UNIVERSITY OF CALGARY

publicité
UNIVERSITY OF CALGARY
Exploring the Relationships between Yoga Practice, Affect and Attention Regulation,
Health Outcomes and Program Adherence in Cancer Survivors
by
Michael John Mackenzie
A DISSERTATION
SUBMITTED TO THE FACULTY OF GRADUATE STUDIES
IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE
DEGREE OF DOCTOR OF PHILOSOPHY
FACULTY OF KINESIOLOGY
CALGARY, ALBERTA
NOVEMBER, 2012
© Michael John Mackenzie 2012
Abstract
Yoga practice has been reported to lead to improvements in health-related quality
of life (HRQL), psychological functioning, and symptom indices in cancer survivors.
Yoga is defined in the Yoga Sutras of Patanjali as a path towards, “stilling the
fluctuations of the mind (YS I.2).” Importantly, meditative states experienced within yoga
practice are correlated to neurophysiological systems that moderate both positive affect
and internalised attention. Despite preliminary findings, little attention has been paid to
the psychophysiological mechanisms by which benefits are accrued via yoga practice.
The impetus for the present studies was to explore the mechanisms by which yoga
practice stills the fluctuations of the mind in cancer survivors. Three studies examined: 1)
the clinical significance of patient–reported outcomes in yoga interventions for cancer
survivors; 2) longitudinal associations between yoga participation, affect, and
mindfulness as predictors of mood, stress and HRQL outcomes in an ongoing
community-based yoga program; 3) associations between valence, activation, attention,
perceived exertion, cardiac activity, and participant descriptions of a single yoga session
for cancer survivors.
The clinical significance review confirmed yoga practice was related to clinically
significant improvements in HRQL, psychological health, and symptom experience.
Longitudinal findings suggest improvements in mood, stress and HRQL were related to
affect and attention (mindfulness) regulation, but not previous yoga experience. Previous
yoga experience, affect and mindfulness were related to yoga practice maintenance. Lab
study findings suggest both affect and attention improved in a single yoga session and
were related to cardiac activity. Participant descriptions confirmed these findings and
ii
further suggested regulation of attention via breath awareness elicited positive affective
responses to yoga.
This research develops current theory in yoga practice, affect and attention
regulation by modelling processes and outcomes, resulting in a detailed exploration of
yoga in cancer settings. Examining these proposed theoretically-based mechanisms for
yoga’s salutary effects enables greater understanding not only of “if” yoga works, but
also “how.” This knowledge can be used to develop innovative yoga interventions with
the express aim of improving mental health and HRQL in cancer survivors.
iii
Preface
In the course of this dissertation, Chapter 2 was prepared and published in
Evidence-Based Complementary and Alternative Medicine
(www.hindawi.com/journals/ecam/2012/642576/abs/). The idea for the manuscript was
initially conceptualised by Drs. Culos-Reed and Danhauer and Michael Mackenzie was
brought in as a collaborator. Mr. Mackenzie was subsequently responsible for completing
the literature review, methods, data analysis and results section and significantly
contributing to both the introduction and discussion. This was completed under the
guidance of Drs. Culos-Reed and Danhauer. All authors provided critical reviews of the
manuscript and contributed intellectual content. The published article was reproduced in
its entirety and included as a chapter in this document with permission from the
publisher. This is an open access article distributed under the Creative Commons
Attribution License, which permits unrestricted use, distribution, and reproduction in any
medium, provided the original work is properly cited
(creativecommons.org/licenses/by/3.0/).
Culos-Reed SN, Mackenzie MJ, Sohl SJ, Jesse MT, Ross-Zahavich AN, Danhauer SC.
Yoga & Cancer Interventions: A review of the clinical significance of patient-reported
outcomes for cancer survivors. Evidence-Based Complementary and Alternative
Medicine: eCAM 2012; 1-17.
iv
Acknowledgements
I am incredibly grateful to many people for their varied contributions and support
of this work over the past few years.
Thank you to my dissertation committee for all their work in moving this project
forward. Particular thanks go to Dr. Nicole Culos-Reed for her mentorship, patience and
gentle hand in guiding this work. Dr. Linda Carlson has been my mentor for many years
and someone I still grow with and learn from. Dr. Dave Paskevich has also been a
wonderful mentor these past three years. His thoughts and opinions have always been
welcome additions to this work. I also thank Dr. Telles and Dr. Page for taking time out
of their extremely busy schedules to be part of this process.
The Health and Wellness Lab has been my home the past three years. Thanks to
Dr. Carolina Chamarro, Lauren Capozzi, Amanda Wurz, Kathryn Wytsma, Katie Krenz,
Jessica Danyluk, Tanya Williamson, Lynette Stephenson, and Marni Armstrong for being
an integral part of this journey. Thanks also to Tyla Arnason, Anne Cox, Sophie
Dufresne, Jeanine Goranson, Catherine Townley, Heather Jackson, Helen Cheung and
Linda Crawford, our amazing Yoga Thrive teachers who have been so central to the
research process. My greatest debt is to those cancer survivors and their support persons
who chose to be a part of our research program. They have shared so much and asked for
little in return. Thank you.
Linda Carlson’s Lab has also been a place of great learning and friendship. Dr.
Laura Labelle, Dr. Sheila Garland and Kristin Zernicke have been indispensible to this
work and are always good for a fresh viewpoint and a warm beverage. Many thanks to
Linda, Laura and Sheila for reading drafts of Chapters 3 and 4.
v
In the Human Performance Lab Dr. David Smith, Dr. Neil Eves, Dr. Eric Groves,
Rosie Neil, have opened my eyes to the beauty of exercise physiology and human
performance. There is no going back now. Thank you. Many thanks to Brandon Hisey for
his artistic flair and MATLAB expertise in creating the Affect Circumplex 3D models in
Chapters 3 and 4.
To Gisela Engels, Queen of Stats and lover of basketball, go my deepest thanks. I
literally could not have done this without you.
To Panteleimon Ekkekakis, Julian Thayer and Shirley Telles, your work has
allowed for my work. Thank you.
Behind every great learning institution there are administrative assistants that
ensure everything runs smoothly. Rosalie Kolstad and Sherri Simpson are impeccable in
their work and have been indispensible in my work. Many thanks to you both.
I would be neglectful if I also didn’t mention Drs. Michael Speca, Barry Bultz and
Tavis Campbell, all long-term mentors and colleagues. Gentlemen I thank you.
To yoga friends Robin Rothenberg, Maggie Reagh, Robert Birnberg, Jan’net
Barchard-Smith, Manon Bussieres, the Desikchar family, the Krishnamacharya Yoga
Mandiram, and Gita Thalashvar’s mom, wherever you may be, thank you for my ongoing
introduction to yoga.
To friends and family, thanks for letting me disappear for three years. With any
luck, I’ll soon be back in the Land of the Living. Tim Seinen deserves an extra shout-out
as the “best man” behind the scenes. You were always up for a chat and a workout and
infinitely patient when, on more than one occasion, neither happened. Many thanks also
vi
to Scott Haggins and Cedarglen Homes for allowing me to use their fitness facilities with
Tim gratis over the past three years. It has been a pleasure.
Kristin what can I say? You’ve been a trooper for three-whole years. Thank you, I
love you and look forward to our new life in Illinois together as husband and wife. Many
thanks go to Dr. Edward McAuley and members of his Exercise Psychology Lab at the
University of Illinois in Urbana-Champaign for this exciting new opportunity as a postdoctoral fellow.
Finally, my thanks to SSHRC, AIHS and the University of Calgary for supporting
this research. It has been my greatest privilege and I thank you.
vii
Table of Contents
Abstract.............................................................................................................................. ii
Preface............................................................................................................................... iv
Acknowledgements ............................................................................................................v
Table of Contents ........................................................................................................... viii
List of Appendices .......................................................................................................... xiii
List of Tables .................................................................................................................. xiv
List of Figures...................................................................................................................xv
List of Symbols, Abbreviations and Nomenclature .................................................... xvi
Epigraph ........................................................................................................................ xvii
CHAPTER 1: INTRODUCTION .....................................................................................1
1.1 Cancer ........................................................................................................................3
1.1.1 Prevalence ...........................................................................................................3
1.1.2 Yoga and Exercise ..............................................................................................4
1.2 Yoga ...........................................................................................................................6
1.2.1 Yoga in Cancer Settings......................................................................................6
1.2.2 Yoga and Exercise ..............................................................................................7
1.2.3 What is Yoga? .....................................................................................................7
1.3. Mechanisms of Action ..............................................................................................9
1.3.1 Positive Affect ..................................................................................................10
1.3.1.1 Circumplex Model of Affect ......................................................................12
1.3.1.2 Dual Mode Theory .....................................................................................13
1.3.2 Attention ...........................................................................................................14
1.3.2.1 Effort-Related Attention Model .................................................................15
1.3.3 Mindfulness.......................................................................................................16
1.3.4 Autonomic Nervous System and Heart Rate Variability ..................................17
1.3.4.1 Neurovisceral Integration Model ..............................................................19
1.3.5 Neurophenomenology ......................................................................................20
1.4 Research Objectives .................................................................................................21
1.4.1 Study 1 (Systematic Review) ............................................................................22
1.4.2 Study 2 (Longitudinal Study)............................................................................22
1.4.3 Study 3 (Lab Study) ..........................................................................................22
1.5 Significance..............................................................................................................23
viii
CHAPTER 2: YOGA AND CANCER INTERVENTIONS: A REVIEW OF THE
CLINICAL SIGNIFICANCE OF PATIENT REPORTED OUTCOMES FOR
CANCER SURVIVORS ..................................................................................................24
2.0 Abstract ....................................................................................................................25
2.1 Introduction ..............................................................................................................26
2.2 Methods....................................................................................................................30
2.2.1 Selection of Publications...................................................................................30
2.2.2 Data Synthesis...................................................................................................31
2.3 Results ......................................................................................................................33
2.3.1 Description of Studies .......................................................................................33
2.3.1.1 Study Designs ............................................................................................33
2.3.1.2 Instruments Reviewed ................................................................................34
2.3.2 Quality of Life Outcomes .................................................................................34
2.3.2.1 Overall HRQL ............................................................................................34
2.3.2.2 Physical HRQL ..........................................................................................35
2.3.2.3 Mental HRQL ............................................................................................36
2.3.2.4 Emotional HRQL .......................................................................................36
2.3.2.5 Social HRQL ..............................................................................................37
2.3.2.6 Functional HRQL.......................................................................................37
2.3.3 Psychosocial Outcomes ....................................................................................38
2.3.3.1 Depression..................................................................................................38
2.3.3.2 Anxiety.......................................................................................................39
2.3.3.3 Positive Affect ...........................................................................................39
2.3.3.4 Negative Affect ..........................................................................................40
2.3.3.5 Spiritual Well-Being ..................................................................................40
2.3.4 Symptom Outcomes ..........................................................................................41
2.3.4.1 Fatigue........................................................................................................41
2.3.4.2 Sleep...........................................................................................................41
2.3.5 Clinical Significance Criteria ............................................................................41
2.3.5.1 Consistency between SEM and SD............................................................41
2.4 Discussion ................................................................................................................42
2.4.1 Baseline Cut-Offs Reported in Literature .........................................................44
