. Dupuis and M. Choiniere, "A systematic literature review of 10 years of research on sex/gender and experimental pain perception - part 1: are there really differences between women and men?," Pain 153(3), 602-618 (2012)

122. J. L. Riley, 3rd, M. E. Robinson, E. A. Wise, C. D. Myers and R. B. Fillingim, "Sex differences in the perception of noxious experimental stimuli: a meta-analysis," Pain 74(2-3), 181-187 (1998)

123. S. M. Berman, B. D. Naliboff, B. Suyenobu, J. S. Labus, J. Stains, J. A. Bueller, K. Ruby and E. A. Mayer, "Sex differences in regional brain response to aversive pelvic visceral stimuli," American journal of physiology. Regulatory, integrative and comparative physiology 291(2), R268-276 (2006)

124. L. A. Henderson, S. C. Gandevia and V. G. Macefield, "Gender differences in brain activity evoked by muscle and cutaneous pain: a retrospective study of single-trial fMRI data," Neuroimage 39(4), 1867-1876 (2008)

125. S. W. Derbyshire, T. E. Nichols, L. Firestone, D. W. Townsend and A. K. Jones, "Gender differences in patterns of cerebral activation during equal experience of painful laser stimulation," The journal of pain : official journal of the American Pain Society 3(5), 401-411 (2002) 122

126. A. R. Hobson, P. L. Furlong, S. F. Worthen, A. Hillebrand, G. R. Barnes, K. D. Singh and Q. Aziz, "Real-time imaging of human cortical activity evoked by painful esophageal stimulation," Gastroenterology 128(3), 610-619 (2005)

127. E. A. Moulton, M. L. Keaser, R. P. Gullapalli, R. Maitra and J. D. Greenspan, "Sex differences in the cerebral BOLD signal response to painful heat stimuli," American journal of physiology. Regulatory, integrative and comparative physiology 291(2), R257-267 (2006)

128. B. D. Naliboff, S. Berman, L. Chang, S. W. Derbyshire, B. Suyenobu, B. A. Vogt, M. Mandelkern and E. A. Mayer, "Sex-related differences in IBS patients: central processing of visceral stimuli," Gastroenterology 124(7), 1738-1747 (2003)

129. T. Straube, S. Schmidt, T. Weiss, H. J. Mentzel and W. H. Miltner, "Sex differences in brain activation to anticipated and experienced pain in the medial prefrontal cortex," Hum. Brain Mapp. 30(2), 689-698 (2009)

130. R. M. Harper, R. Bandler, D. Spriggs and J. R. Alger, "Lateralized and widespread brain activation during transient blood pressure elevation revealed by magnetic resonance imaging," J. Comp. Neurol. 417(2), 195-204 (2000)

123

VITA ZEINAB BARATI

EDUCATION Drexel University, Philadelphia, PA, USA Doctoral degree in Biomedical Engineering – Winter 2009 to Summer 2013

Iran University of Science and Technology, Tehran, Iran Master of Science in Biomedical Engineering – Fall 2003 to Summer 2006

Amirkabir University of Technology, Tehran, Iran Bachelor of Science in Biomedical Engineering – Fall 1998 to Fall 2002

PROFESSIONAL Siemens representative in Iran (Matlab Kish Co.), Tehran, Iran EXPERIENCE Sales Engineer at the Department of Molecular Imaging and Oncology Care Systems - April 2007 to November 2008

Chamran Hospital, Tehran, Iran Intern at the Department of Biomedical Engineering – Summer 2000

TEACHING Bioelectronics Lab - Winter and Summer 2012 EXPERIENCE Medical Device Development - Spring 2012 Intermediate Biostatistics - Winter 2011 Biomedical Systems and Signal Processing - Fall 2009 and 2011

PUBLICATIONS  Zeinab Barati, Patricia A. Shewokis, Meltem Izzetoglu, Robi Polikar, George Mychaskiw, and Kambiz Pourrezaei. “Hemodynamic response to repeated noxious cold pressor tests: a near infrared spectroscopy study on forehead”, Annals of Biomedical Engineering 41(2), 223-237 (2013)  Zeinab Barati, Issa Zakeri, and Kambiz Pourrezaei. “Functional Data Analysis View of fNIRS Data”, submitted to the Journal of Biomedical Optics  Zeinab Barati and Ahmad Ayatollahi, “Baseline wandering removal by using independent component analysis to single-channel ECG data”, proceedings of IEEE International Conference on Biomedical and Pharmaceutical Engineering, Singapore, December 2006

AWARDS  Travel award from the American Pain Society (2013)  Fellowship from the American Pain Society for “Fundamentals of Pain Management: A Primer for Residents and Fellows ©” (2011)  Travel award from the International Association for the Study of Pain (2010)  Teaching Assistantship from the School of Biomedical Engineering, Science, and Health Systems, Drexel University (2009 to 2012)  Research Assistantship from CONQUER CollabOrative lab, School of Biomedical Engineering, Science, and Health Systems, Drexel University (2009 to 2012)  Calhoun Fellowship from the School of Biomedical Engineering, Science, and Health Systems, Drexel University (2009 and 2010)