2.4.2 MIDs Reported in Literature.............................................................................45
2.4.3 Limitations and Future Research ......................................................................45
2.4.4 Conclusion ........................................................................................................48
CHAPTER 3: MODELLING ASSOCIATIONS BETWEEN YOGA
PARTICIPATION, AFFECT, MINDFULNESS AND HEALTH OUTCOMES .....62
3.1 Introduction ..............................................................................................................63
3.1.1 Yoga in Cancer Settings....................................................................................63
3.1.2 Theoretical Mechanisms of Action ...................................................................64
3.1.2.1 Affect Regulation .......................................................................................64
3.1.2.2 Circumplex Model of Affect ......................................................................65
3.1.2.3 Dual Mode Theory .....................................................................................65
3.1.2.4 Effort-Related Attention Model .................................................................66
ix
3.1.3 Mindfulness.......................................................................................................66
3.1.3.1 Facets of Mindfulness ................................................................................67
3.2 Objectives ................................................................................................................68
3.2.1 Acute Class Effects ...........................................................................................69
3.2.2 Longitudinal Program Effects ..........................................................................69
3.2.3 Yoga Practice Maintenance ..............................................................................69
3.3 Methods....................................................................................................................70
3.3.1 Participants........................................................................................................70
3.3.2 Program .............................................................................................................70
3.3.3 Recruitment .......................................................................................................71
3.3.4 Instruments ........................................................................................................72
3.3.4.1 Baseline ......................................................................................................72
3.3.4.2 Acute Effects ..............................................................................................72
3.3.4.3 Longitudinal Effects...................................................................................73
3.3.4.4 Yoga Practice Maintenance .......................................................................75
3.3.5 Data Analysis ....................................................................................................75
3.3.5.1 Multilevel Models ......................................................................................75
3.3.5.2 Clinical Significance ..................................................................................76
3.3.5.3 Acute Effects ..............................................................................................77
3.3.5.4 Longitudinal Effects...................................................................................77
3.3.5.5 Yoga Practice Maintenance .......................................................................78
3.4 Results ......................................................................................................................78
3.4.1 Demographics ...................................................................................................79
3.4.2 Acute Class Effects ...........................................................................................79
3.4.2.1 Valence ......................................................................................................80
3.4.2.2 Activation...................................................................................................80
3.4.2.3 Attention ....................................................................................................80
3.4.3 Longitudinal Program Effects ...........................................................................81
3.4.3.1 Covariates ..................................................................................................81
3.4.3.2 Affect .........................................................................................................81
3.4.3.3 Mindfulness................................................................................................82
3.4.3.4 Mood Disturbance, Stress Symptoms and Quality of Life ........................82
3.4.3.5 Mood Disturbance......................................................................................83
3.4.3.6 Stress Symptoms ........................................................................................83
3.4.3.7 Quality of Life............................................................................................83
3.4.4 Yoga Practice Maintenance ..............................................................................84
3.5 Discussion ................................................................................................................85
3.5.1 Acute Class Effects ...........................................................................................86
3.5.2 Longitudinal Program Effects ...........................................................................88
3.5.3 Yoga Practice Maintenance ..............................................................................91
3.5.4 Limitations ........................................................................................................92
3.5.5 Conclusion ........................................................................................................93
x
CHAPTER 4: MODELLING ASSOCIATIONS BETWEEN AFFECT,
ATTENTION AND HEART RATE VARIABILITY IN A SINGLE YOGA
SESSION: A NEUROPHENOMENOLOGICAL APPROACH ..............................118
4.1 Introduction ............................................................................................................119
4.1.1 Affect ..............................................................................................................120
4.1.2 Attention .........................................................................................................120
4.1.3 Heart Rate Variability .....................................................................................122
4.1.4 Neuro-Affective Theories and Approaches ....................................................124
4.1.4.1 Circumplex Model of Affect ....................................................................124
4.1.4.2 Dual Mode Theory ...................................................................................125
4.1.4.3 Effort-Related Attention Model ...............................................................125
4.1.4.4 Neurovisceral Integration Model .............................................................126
4.1.4.5 Neurophenomenology ..............................................................................126
4.2 Objectives ..............................................................................................................127
4.3 Methods..................................................................................................................128
4.3.1 Subject Recruitment ........................................................................................128
4.3.2 Procedures .......................................................................................................128
4.3.3 Instruments – Baseline ....................................................................................130
4.3.4 Instruments – In-Task Measures .....................................................................131
4.3.5 Instruments – Pre-post Supine Meditation and Post Yoga Supine Resting ....132
4.3.6 Data Analysis ..................................................................................................132
4.3.6.1 Multilevel Modelling ...............................................................................133
4.3.6.2 Clinical Significance ................................................................................134
4.3.6.3 Qualitative Analysis .................................................................................135
4.3.6.4 Mixed Methods ........................................................................................136
4.4 Results ....................................................................................................................136
4.4.1 Demographics .................................................................................................137
4.4.2 Valence, Activation, Attention, Perceived Exertion and Cardiac Activity
During Yoga Session ...............................................................................................137
4.4.2.1 Valence ....................................................................................................137
4.4.2.2 Activation.................................................................................................138
4.4.2.3 Attention ..................................................................................................139
4.4.2.4 Rating of Perceived Exertion ...................................................................139
4.4.2.5 Cardiac Activity .......................................................................................140
4.4.3 In-Task Yoga Session Qualitative Findings ...................................................141
4.4.3.1 Description of Events...............................................................................141
4.4.3.2 Explanatory Mechanisms .........................................................................145
4.4.4 Pre-Post Yoga Session Affect Regulation ......................................................146
4.4.4.1 Energy ......................................................................................................146
4.4.4.2 Tension.....................................................................................................147
4.4.4.3 Tiredness ..................................................................................................147
4.4.4.4 Calmness ..................................................................................................147
4.4.5 Pre-Post Yoga Session Qualitative Findings ..................................................148
4.4.5.1 Energy ......................................................................................................148
4.4.5.2 Tension.....................................................................................................148
4.4.5.3 Tiredness ..................................................................................................148
xi
4.4.5.4 Calmness ..................................................................................................149
4.5 Discussion ..............................................................................................................149
4.5.1 In-Task Effects ................................................................................................150
4.5.1.1 Valence ....................................................................................................150
4.5.1.2 Activation.................................................................................................151
4.5.1.3 Attention ..................................................................................................151
4.5.2 Pre-Post Yoga Session ....................................................................................152
4.5.3 Integration .......................................................................................................152
4.5.4 Limitations ......................................................................................................154
4.5.5 Conclusion ......................................................................................................155
CHAPTER 5: GENERAL DISCUSSION ..................................................................174
5.1 The Present Studies ................................................................................................175
5.1.1 Integration of Theoretical Models, Constructs and Approaches ....................175
5.2 Study Findings .......................................................................................................176
5.2.1 Study 1 (Systematic Review, Chapter 2) ........................................................177
5.2.2 Study 2 (Longitudinal Study, Chapter 3) ........................................................177
5.2.3 Study 3 (Lab Study, Chapter 4) ......................................................................178
5.3 Theoretical Integration ...........................................................................................178
5.3.1 Acute Effects ...................................................................................................178
5.3.2 Longitudinal Measures of Affect ....................................................................179
5.3.3 The Role of Attention in Affective Experience ..............................................180
5.3.4 Autonomic Concomitants of Affect and Attention .........................................182
5.3.5 Integration of Neurophenomenological Approach .........................................183
5.4 Yoga’s Role in Exercise Settings...........................................................................184
5.4.1 Yoga as Cardiovascular Exercise....................................................................184
5.4.2 Yoga as Relaxation .........................................................................................186
5.4.3 Yoga as Yoga ..................................................................................................186
5.5 Concluding Thoughts .............................................................................................187
BIBLIOGRAPHY ..........................................................................................................190
xii
List of Appendices
Appendix A. Consent Forms........................................................................................232
Appendix B. Demographics Questionnaire .................................................................238
Appendix C. Medical Information Questionnaire .......................................................239
Appendix D. Previous Yoga Experience Questionnaire ..............................................240
Appendix E. Beliefs about Yoga Scale (BAYS) .........................................................241
Appendix F. Feeling Scale (FS) ...................................................................................242
Appendix G. Felt Arousal (FAS) .................................................................................242
Appendix H. Associative-Dissociative Scale (A/D S) .................................................242
Appendix I. Godin Leisure Time Exercise Questionnaire (GLTEQ) ..........................243
Appendix J. Activation-Deactivation Adjective Questionnaire (AD ACL) ................244
Appendix K. Five Facet Mindfulness Questionnaire (FFMQ) ....................................245
Appendix L. Profile of Mood States – Short Form (POMS-SF) .................................247
Appendix M. Functional Assessment of Cancer Therapy (FACT-G) .........................249
Appendix N. Calgary Symptoms of Stress Inventory (SOSI-C) .................................251
Appendix O. Borg Rate of Perceived Exertion (RPE) .................................................255
Appendix P. Qualitative Research Questions ..............................................................256
Appendix Q. Yoga Class Attendance Log ...................................................................257
Appendix R. Yoga 3 and 6 Month Follow-Up ............................................................258
Appendix S. Yoga for Cancer Survivors 7-Week Protocol (Longitudinal Study) ......259
Appendix T. Study 3 Yoga Protocol (Lab Study)........................................................267
xiii
List of Tables
Table 2.1 Studies Reviewed from the Yoga-Cancer Literature – Control Design .....50
Table 2.2 Studies Reviewed from the Yoga-Cancer Literature – Single Group Design ............... 52
Table 2.3 Health-Related Quality of Life Measures – Control Design.......................................... 54
Table 2.4 Psychosocial Measures – Control Design ..................................................................... 56
Table 2.5 Symptom Measures – Control Design ........................................................................... 57
Table 2.6 Health-Related Quality of Life Measures – Single Group Design ................................ 58
Table 2.7 Psychosocial Measures – Single Group Design............................................................. 60
Table 2.8 Symptom Measures – Simple Group Design ................................................................. 61
Table 3.1 Demographics ................................................................................................................ 95
Table 3.2 Pre-Post Class Estimated Marginal Means Models ....................................................... 96
Table 3.3 Acute Associations: Pre-Post Class Multilevel Regression Model ............................... 97
Table 3.4 Longitudinal Estimated Marginal Means Model ........................................................... 98
Table 3.5 Longitudinal Multilevel Regression Model ................................................................. 100
Table 3.6 Yoga Practice Maintenance Estimated Marginal Means Model .................................. 102
Table 3.7 Predictors of Yoga Practice Maintenance .................................................................... 103
Table 4.1 Demographics .............................................................................................................. 156
Table 4.2 Estimated Marginal Means Model – Valence, Activation, Attention and Rating of
Perceived Exertion ...................................................................................................................... 157
Table 4.3 Estimated Marginal Means Model – Heart Rate, HF HRVnu, SDNN, RMSSD and
SamEn .......................................................................................................................................... 158
Table 4.4 Multilevel Regression Analyses – Valence, Activation, Attention and Rating of
Perceived Exertion ....................................................................................................................... 159
Table 4.5 Estimated Marginal Means Model – AD ACL ............................................................ 161
Table 4.6 Multilevel Regression Analyses – AD ACL ................................................................ 162
xiv
List of Figures
Figure 3.1 Circumplex Model of Affect ...................................................................................... 104
Figure 3.2 Participant Recruitment Flow Diagram ...................................................................... 105
Figure 3.3 Affect Circumplex 2D Model (valence, activation) ................................................... 106
Figure 3.4 Affect Circumplex 3D Model (valence, activation, attention) ................................... 107
Figure 3.5 Continuous Moderators: total physical activity (LSI), moderate-vigorous physical
activity (GLTEQ) and previous Yoga Thrive experience ............................................................ 108
Figure 3.6 Longitudinal Changes in Affect (AD ACL) ............................................................... 110
Figure 3.7 Longitudinal Changes in Mindfulness (FFMQ) ......................................................... 112
Figure 3.8 Longitudinal Changes in Mood Disturbance (POMS), Stress Symptoms (C-SOSI) and
Quality of Life (FACT-G) ........................................................................................................... 115
Figure 3.9 Yoga Practice Maintenance ........................................................................................ 117
Figure 4.1 Circumplex Model of Affect ...................................................................................... 163
Figure 4.2 Simple Slopes of the Regression for Dissociative Attention on Condition at Low,
Medium and High Levels of Sample Entropy ............................................................................. 164
Figure 4.3 Estimated Marginal Means Models for Valence, Activation, Attention and Perceived
Exertion ........................................................................................................................................ 165
Figure 4.4 Estimated Marginal Means Models for Heart Rate, HF HRVnu, SDNN, RMSSD and
Sample Entropy............................................................................................................................ 167
Figure 4.5 Estimated Marginal Means Models for Activation-Deactivation Adjective Checklist
..................................................................................................................................................... 170
Figure 4.6 Affective Response (Mean FS and FAS) to a Single Yoga Session Plotted on a TwoDimensional Circumplex Model .................................................................................................. 172
Figure 4.7 Affective Response (Mean FS, FAS, ADTS) to a Single Yoga Session Plotted on a
Three-Dimensional Circumplex Model ....................................................................................... 173
Figure 5.1 Yoga Mechanisms Integrated Model .......................................................................... 189
xv
List of Abbreviations, Symbols and Nomenclature
Symbol
Definition
ACSM
AD ACL
A/D S
ANS
BMI
Bpm
BAYS
CI
CSEP
C-SOSI
ES
FACIT-F
FACIT-Sp
American College of Sports Medicine
Activation-Deactivation Adjective Checklist
Association / Disassociation Scale
Autonomic Nervous System
Body Mass Index
Beats Per Minute
Beliefs about Yoga Scale
Confidence Interval
Canadian Society of Exercise Physiology
Calgary Symptoms of Stress Inventory
Effect Size
Functional Assessment of Chronic Illness Therapy
Functional Assessment of Chronic Illness Therapy – Spiritual
Wellbeing
Functional Assessment of Cancer Therapy
Functional Assessment of Cancer Therapy – General
Felt Arousal Scale
Five Facet Mindfulness Questionnaire
Feeling Scale
Generalized Estimated Equation
Godin Leisure Time Exercise Questionnaire
High Frequency Heart Rate Variability normalized units
Heart Rate
Health-Related Quality of Life
Heart Rate Variability
Intraclass Correlation Coefficient
Leisure Score Index (total physical activity)
Mindfulness-Based Stress Reduction
Minimally Clinically Important Difference
Physical Activity
Profile of Mood States
Profile of Mood States – Short Form
Patient Reported Outcomes Measurement Information System
Quality of Life
Root Mean Square of Successive Differences
Rate of Perceived Exertion
Sample Entropy
Standard Deviation of NN (normal-to-normal) heart rate intervals
Short Form Health Survey (12-item)
Symptoms of Stress Inventory
Visual Analog Scale
FACT
FACT-G
FAS
FFMQ
FS
GEE
GLTEQ
HF HRVnu
HR
HRQL
HRV
ICC
LSI
MBSR
MCID
PA
POMS
POMS-SF
PROMIS
QoL
RMSSD
RPE
SamEN
SDNN
SF-12
SOSI
VAS
xvi
Epigraph
Yoga is the stilling of the fluctuations of the mind.
Yoga Sutras of Patanjali, Chapter 1, Verse 2
xvii
1
CHAPTER 1: INTRODUCTION
2
Yoga appears to be a promising complementary exercise choice for cancer
survivors. Positive effects have been seen on indices of mood, stress-related symptoms,
health-related quality of life (HRQL), fatigue and sleep (1-3). Despite preliminary
findings, little attention has been paid to the psychological mechanisms by which benefits
are accrued via yoga practice. The current research examines: 1) the clinical significance
of patient–reported outcomes from yoga interventions for cancer survivors; 2)
associations between yoga participation, affect, mindfulness and mood, stress and HRQL
outcomes in cancer survivors longitudinally in an ongoing community-based yoga
program; and 3) associations between valence, activation, attention, perceived exertion
and cardiac activity in a single yoga session for cancer survivors. The larger aim of this
set of studies was to develop new models to better describe the mechanisms underlying
how yoga practice elicits improvements in HRQL and psychological outcomes in cancer
survivors.
The present chapter provides a concise overview of the role of exercise in
improving mental health and HRQL in cancer survivors, a brief introduction to the use of
yoga in cancer settings, a survey of models, constructs, and approaches helpful to the
present work and a synopsis of study objectives. Chapter 2 is a systematic review of yoga
interventions for cancer survivors with an emphasis placed on the clinical significance of
these outcomes. This chapter was published in its entirety in Evidence-Based
Complementary and Alternative Medicine. Chapter 3 is an original longitudinal study
examining the effects of yoga on measures of mood disturbance, stress symptoms and
HRQL and the associations between these outcomes, affect, mindfulness and yoga
practice. Chapter 4 is an original laboratory study examining the effect of yoga on
3
valence, activation, attention and perceived exertion during a single yoga session and the
relations between these measures and cardiac activity, particularly heart rate and heart
rate variability (HRV). Chapter 5 summarises and integrates the research findings from
the previous three chapters to provide an encompassing model of how yoga elicits change
in cancer survivors.
1.1 Cancer
1.1.1 Prevalence
Over 500 Canadians are diagnosed with cancer daily and 200 die from cancer
each day. There will be an estimated 186,400 new diagnoses of cancer and an estimated
75,700 cancer-related deaths in Canada in 2012, with 69% of new cases and 62% of
cancer deaths occurring among those 50 to 79 years of age. Of new diagnoses, over half
will be either lung, colorectal, prostate or breast cancers (4).
Receiving a cancer diagnosis, undergoing treatment and subsequent recovery takes a
great toll on many cancer survivors. The cancer experience is associated with high
emotional distress and lowered HRQL (5). HRQL can be defined as a state of wellbeing
comprised of two components: 1) the ability to perform daily activities that reflect
physical, functional, psychological and social wellbeing 2) individual satisfaction with
levels of functioning and control of disease and/or treatment-related symptoms. HRQL
usually includes functional (e.g., whether a patient is able to engage in activities of daily
living), physical (e.g., pain, shortness of breath, fatigue), psychological (e.g., depressive
symptoms, positive affect), and social (e.g., involvement with others) domains (6).
Distress stemming from the cancer experience is a significant problem for up to half
of all cancer patients (7). Cancer survivors may exhibit signs of clinical psychological
4
disorders and a host of more general mood, stress, fatigue and sleep-related symptoms (812). In addition, there are often chronic and late-appearing effects of illness and/or
treatment, and increased risk of cancer recurrence or other diseases (13).
1.1.2 Cancer and Exercise
An ever-growing area of research highlights the positive role physical activity
(PA) and exercise may play in alleviating many of the negative side-effects of cancer and
its subsequent treatments. PA can be defined as any movement of the body resulting in
increased energy expenditure (6). PA encompasses all bodily movements, from sports to
activities of daily living. Exercise is a specific type of PA consisting of planned,
structured bodily movement with the express goals of improving or maintaining physical
fitness (6).
Regardless of intervention specifics, exercise has been shown to improve a variety
of HRQL, psychosocial and health outcomes in various cancer survivor groups both
during and after cancer treatment, and may also help to manage the long-term effects of
treatment (14,15). In a national survey (16), those meeting American Cancer Society
recommendations for exercise had higher HRQL. A literature review of 25 intervention
studies indicated increased exercise led to increased HRQL (17). Those with greater
exercise-related PA reported higher HRQL and better mental health outcomes throughout
the cancer experience as opposed to their sedentary counterparts (18). Current research
evidence also suggests exercise improves cancer-related fatigue across several types of
cancer (9,19). Moreover, increased PA may decrease the risk of cancer recurrence and
extend survival (20-22). The sooner cancer survivors maintain, re-establish or improve
5
upon pre-diagnosis PA and body composition within their cancer experience; the more
likely they are to report improved HRQL and mental health outcomes.
Within the broad scope of oncology, there is a growing interest in promoting PA
in general and the development of exercise interventions specifically to ameliorate
cancer’s destructive role in survivors’ overall HRQL and mental health. The ACSM
roundtable on Exercise Guidelines for Cancer Survivors (23) concluded exercise is safe
both during and after cancer treatments, resulting in improvements in HRQL, physical
functioning, and cancer-related fatigue in several cancer survivor groups. ACSM
Guidelines suggest cancer survivors, regardless of where they are in the treatment
continuum, should avoid inactivity and that any level of PA carries with it some benefit
(23). Current recommendations suggest an overall volume of weekly activity of 150
minutes of moderate-intensity exercise, or 75 minutes of vigorous-intensity exercise, or
an equivalent combination. These recommendations are not dissimilar from general
population guidelines and have been shown to be associated with a variety of health
outcomes (24). These cancer-specific guidelines are embedded within the proviso that
exercise programs must be adapted for the individual on the basis of their health status,
treatments, and anticipated disease trajectory (23).
Despite reported benefits and recommendations, it is widely acknowledged many
cancer survivors never return to either initial rates or healthy levels of PA (24). Data
suggests only 22% of cancer survivors are active enough to achieve health benefits (25)
and only 50% of cancer patients offered an exercise program will undertake and complete
the intervention (26). Even in the general population, the amount of time spent engaged
6
in PA declines across the lifespan and the amount of time spent in sedentary activities
increases (27,28).
Given poor exercise adoption, adherence and maintenance in cancer and general
populations, determinants of health behaviours are considered paramount. These include
physiological, psychosocial and environmental factors (6). Addressing these issues often
leads to improved health outcomes. Current research advocates for the translation of
research on the HRQL and mental health benefits of PA into optimal theoretical models
of behaviour change in which determinants of PA, including exercise preferences, are
incorporated into exercise interventions (29).
1.2 Yoga
1.2.1 Yoga in Cancer Settings
Within the growing body of cancer and exercise research, there have been calls to
examine modes of exercise from the area of complementary medicine (30,31). In this
regard, yoga has emerged as an important complementary therapeutic approach in cancer
settings, often considered a gentle, low-intensity form of exercise (22,32). Preliminary
results from the emerging literature on yoga and cancer are promising and provide
support for the utility of yoga interventions for cancer survivors. Research examining the
effects of these interventions on cancer patients suggest yoga participants compared to
waitlist controls or those in supportive therapy groups show greater improvements in
overall HRQL, psychological health, stress-related symptoms, fatigue and sleep indices
(1,2,33-36), though reported improvements have not been uniform (37).
7
1.2.2. Yoga and Exercise
Recent reviews suggest yoga has positive effects on general physical and mental
health (38-41), fitness in older adults (42), mood (43-45), symptoms of stress (46), and
HRQL in older adults (47). Studies comparing the effects of yoga and exercise indicate
that in both healthy individuals and those with various health conditions, yoga may be as
effective as walking, cycling, jogging or aerobics at improving a variety of health-related
outcome measures and may carry added effects in the alleviation of stress (48).
1.2.3 What is Yoga?
The term yoga stems from traditional philosophies from the Indian subcontinent.
Most commonly referenced are the Yoga Sutras of Patanjali, an ancient meditation text of
indeterminate origin (49) and its various commentaries. The Yoga Sutras four chapters
outline both a philosophical and pragmatic path to enlightenment, including attendant
ethical prescriptions (yamas), personal observances (niyamas) and the development of
various accomplishments (siddhis) along the way.
In the 21st century, yoga practice as a system of physical exercise has become
prominent (50). This unique integration of both moving and static sequences (asana),
breathing exercises (pranayama), and different meditation tools to withdraw the senses
(pratyahara), concentrate the mind (dharana) and develop abilities of impartial
awareness (dhyana) have all been used in modern settings as means of increasing
performance and recovery in both general and clinical populations (40). These practices
are routinely modified based on desired outcomes as well as participant health status (51).
Yoga’s transition from a system designed for those seeking enlightenment to
contemporary yoga, a system comprised of physical, breathing and meditation
8
techniques, remains controversial. There are those that maintain today’s yoga is part of an
extant system passed down from antiquity, to those that suggest contemporary yoga is
largely a modern invention (50,52-54). Given the complexity of yoga, both in terms of its
wide and varied practices and the historical complexity from which it emerged, it is
perhaps best thought of as the “source” of a variety of different practices based on
geography, historical origins, lineage and purpose (55). From this viewpoint yoga
practice may have numerous applications, ranging from enlightenment (IV.34), as per the
Yoga Sutras, to ordinary physical and mental training (55).
A common definition of yoga is to link or “yoke” things together. This definition
has been widely interpreted, from the linking of The Universal (Brahman) and the
personal (atman) (56), to linking the mind to an object of attention (57). In contemporary
yoga practice, the object of attention could include movement of the body, sensation of
the breath, the contents of the mind, a mantra, or a variety of other tools. Perhaps a
clearer definition of what yoga “is” or “does” is that found in the Yoga Sutras of
Patanjali, which characterises yoga as the “stilling of the fluctuations of the mind (YS
1.2).” Control of these mental fluctuations comes from persevering practice aimed at both
attaining and maintaining states of mental stability over long, uninterrupted periods of
time (YS 1.12-1.14). Importantly, meditative states experienced within yoga practice are
correlated to neurophysiological systems that mediate both positive affect and
internalised attention (58,59). The Yoga Sutras trace all benefits of yoga to developing
mastery over ones’ mind (60), regardless of whether the personal goals of those
practicing yoga are to increase flexibility, relax more, alleviate health concerns or, in the
present case, cope with cancer diagnosis, treatment and/or recovery.
9
The previous section has provided a brief overview of yoga in cancer settings, the
relationship between yoga and exercise and a working definition of what yoga “is” or
“does.” Further details on yoga practice in cancer settings are provided in Chapter 2. The
next section examines potential mechanisms for yoga’s health-related effects.
1.3 Mechanisms of Action
Therapies like yoga, directed toward addressing functional links between mind
and body, may be particularly effective in treating symptoms associated with chronic
illness (61). As research in the use of yoga in cancer settings continues to grow and yoga
is integrated into cancer care, it is imperative the clinical benefits of yoga for cancer
survivors are better understood. Identifying mechanisms that explain how yoga leads to
clinically significant outcomes is the next step in understanding the ability of yoga to
target desired outcomes. In this capacity, there have been recommendations to integrate
patient-reported outcomes with biomedical endpoints as a means of better describing the
complexity of these measures (62). While these outcomes may purport treatment effect,
they do little in explaining patients’ baseline characteristics, the interventions themselves,
and purported mechanisms of action.
Recent recommendations from the UK Medical Research Council (63) suggest
time spent examining these mechanisms, both theoretically and practically, helps to
strengthen the causal chain of evidence and further refine both the interventions
themselves and research design. Reviews in the area of exercise and cancer also advocate
for further research to define the relations between the psychology and physiology of
exercise and their respective contributions to health outcomes (15,22,26,29,64-66).
10
Future research must explore the complementary psychophysiological mechanisms by
which exercise and yoga improve cancer survivor health outcomes.
In the following studies, three primary mechanisms for yoga’s benefits are
suggested. They are: 1) positive affect; 2) attention-mindfulness; 3) cardiac activity as
measured via heart rate and heart rate variability (HRV). Within each proposed
mechanism, supportive theories, models, constructs, and approaches are described and
serve as a foundation for the subsequent research.
1.3.1 Positive Affect
A current area of clinical research interest is the study of positive affect as an
independent, adaptive pathway in the cancer experience. Positive affect can be broadly
defined as feelings that elicit pleasurable experience and may include descriptors such as
happiness, joy, contentment, and peacefulness among others (67). Positive affect provides
a resource for cancer survivors to cope with the demands of cancer diagnosis, treatment
and recovery (68). In a study examining 215 recently diagnosed colorectal cancer
patients, those reporting lower positive affect reported higher anxiety and depression. In
addition, changes in positive affect significantly moderated associations between
increased symptom distress, anxiety and depression (68). A study examining both
positive affect and negative affect in a sample of 133 lung cancer patients (69) further
suggested negative affect was associated with impaired physical and social functioning,
emotional problems, pain and poorer general health, while positive affect was associated
with adaptive social functioning, fewer emotional problems, less severe pain and greater
general health. In a study of 50 early-stage breast and prostate cancer patients receiving
radiotherapy, patients with higher levels of positive affect had an improved acute
11
inflammatory response to radiation treatment, believed to enhance the tissue repair
process (70). In sum, reduction of negative affect and maintenance or improvement of
positive affect are important therapeutic goals, likely to lead to enhanced HRQL and
perhaps improved treatment outcomes.
In general, exercise of low- to moderate-intensity increases positive affect and
reduces negative affect (71,72). Positive affect tends to increase pre- to post-exercise
following exercise intensities that are not exhaustive (73). While a majority of research
examining psychological changes associated with exercise is focused on reductions in
negative emotions (depression, anxiety), actual exercise recommendations tend to focus
on making activity more enjoyable rather than on surmounting mental health concerns
(74). The assumption is that if exercise is pleasant and enjoyable, adoption and adherence
rates might improve and exercise benefits will be more readily experienced (74).
Positive affect experienced during exercise is a strong predictor of future exercise
engagement and may enhance adherence to exercise programs, further solidifying the
benefits obtained through regular exercise (73). For example, in a sample of 37 healthy
adults, affect experienced during an acute bout of moderate-intensity exercise was
predictive of exercise participation 6 and 12 months later (75). This evidence suggests the
influence of affective responses resulting from exercise may be important with respect to
longer term exercise maintenance. In cancer settings, fostering positive affective
reactions after a structured exercise program is an important target for interventions
designed to facilitate post-program exercise adherence (76).
Clinical research has begun to further explore the specific relations between yoga
and positive affect. In two nonrandomized studies, single yoga sessions were associated
12
with significant improvements in positive affect and reductions in negative affect and
perceived stress, comparable to changes seen with aerobic exercise (77,78). A pilot RCT
examining the effects of restorative yoga in 44 breast cancer survivors either on or off
treatment also reported significant benefits favouring the yoga group for positive affect,
mental health, depression and spirituality outcomes (79). A second pilot RCT examining
the effects of an integrated yoga program in 88 breast cancer outpatients undergoing
adjuvant radiotherapy reported significant improvement in positive affect, emotional and
cognitive function, and decreased negative affect in the yoga group as compared to
controls (80). These findings suggest yoga practice is associated with increased positive
affect, reduced negative affect and improved health outcomes.
1.3.1.1 Circumplex Model of Affect. Given the proposed importance of positive
affect and yoga’s influence on this phenomenon, the Circumplex Model of Affect (81)
offers an important parsimonious explanatory approach. At the heart of mood and
emotion are underlying psychophysiological states experienced as simply feeling good or
bad, energized or fatigued. According to this model, affect is best understood as an
underlying state comprised of two orthogonal dimensions: 1) valence, or whether states
are perceived as good or bad, and 2) activation, or whether states are experienced as
energizing or fatiguing (82). These two dimensions of valence and activation, termed
core affect, underlie all affective states. Upon this affective base are layered various
cognitive processes that interpret and refine emotions and moods (81).
Affect attributed to an immediate cause is considered an emotion. For example,
emotion is directed at something (Ted was afraid of the bear). Emotion requires an
antecedent event, which elicits core affect. This core affect is then attributed to its
13
antecedent, requiring further appraisal of the antecedent itself and action to be taken in
the direction of the antecedent (in the case of the bear, one might consider running...)
(81). Mood is described by Russell (2003) as “long-term core affect” and is objectless in
the sense that mood is not outwardly directed toward anything. However, based on the
principle of mood congruency, what we attend to and cognize in the field of
consciousness is based on these underlying affective states (81). Independent of the
temporal nature of affect, at the intersection of valence and activation, lay four affective
quadrants which can be described as: high activation pleasure (energy), low activation
pleasure (calmness), high activation displeasure (tension), and low activation displeasure
(tiredness) (83). Based on these two dimensions and four quadrants, a host of affective
responses, emotions and moods are possible.
1.3.1.2 Dual Mode Theory. The Circumplex Model has been further developed
for exercise settings utilising Dual Mode Theory (84), which postulates changes in core
affect, valence and activation, within a bout of exercise are regulated via the continuous
interplay of two mechanisms: 1) cognitive processes, including exercise goals, selfefficacy, expectations (75); and 2) interoceptive cues, or physical sensations from the
body. The relative importance of these two factors in the regulation of core affect is
hypothesized to shift systematically as a function of exercise intensity, with cognitive
factors being dominant at low and moderate intensities and interoceptive cues becoming
more prominent at higher intensities (85). Intensities above the lactate or ventilatory
threshold often lead to negative affective experiences of exercise, as workload increases
to a point that exercise cannot be physiologically maintained and interoceptive cues
indicate one should stop. Exercise below this threshold is said to lead to positive affective
14
experiences, as individuals are able to maintain this workload, which bolsters cognitive
factors including increased self-efficacy and positive exercise beliefs (86). There is large
inter-individual variability close to the ventilatory or lactate threshold (73). The lactate
threshold is the point at which lactate concentrations in blood exceed the rate at which
lactate can be removed, resulting in excess lactate and a shift towards anaerobic
metabolism, while the ventilatory threshold is the point at which ventilation (air in and
out of the lungs) deviates from a linear sequence and carbon dioxide cannot be removed
from the body at the rate it is produced. Ventilatory threshold is closely related with
lactate threshold, reflecting the same basic shift toward anaerobic metabolism during
greater exercise intensity (87).
The previous section provided an overview of the role of affect in cancer settings
and the relations between affect, exercise and exercise adherence. The Circumplex Model
of Affect provides a dimensional model of affective processes and Dual Mode Theory
suggests how affect is regulated in exercise settings. These two theories figure
prominently in Chapters 3 and 4 of the present work.
1.3.2 Attention
Within the exercise psychology field, the role of attention in performance has
been researched extensively. A seminal study in this field by Morgan and Pollock (88)
examines the attention strategies of distance runners. Athletes could be subdivided into
elite and non-elite based on these attention strategies. Non-elite runners were
characterized by dissociative strategies designed to dissociate from painful running
stimulus, while elite runners utilized an associative attention strategy, paying attention to
physiological cues to monitor running pace and other running-specific cues.
15
Individual focus of attention during exercise can influence changes in affective
responses (89). Associative attention is defined as focusing attention on present-moment
physical sensations related to exercise, in which exercisers seek to monitor sensory input
and adjust their effort accordingly. Dissociative attention is defined as focusing on nonexercise related stimuli that divert attention away from internal sensations and present
moment exercise experience, including listening to music, watching TV, talking on the
phone or attending to other thoughts and feelings unrelated to exercise (90). For fit
individuals, associative attention allows for better regulation of effort, while dissociative
strategies have been recommended for individuals beginning exercise programs and
unused to the physiological strain of exercise (89).
1.3.2.1 Effort-Related Attention Model. The Effort-Related Attention Model
spearheaded by Morgan (88) and refined by Tenenbaum (91), suggested attention
strategies during exercise vary based on intensity. Specifically, during low or moderate
intensity workloads, attention is voluntary and can switch between dissociative and
associative attentional strategies, as external stimuli do not compete with internal stimuli.
At higher exercise intensities, increased physiological cues demand attention. As a result,
attention narrows, shifts internally and becomes associative (92). Based on exercise
intensity, dissociative attention strategies may make exercise “feel easier” at low tomoderate intensities (91), but this coping strategy is “load dependent” and ineffectual at
higher intensities (93).
The previous section provides an overview to associative and dissociative
attention strategies in exercise settings and suggests focus of attention is related to
affective experiences of exercise. The Effort-Related Attention Model proposes attention
16
and affect are linked via exercise intensity. The Effort-Related Attention Model was used
to build upon the Circumplex Model of Affect and Dual Mode Theory and is featured in
both Chapters 3 and 4 of the current work.
1.3.3 Mindfulness
Mindfulness is the systematic development of the ability to direct attention towards
events in the field of consciousness non-judgementally in the present moment (94).
Salmon and colleagues (2010) suggested mindful awareness is similar to associative
attentional processing, but that the inclusion of these non-judgmental qualities enhance
accurate observation and acceptance of both one’s internal bodily cues, the content of
thoughts and the external environment (95). Within exercise settings, mindfulness builds
upon associative attention by refining perceptions of perceived exertion on a moment-bymoment basis, via increased sensitivity to a host cognitive and interoceptive cues.
Mindfulness concurrently limits emotional reactivity to these cues via acceptance (95).
It is hypothesised the strong mind-body interaction within contemporary yoga
practice adds a unique contemplative dimension to exercise and has been referred to as
‘mindfulness in motion’ (96). The practice of yoga provides an opportunity for sustained
attention to the body, breath and mind through a progressive sequence of dynamic
movements, restful postures, breathing exercises and periods of meditative awareness.
Froeliger and colleagues (97) suggest the object of mindfulness in contemporary yoga
practice comes from both physical sensations of the body moving and awareness of
breath.
However, research evidence linking yoga in eliciting these changes in
mindfulness remains equivocal. Lengacher and colleagues (98) found total minutes of
17
yoga practice as part of an overall MBSR program was not significantly related to
positive changes in psychosocial status and HRQL. These findings are in contrast to the
work of Carmody and Baer (99), whose research in a non-cancer heterogeneous medical
population found the yoga component of a mindfulness program to be most strongly
related to improvements of mindfulness measures, psychological symptoms and
perceived stress. These findings have more recently been corroborated in a study by
Sauer-Zavala and colleagues (100), which reports yoga practice was associated with
greater psychological wellbeing independent of body scan and sitting meditation
practices.
The previous section has provided a brief introduction to mindfulness, the link
between mindfulness and associative attention strategies, and the role of mindfulness in
exercise settings. It is hypothesized yoga practice is a form of mindfulness but findings in
this area are preliminary and equivocal. The construct of mindfulness in relation to yoga
practice features prominently in Chapter 3 of the current work.
1.3.4 Autonomic Nervous System & Heart Rate Variability
A commonly reported neurophysiological mechanism for relations between yoga
practice, positive affect and improved attention regulation via mindfulness lies in the
Autonomic Nervous System (ANS). The ANS is the most prominent physiological factor
in determining heart functions and is comprised of both parasympathetic and sympathetic
input. The amount of input received from each is examined through the study of heart
rate variability (HRV), the modulation of the heart rate about its mean value (101).
Primarily, the heart is under tonic inhibitory control by parasympathetic influences, and
resting cardiac autonomic balance favours energy conservation by parasympathetic
18
dominance over sympathetic influences (102). The basic data for calculating HRV
measures is the sequence of time intervals between heart beats, also known as the R-R
interval. Increases in sympathetic activity, which cause the time between heart beats to
become shorter, are associated with heart rate increases, while increases in
parasympathetic activity cause the heart rate to decrease by increasing the R-R interval.
Sympathetic effects, usually taking seconds to respond, are slow in comparison to the
parasympathetic effects, which take milliseconds. Therefore, parasympathetic influence
is responsible for making the rapid changes in the timing of the heart (102). An increase
in parasympathetic activity is also expected with deep breathing and low respiration rates
(103). HRV offers a non-invasive quantitative method of investigating autonomic effects
on the heart and has been increasingly used as an index of affect regulation (104-106). In
general, higher vagally-mediated HRV is associated with positive health outcomes
(106,107).
Yoga programs are reported to result in increased parasympathetic nervous
system (PNS) activity and significantly improved cardiac autonomic function (108).
Streeter and colleagues (2012) hypothesize yoga practice corrects PNS under-activity via
stimulation of the vagus nerves to ameliorate stress-related illness symptoms (109). One
potential mechanism of action is that the conscious regulation of breathing practiced in
yoga may influence autonomic function by regulating the balance of sympathetic and
parasympathetic activation, the latter increasing as the breath deepens and slows
(110,111). Yoga practice also appears to result in increased vagal modulation and control
over autonomic responses such as heart rate (112,113). During active poses, heart rate
and respiration may increase consistent with exercise (114). However, during and
19
following the restful supine and seated meditative postures that conclude most yoga
classes, parasympathetic activity consistent with a reduction in physiological arousal is
apparent (115).
1.3.4.1 Neurovisceral Integration Model. The Neurovisceral Integration Model
integrates affective, attentional and autonomic systems into a theoretical structure that
aids in understanding affect regulation (106). This model suggests attention, valence and
activation are vagally mediated. Specifically, the Neurovisceral Integration Model links
the dimensions of valence, activation and attention via cardiac autonomic function,
specifically PNS function, as measured by heart rate and HRV. Research utilising the
Neurovisceral Integration Model has demonstrated those with lower HRV are at risk for
both psychological and physiological dysfunction (116) and, ultimately, increased
morbidity and mortality (117). In exercise settings, evidence suggests sedentary lifestyles
result in negative mood symptoms, especially feelings of fatigue and depressive
symptoms, which are related to decreases in PNS function (118). Conversely, regular
exercise induces training adaptation in ANS, particularly greater parasympathetic
modulation of cardiac function (119). Daily exercise also improves positive affect by
shifting autonomic balance to parasympathetic predominance (118).
We chose to examine the Neurovisceral Integration Model for two reasons: 1)
given the linkage between yoga’s benefits and HRV, it was important to include a model
which used HRV as its primary indicator; and 2) both Dual Mode Theory and the EffortRelated Attention Model suggest affective experiences in exercise settings are driven by
the physiological intensity of exercise. Common physiological indices of this intensity
are ventilatory and lactate thresholds. However, from an exercise physiology perspective,
20
yoga practice is generally considered low-intensity and occurs below lactate or
ventilatory threshold (120). Given this intensity, we required an additional theoretical
model and accompanying physiological measure sensitive to lower-intensity exercise
changes. The Neurovisceral Integration Model figures prominently in Chapter 4.
1.3.5 Neurophenomenology
The current research is concerned with the phenomenological lived experience of
contemporary yoga practice, related psychosocial health outcomes and potential
mechanisms of action. We are particularly interested in examining these phenomena
within the practice of yoga via the lens of neurophenomenology, a convergent mixed
methods research approach that combines qualitative subjective first-person narratives,
quantitative self-reports and physiological data to develop a layered descriptive research
approach to conscious experience (121).
Specifically, the inclusion of subjective first-person descriptions of the
phenomena in question are with the express aim of further describing these events of
consciousness and how they pertain to the quantitative measures (122). These reports are
useful for identifying physiological variability in responses to stimuli from moment to
moment. It is believed these descriptions provide unique information that aid in the
interpretation of physiological processes. The strategy of neurophenomenology is to
provide mutually informative insights between first-person and neurophysiological
accounts as a means of generating new data in the exploration of consciousness (123).
The importance of contemplative mental training is central to this novel research
approach and has been used extensively in examining contemplative practices (55) using
fMRI and EEG investigative methods (124-128).
21
Despite use of neurophenomenology in the field of meditation and contemplative
studies (129), this approach had not been used previously in a yoga setting prior, nor had
HRV been used as the primary neurophysiological indices. This overarching
methodological approach features prominently in Chapter 4.
1.4 Research Objectives
Taking Patanjali’s definition that yoga is the, “stilling of the fluctuations of the
mind (YS 1.2)” as a guide, we suggest the benefits of yoga practice experienced in cancer
settings are related to mechanisms of affect, attention-mindfulness and regulation of the
autonomic nervous system. The Circumplex Model of Affect, Dual Mode Theory, EffortRelated Attention Model, construct of Mindfulness, Neurovisceral Integration Model, and
Neurophenomenology helped further encapsulate these mechanisms of action. Based on
the premise that: 1) yoga’s primary purpose is to control the mind via 2) affect and
attention regulation and its autonomic correlates, leading to 3) improved HRQL and
mental health outcomes, three studies were developed to examine the relations between
yoga practice, affect and attention regulation in cancer settings. The current research
examines: 1) the clinical significance of patient–reported outcomes from yoga
interventions for cancer survivors; 2) associations between yoga participation, affect,
mindfulness and health outcomes in cancer survivors longitudinally in an ongoing
community-based yoga program; and 3) associations between core affect, attention,
perceived exertion and cardiac activity in a single yoga session for cancer survivors.
Study objectives were as follows:
22
1.4.1 Study 1 (Clinical Significance Review)
Objective 1a: Examined how yoga programs have been offered in cancer settings
and what outcomes were measured. Objective 1b: Examined whether yoga improved
measured outcomes and by what magnitude using measures of clinical significance.
1.4.2 Study 2 (Longitudinal Study)
Objective 1a: Examined potential changes in valence, activation, and attention
pre-post each class of a seven-week yoga intervention. Objective 1b: Examined whether
potential changes in valence, activation and attention were related to one another or
moderated by baseline affect or mindfulness Objective 2a: Examined changes pre-post
and at three- and six-month follow-up, in measures of mood disturbance, stress symptoms
and HRQL as well as proposed predictor variables, including affect and mindfulness
Objective 2b: Examined whether potential improvements in mood disturbance, stress
symptoms and HRQL were longitudinally associated with concurrent changes in affect
and mindfulness Objective 3a: Examined whether participants maintained their yoga
practice longitudinally, pre-post and three- and six-month follow-ups. Objective 3b:
Examined whether maintenance of yoga practice was associated with mood disturbance,
stress symptoms and HRQL as well as proposed predictor variables, including affect and
mindfulness.
1.4.3 Study 3 (Lab Study)
Objective 1a: Examined changes in valence, activation, attention and perceived
exertion during a single yoga session as well as associated measures of cardiac activity,
including heart rate and HRV. Objective 1b: Examined whether changes in valence,
activation, attention and perceived exertion are related to one another and/or changes in
23
cardiac activity. Objective 1c: Examined whether changes in valence, activation,
attention, perceived exertion and cardiac activity were related to subjective first-person
descriptions of the single yoga session. Objective 2a: Examined whether significant
changes in dimensions of affect, including measures or energy, tension, tiredness and
calmness, occur pre-post yoga class. Objective 2b: Examined whether changes in
dimensions of affect were related to one another and/or to concurrent changes in valence,
activation, attention, perceived exertion, and cardiac activity. Objective 2c: Examined
whether changes in affect were related to subjective first-person descriptions of the yoga
session.
1.5 Significance
Little attention has been paid to the mechanisms by which health benefits from
yoga practice accrue and the experience of yoga has not been studied within these
theoretical frameworks or using this integrated methodology to date. This research
develops current theory in yoga practice, affect and attention regulation by modelling
processes and outcomes, resulting in a detailed exploration of yoga in cancer settings.
Examining these proposed theoretically-based mechanisms for yoga’s salutary effects
enables understanding not only of “if” yoga works, but also “how.” This research
provides a foundation to more fully examine the use of yoga in cancer settings. This
knowledge can then be transferred directly back into the community to further develop
innovative yoga programs and best practices, with the express aim of improving HRQL
and mental health in cancer survivors.
24
CHAPTER 2: YOGA AND CANCER INTERVENTIONS: A REVIEW OF THE
CLINICAL SIGNIFICANCE OF PATIENT REPORTED OUTCOMES FOR
CANCER SURVIVORS
25
2.0 Abstract
Limited research to date suggests yoga may be a viable gentle physical activity
option with a variety of health-related quality of life, psychosocial and symptom
management benefits. The purpose of this review was to determine the clinical
significance of patient-reported outcomes from yoga interventions conducted with cancer
survivors. A total of 25 published yoga intervention studies for cancer survivors from
2004-2011 had patient-reported outcomes, including quality of life, psychosocial or
symptom measures. Thirteen of these studies met the necessary criteria to assess clinical
significance. Clinical significance for each of the outcomes of interest was examined
based on 1 standard error of the measurement, 0.5 standard deviation, effect sizes and
their respective confidence intervals. This review describes in detail these patientreported outcomes, how they were obtained, their relative clinical significance and
implications for both clinical and research settings. Overall, clinically significant changes
in patient-reported outcomes suggest yoga interventions hold promise for improving
cancer survivors’ well-being. This research overview provides new directions for
examining how clinical significance can provide a unique context for describing changes
in patient-reported outcomes from yoga interventions. Researchers are encouraged to
employ indices of clinical significance in the interpretation and discussion of results from
yoga studies.
26
2.1 Introduction
Physical activity is widely accepted as beneficial for cancer survivors both on and
off treatment (130). It is recommended adult cancer survivors engage in the same amount
of physical activity as healthy older adults, with adaptations made as necessary (23). This
recommendation is supported by a substantive amount of evidence documenting the
beneficial effects of physical activity on quality of life, psychosocial and
physical/symptom outcomes (130). Within this growing body of physical activity and
cancer research, there have been calls to examine modes of physical activity from the
area of complementary medicine as a means of improving participation rates and longterm adherence (29). This approach is particularly important when considering potential
additional barriers experienced during a cancer diagnosis, its subsequent treatments, and
ongoing recovery (29,131).
Yoga is quickly emerging as an important complementary medicine therapeutic
approach for cancer survivors. Ross and Thomas’ (48) review of exercise and yoga
highlights that yoga is as beneficial as more traditional types of physical activity at
improving a variety of health-related outcome measures (with the exception of physical
fitness outcomes) in both healthy individuals and those with various health conditions
(e.g., cancer). It concludes that yoga is a potentially beneficial gentle form of physical
activity for cancer survivors and continues to receive growing research attention in cancer
populations (48).
Additionally, four recent reviews have summarized findings specifically on yoga
for cancer (2,33,51,132). Smith and Pukall (33) conducted the first systematic review of
the yoga and cancer research to date. This review reported both the characteristics and
27
effect sizes of ten studies, including six randomized controlled trials. The authors
documented large variability across studies (i.e., yoga type, population, sample size,
intervention duration) as well as methodological limitations. Despite these issues,
generally positive results, especially in terms of psychological outcomes, were noted
based on a calculation of effect sizes and narrative summary of reported results. In
addition, a recently published meta-analysis aimed to determine effects of yoga on
quality of life, psychological and physical health in cancer survivors (2). Ten studies
were examined, including both yoga and mindfulness-based stress reduction (MBSR)
interventions, due to the inclusion of yoga in MBSR. Although the results are preliminary
and should be interpreted cautiously, intervention groups showed significantly greater
improvements in psychological health than waitlist or supportive therapy control groups
(2). Results were inconclusive for quality of life and physical health outcomes.
Undoubtedly, research examining yoga for cancer survivors is in its infancy. However, as
research in this area continues to rapidly grow and yoga is increasingly integrated into
cancer care, it is imperative the clinical benefits of yoga for cancer survivors are better
understood and emphasized within clinical and post-treatment survivorship care. Thus the
focus of the current study is to better determine the implications of the existing research
on yoga for cancer survivors by evaluating additional indicators of clinical significance.
Reviews of the research on yoga for cancer survivors to date have relied on the
interpretation of effect sizes and subjective observation of trends. Subjective
interpretation of trends is often heavily influenced by reporting of the p-value, which is
commonly used as a statistical indicator of the benefits of an intervention. However, the
P-value, often set at .05, indicates only whether or not observed changes are large enough
28
to conclude that such differences were not caused by chance (133). In addition, the Pvalue is largely contingent on sample size, with larger studies more likely to report
statistical significance. Clinical significance, or minimal clinically important difference
(MCID), may not involve statistical significance (134) and is considered a marker of the
effectiveness of interventions that takes into account the practical importance of
treatment effects (135). Clinical significance also gives meaning to observed changes in
terms of their implications for patient care (136,137). In the case of the current review,
clinical significance may also be used as a comparative metric of treatment effects
between studies.
There are a number of widely accepted assessments of the clinical significance of
change in an intervention, using both anchor-based (i.e., clinical) and distribution-based
(i.e., statistical) assessments (135,137). Anchor-based approaches are methods that relate
change to an external event, rating, or condition, while distribution-based methods link
clinical significance to a statistical parameter of group or individual data (137).
Examination of the yoga and cancer intervention literature reveals minimal reporting of
anchor-based metrics. Given this an anchor-based approach for understanding clinical
significance was not employed in the current overview of the literature.
Commonly reported distribution-based methods include one standard error of the
measurement (1 SEM), 0.5 standard deviation (0.5 SD), and effect sizes. The SEM
provides an index that incorporates both the variation and reliability of a sample on any
measure expressed in the original metric of the measure it describes (138). When
examining the mean difference from pre to post intervention, a value greater than 1 SEM
is considered clinically significant. One SEM has been shown to consistently correlate
29
with anchor-based measures of important difference and with an effect size of 0.5, if the
reported reliability is ≥0.75 (138). Caution must be used in interpreting the SEM, as this
criterion is often predetermined from other studies and may not be derived directly from
the study sample.
Second, the assessment of the clinical significance is often determined using 0.5
SD of the baseline measure as a “rule of thumb” or guideline (139,140). Specifically, prepost mean differences larger than 0.5 SD on that scale are considered clinically
significant. It has been suggested (135) the 0.5 SD is best used as an indicator of
“meaningful difference” versus “minimal difference,” suggesting a change that cannot be
ignored. Third, effect size (ES) is a simple way of quantifying the differences between
two groups, and by association, how large a clinical response is observed (141). Helpful
exploratory criteria are Cohen’s convention of 0.20 for a small effect, 0.50 for a moderate
effect and 0.80 for a large effect (142). Of note, 0.5 SD of the pre-intervention score (in
the case of pre-post within subjects) or 0.5 of control baseline (in the case of betweengroups) can equate with 0.50 ES (provided these measures are calculated using these
same baseline measures as the denominator).
Finally confidence intervals (CI) provide a range of possible scores for a
population parameter. Thus, in the case of the oft reported 95% CI, we can be 95%
confident that the CI includes the population parameter (143). The width of the CI
reflects the precision of the data, with more narrow confidence intervals being more
precise. CIs are not just a surrogate for P-value reported statistical significance, as they
also give us a window into both the size of the difference and precision in estimating the
true difference (144).
30
In all cases, it is important not to use these indices (1 SEM, 0.5 SD, ES, CI) in the
way P-values often are: as arbitrary cut-off values of a study’s given significance (145).
Rather, these values should be used concurrently to triangulate on and describe the range
of one’s findings, their relative magnitude of effect and generalizability. Such a thorough
examination of the clinical significance of published studies on yoga for cancer survivors
has not been completed to date. The purpose of this paper was to clarify how findings
from yoga interventions for cancer survivors should be interpreted by implementing
multiple methods for calculating clinical significance of patient-reported outcomes,
including measures of general quality of life, psychosocial factors and symptoms.
Interpretation of the studies’ findings consisted of examining both within group pre-post
change in both control and pre-post design studies, as well as determining the magnitude
of difference between treatment and control groups in control design studies. Factors that
may influence clinical significance related to the study design, sample, specific
intervention and method of measurement were also documented.
2.2 Methods
2.2.1 Selection of Publications.
This literature review of yoga interventions for cancer survivors, both on and off
treatment, was completed in July 2011. To this end, a systematic search was conducted of
the relevant databases, including PubMed, PsychInfo, Medline, Annotated Bibliography
of Indian Medicine, Cochrane Library, Web of Science, and Google Scholar. Key search
terms included yoga, cancer, survivor, patient, intervention, quality of life, well-being,
clinical significance, important change(s), important difference(s), and/or patient
reported outcomes. The following inclusion criteria were applied to the search: (1) must
31
include a yoga intervention; (2) have a sample of exclusively adult cancer survivors (on
or off treatment); (3) report at least pre- and post-intervention assessments; and (4)
include patient-reported outcomes related to quality of life, psychosocial and/or symptom
outcomes. Publications were excluded if they incorporated intervention components in
addition to yoga (e.g., MBSR, comprehensive lifestyle change programs, physiotherapy),
were published in a non peer-reviewed journal, or were not written in English.
2.2.2 Data Synthesis
The common indices of clinical significance used in the current overview were:
(1) 1 standard error of the measurement (1 SEM); (2) 0.5 standard deviation (0.5 SD);
and (3) both within and between-group difference effect sizes (ES - Hedge’s g) and their
respective confidence intervals (CI). In order to examine the clinical significance of
findings for each yoga intervention study, both pre-post means and the SDs or SEs had to
be available in order to calculate 1 SEM, 0.5 SD, ES and CI. No follow-up time periods
(if included) were analyzed, and only published data were accessed. From the initial 19
eligible publications, a further six were dropped because: (a) SDs or SEs were not
available; or (b) the same data were reported in duplicate publications. Data were
extracted to examine changes pre-post intervention in both the yoga and control groups,
as well as the relative difference between groups in the case of control design studies.
Clinical significance was determined using the following criteria. First, the SEM was
calculated as (pre-intervention SD * √1 – Cronbach’s alpha). If the Cronbach’s alpha was
not cited in the study it was derived from either a clinically relevant population, or from
prior psychometric evaluations of the instrument. Second, 0.5 SD values were calculated
from the baseline SD. Third, ES were calculated using Hedges g for both within-groups
32
effects (post-pre/pooled SD) and between-groups effects (treatment mean differencecontrol mean difference/pooled change SD). These effects were interpreted following
Cohen’s convention of small (0.2), medium (0.5), and large (0.8) (142). Fourth, CIs were
calculated for both treatment and control within-subjects mean differences, betweengroups mean differences and their accompanying ES. ES and CIs were calculated using
Comprehensive Meta-Analysis (Version 2) software (146). Pre-post differences were
evaluated based on their respective 1 SEM, 0.5 SD ES, and CI, while between-group
differences were evaluated based on the control group 1 SEM, 0.5 SD and the betweengroups calculated ES and CI. Finally, given the aforementioned correlation between 1
SEM and an effect size of 0.5 if the reported reliability is ≥ 0.75 (138), the Pearson
product-moment correlation coefficients were calculated to determine the strength of the
linear dependence between the 1 SEM and 0.5 SD. As the primary purpose was to
evaluate patient-reported outcomes in the yoga literature and given the comparatively
small number of studies, no computation of study quality was completed in the current
review.
In addition, given the focus on a number of different indices of clinical
significance, the small number of studies and heterogeneity in terms or type of yoga,
dose-response, length of intervention and outcome measures, average effect sizes across
studies were not calculated nor meta-analyses performed. Although we intended that the
literature search be inclusive of all existing studies, it did not follow other methods
necessary to be considered a systematic review so as not to be redundant with other
recent publications (2,33).
33
2.3 Results
2.3.1 Description of Studies
The initial literature search resulted in a total of 25 publications. A review of
abstracts reduced the final number to 19 publications based on the inclusion and
exclusion criteria. Of the 19 publications reviewed, 14 contained enough data to be
included in this review and are summarized in Tables 2.1 and 2.2. Note that Vadiraja
2009a (80) and Vadirja 2009b (147) are from the same study sample and are therefore
reported as one study. The resulting 13 studies included a variety of yoga interventions
(with respect to type of yoga and duration), cancer types, and timing and content of
assessments. Seven of the 13 studies employed a randomized controlled trial design,
including a control group for comparison to the yoga intervention (n=6 with a waitlist
control group, n=1 with an active control group). Six studies were single-group pre-post
designs. Tables 2.3 through 2.8 details all study results.
2.3.1.1 Study Designs. In the randomized controlled trial studies, mean age
ranged from 46-60 years. Sample size in the treatment group at time 2 ranged from 13-45
participants. Cancer diagnoses were comprised primarily of breast cancer, with one study
focused on lymphoma. Many participants were on active treatment or, in some cases,
three months or more post-treatment. In the case of Moadel et al. (148) we reported on
their secondary analysis that focused on women with breast cancer who were not
receiving chemotherapy. Yoga classes were 60-90 minutes in length, and programs lasted
6 to 26 weeks. Yoga styles included Hatha, Integral, Iyengar, Tibetan, Viniyoga, and
Vivekananda. The majority of control designs employed a waitlist control group except
34
the Vadiraja study (80,147), which utilized an active supportive therapy with education
control group.
In the single group design studies, mean age ranged from 41-59 years. Sample
size at time 2 ranged from 10-43 participants. Cancer diagnoses were comprised
primarily of breast cancer, with one study focused on ovarian cancer. Many participants
were on active treatment or, in some cases, six months or more post-treatment. Yoga
classes were 60-90 minutes in length and programs lasted 6 to 10 weeks. Yoga styles
included Integral, Iyengar, and “classical” yoga.
2.3.1.2 Instruments Reviewed. A total of 18 different instruments, assessing
patient-reported outcomes in health-related quality of life (HRQL; six instruments),
psychosocial (eight instruments) and symptom domains (four instruments) were
reviewed. We acknowledge there is some overlap in HRQL and psychosocial constructs.
For the sake of simplicity and parsimony, we left HRQL subscales in the HRQL category
even if they capture psychosocial constructs such as social or emotional well-being.
Further, in order to minimize the number of categories of patient-reported outcomes, we
included spiritual well-being in the psychosocial category. Tables 2.3 through 2.8 present
a summary of the clinical significance for the patient-reported outcomes in the 13
reviewed studies and is divided into control design and pre-post single group design
studies. Each table further distinguishes between patient-reported outcome categories –
HRQL, Psychosocial and Symptom variables.
2.3.2 Quality of Life Outcomes
2.3.2.1 Overall HRQL. Two control design studies, Danhauer et al. (79) and
Culos-Reed et al. (149) met the 1 SEM and 0.5 SD criteria both pre-post and between
35
yoga intervention and waitlist control. Medium between-group ES, ranged from 0.49
[95% CI -0.25, 1.24; p=NS] (79) to 0.67 [95% CI 0.01, 1.32; p<.05] (149). Moadel et al.
(148), while showing no evidence of clinically significant change pre-post, met the 1
SEM criteria with an overall small ES of 0.43 [95% CI -0.05, 0.91; p=NS], reflecting
improvement in overall HRQL in the yoga group versus a worsening in the control group.
Littman et al. (150) showed no significant differences either pre-post or between-groups.
In single group design studies (151-156) clinically significant outcomes ranged from a
small ES of 0.39 [95% CI -0.08, 0.86; p=NS] meeting the 1 SEM criteria (155), to a large
ES of 0.81 [95% CI 0.14, 1.48; p<.05] (154), that met both 1 SEM and 0.5 SD criteria.
Interestingly, overall HRQL in Danhauer et al. (152), while statistically significant
(p<.05) with a small ES of 0.32 [95% CI 0.02, 0.63], met neither the 1 SEM nor 0.5 SD
criteria.
2.3.2.2 Physical HRQL. Only two control design studies, Chandwani et al. and
Danhauer et al. (79,157), showed evidence of clinically significant differences. While
both studies did not exhibit clinically significant pre-post changes, both studies showed
clinically significant effects between groups that met the 1 SEM criteria, with small ES,
ranging from 0.36 ES [95% CI -0.38, 1.10; p=NS] (79) to 0.46 ES [95% CI -0.06, 0.97;
p=NS] (157). Interestingly, although there was clinically significant change on the FACT
physical wellbeing subscale in Danhauer et al. (79), there was no significant change on
the SF-12 physical wellbeing subscale. Several other control design studies (80,148,150)
exhibited no clinically significant differences in physical HRQL, either pre-post or
between groups. Within the single-group design studies, only the Ülger and Yağli study
(156) showed a large ES of -0.92 [95% CI -1.43, -0.41; p<.001] and met both the 1 SEM
36
and 0.5 SD criteria, indicating significant improvements in physical HRQL pre-post yoga
intervention. Speed-Andrews et al. (155) met the 1 SEM criteria with a small ES of 0.32
[95% CI -0.14, 0.79; p=NS]. Interestingly, Danhauer et al. (152) met neither 1 SEM nor
0.5 SD but was statistically significant (p<.05) with an effect size of 0.38 [95% CI 0.07,
0.68] on the FACT physical wellbeing subscale. There was no significant change on the
SF-12 physical wellbeing subscale.
2.3.2.3 Mental HRQL. Large effects were seen in control group designs for
Danhauer et al. (79) with pre-post and between-group mean differences meeting both 1
SEM and 0.5 SD criteria, with a between group ES of 1.00 [95% CI 0.22, 1.78; p<.01].
Vadiraja et al. (80) did not meet the 1 SEM or 0.5 SD criteria but reported a statistically
significant (p<.05) ES of 0.30 [95% CI 0.00, 0.61] pre-post yoga intervention. There
were no clinically significant differences between groups. There were also no significant
differences either pre-post or between groups in Chandwani et al. (157). In single group
designs, Danhauer et al. (152) pre-post differences did not meet the 1 SEM or 0.5 SD
criteria but reported a statistically significant (p<.01) ES of 0.41 [95% CI 0.11, 0.72].
Speed Andrews et al. (155) demonstrated a small clinically significant difference meeting
the 1 SEM criteria with an effect size of 0.45 [95% CI -0.03, 0.93; p=NS].
2.3.2.4 Emotional HRQL. Values ranged from meeting both the 1 SEM and 0.5
SD criteria pre-post and between groups, with ES from 0.53 [95% CI 0.07, 0.99; p<.05]
(80), to 0.61[95% CI -0.15, 1.35; p=NS] (79), between groups. There were no clinically
significant differences pre-post for either Moadel et al. (148) or Culos-Reed et al. (149),
however both exhibited between-group differences. Culos-Reed et al. (149) exhibited a
small ES of 0.48 [95% CI -0.16, 1.13; p=NS] between groups that met both the 1 SEM
37
and 0.5 SD criteria. Moadel et al. (148) met neither 1 SEM nor 0.5 SD criteria but
demonstrated a small statistically significant (p<.01) pre-post ES of 0.40 [95% CI 0.10,
0.70]. There were no differences between groups. Littman et al. (150) demonstrated no
clinically significant differences either pre-post or between groups. Within single subject
designs, Ülger and Yağli (156) reported large pre-post differences that met both 1 SEM
and 0.5 SD with an effect size of -1.15 [95% CI -1.70, -0.60; p<.001]. There was no
clinically significant differences pre-post yoga intervention for either Danhauer et al.
(152) or Speed-Andrews et al. (155).
2.3.2.5 Social HRQL. The majority of control design studies (79,80,150) report
no clinically significant differences either pre-post or between yoga and waitlist control
groups. Interestingly, Moadel et al. (148) demonstrated a moderate, clinically significant
decline in the waitlist control group, with an ES of -0.60 [95% CI -1.00, -0.19; p<.01].
This led to a moderate between group difference of 0.60 ES [95% CI 0.11, 1.09; p<.05],
which met both the 1 SEM and 0.5 SD criteria, indicating a relative improvement in
social HRQL for the yoga group versus a worsening of the waitlist control group. Within
the single group designs, Ülger and Yağli (156) demonstrated a moderate clinically
significant difference of -0.74 ES [95% CI -1.22, -0.26; p<.01], meeting both the 1 SEM
and 0.5 SD criteria, while Danhauer et al. (152) and Speed-Andrews et al. (155)
demonstrated no clinically significant differences in social HRQL pre-post yoga
intervention.
2.3.2.6 Functional HRQL. Only Danhauer et al. (79) reported pre-post
improvement in Functional HRQL, with a moderate between-group ES of 0.58 [95% CI 0.17, 1.34; p=NS] that met both 1 SEM and 0.5 SD criteria. Littman, Moadel, and
38
Vadiraja, (80,148,150) indicated no clinically significant differences either pre-post or
between yoga and waitlist control groups. Similarly no clinically significant differences
are reported pre-post within single group designs (152,155).
2.3.3 Psychosocial Outcomes
2.3.3.1 Depression. Clinically significant differences ranged from small, with a
between-groups ES of -0.39 [95% CI -1.04, 0.25; p=NS], that met the 1 SEM criteria
(149) to medium, with a between groups ES of -0.69 [95% CI -1.45, 0.06; p=NS],
meeting both 1 SEM and 0.5 SD criteria (79). Interestingly in the case of Vadiraja et al.
(147) large clinically significant differences within the yoga group, -0.89 ES [95% CI 1.25, -0.54; p<.001], and small clinically significant differences within the supportive
therapy control group, -0.40 ES [95% CI -0.74, -0.05; p<.05], indicated significant
depression reduction in both groups. This led to a medium overall ES, -0.52 ES [95% CI
-0.98, -0.06; p<.05] that met both the 1 SEM and 0.5 SD criteria, favoring overall
depression reduction in the yoga group relative to the supportive therapy control. Within
the Chandwani et al. study (157), small clinically significant differences in the yoga
group, -0.48 ES [95% CI -0.86, -0.09; p<.05], and small clinically significant differences
in the waitlist control group, -0.32 ES [95% CI -0.67, 0.03; p=NS], also indicated
significant depression reduction in both groups. However, there were no overall clinically
significant differences in depression reduction between yoga and waitlist control groups.
Within the Cohen et al. (158) study there were no clinically significant differences in
depression reduction pre-post or between groups. Within single group designs, ES ranged
from small, -0.36 ES [95% CI -0.66, -0.05; p<.05] meeting the 1 SEM criteria (152), to
large, -1.00 ES [95% CI -1.69, -0.32; p<.01] meeting both the 1 SEM and 0.5 SD (151).
39
In the case of Speed-Andrews et al. (155) no clinically significant differences in
depression were indicated pre-post yoga intervention.
2.3.3.2 Anxiety. Within the control designs, Vadiraja et al. (147) indicated
moderate clinically significant differences, with an overall between-groups ES of -0.51
[95% CI -0.97, -0.05; p<.05] that met the 1 SEM criteria. Chandwani et al. (157)
indicated moderate clinically significant differences within the yoga group, -0.63 ES
[95% CI -1.04, -0.23; p<.01] and small clinically significant differences in the waitlist
control group, ES -0.20 [95% CI -0.55, 0.15; p=NS], indicating anxiety reduction in both
groups. Overall, small clinically significant differences between yoga and waitlist
control, with an ES of -0.46 [95% CI -0.98, 0.05; p=NS] that met both the 1 SEM and 0.5
SD criteria, suggest relative overall anxiety reduction in the yoga group relative to the
waitlist control. Interestingly, within the Cohen et al. study (158), small clinically
significant differences, with an ES of -0.30 [95% CI -0.80, 0.21; p=NS] that met the 1
SEM criteria; indicate significant overall anxiety reduction in the waitlist control versus
the yoga group. Within single group designs Ülger and Yağli (156) demonstrated large
clinically significant reductions in anxiety, -1.51 ES [95% CI -2.14, -0.88; p<.001] that
met both the 1 SEM and 0.5 SD criteria. Danhauer et al. (152) indicated small clinically
significant reductions in anxiety, with an ES of -0.22 [95% CI -0.52, 0.08; p=NS] that
met the 1 SEM criteria. No clinically significant differences were reported in SpeedAndrews et al. (155).
2.3.3.3 Positive Affect. Danhauer et al. (79) demonstrated a small clinically
significant effect of 0.45 ES [95% CI -0.29, 1.19; p=NS] that met the 1 SEM criteria
between yoga and waitlist control groups. Vadiraja et al. (80)reported a moderate
40
clinically significant effect pre-post yoga intervention of 0.52 ES [95% CI 0.20, 0.84].
However, although statistically significant (p<.001), the between-group difference met
neither the 1 SEM nor 0.5 SD criteria. Within the single group designs, there were no
significant differences in positive affect pre-post yoga intervention.
2.3.3.4 Negative Affect. Within the Danhauer et al. study (79), moderate
clinically significant differences between groups, -0.73 ES [95% CI -1.49, 0.03; p=NS]
meeting both the 1 SEM and 0.5 SD criteria, indicated significant overall reduction in
negative affect in the yoga group. Within the Vadiraja et al. study (80), large clinically
significant differences within the yoga group, -0.86, ES [95% CI -1.21, -0.51; p<.001]
and small differences within the supportive therapy group, -0.33 ES [95% CI -0.67, 0.01;
p=NS] suggested there was a clinically significant reduction of negative affect in both
groups. However, moderate clinically significant differences between groups, -0.57 ES
[95% CI -1.03, -0.11; p<.05], that met both the 1 SEM and 0.5 SD criteria, suggest a
significant overall reduction of negative effect in the yoga group relative to the
supportive therapy control group. In the single group study designs, Danhauer et al. (152)
met neither the 1 SEM nor 0.5 SD criteria but did exhibit a statistically significant
(p<.05) ES of -0.35 [95% CI - 0.65, -0.04].
2.3.3.5 Spiritual Well-Being. In Danhauer et al. (79) moderate clinically
significant differences, 0.56 ES [95% CI -0.19, 1.30; p=NS] meeting both the 1 SEM and
0.5 SD criteria, indicated an increase in spiritual well-being in the yoga versus waitlist
control group. Within the Moadel et al. study (148) there were no clinically significant
differences in spiritual well-being pre-post or between groups. Within the single group
designs moderate clinically significant differences, 0.60 ES [95% CI 0.16, 1.04; p<.01]
41
were reported pre-post yoga intervention (153). There were no clinically significant
differences in spiritual well-being pre-post yoga intervention in Danhauer et al. (152).
3.4 Symptom Outcomes
2.3.4.1 Fatigue. Within the control group designs (79,148,150,157,158) clinically
significant differences between the yoga intervention and control groups ranged from
small ES, -0.17 ES [95% CI -0.68, 0.34; p=NS], meeting the 1 SEM criteria (157) to
medium ES, 0.71 ES [95% CI -0.04, 1.47; p=NS], meeting both the 1 SEM and 0.5 SD
criteria (79). In the case of Cohen et al. (158) there were no clinically significant
differences in fatigue pre-post or between yoga and control groups. Within the single
group designs, effect sizes ranged from small, 0.22 ES [95% CI -0.08, 0.52; p=NS]
meeting the 1 SEM criteria (152), to very large, -2.34 ES [95% CI -3.47 to -1.22; p<.001]
meeting both the 1 SEM and 0.5 criteria (151).
2.3.4.2 Sleep. In the case of Danhauer et al. (79) despite small clinically
significant differences pre-post, ES -0.46 ES [95% CI -1.00, 0.08; p=NS] there were no
overall clinically significant differences in sleep between the yoga and waitlist control
groups. There were no clinically significant differences in sleep pre-post or between
groups in the Chandwani or Cohen studies (157,158). In the single group designs, there
was a large effect, -0.88 ES [95% CI -1.38, -0.38; p<.001] meeting both 1 SEM and 0.5
SD requirements, for sleep pre-post yoga intervention (156). In the case of Bower et al.
(151) there were no clinically significant differences in sleep pre-post yoga intervention.
3.5 Clinical Significance Criteria
2.3.5.1 Consistency between SEM and SD. Within the control group designs, the
correlation between the 1 SEM and 0.5 SD criteria for HRQL indices was r=.92, p<.01
42
for the treatment group and r=.93, p<.01 for the control group. For the single group
designs the correlation between 1 SEM and 0.5 SD was r=.92, p<.01, indicating these
criteria are highly correlated. For psychosocial indices within the control group designs,
the overall correlation between the 1 SEM and 0.5 SD criteria was r=.74, p<.01for the
treatment group and r=.83, p<.01 for the control group. For the single group designs the
correlation between 1 SEM and 0.5 SD was r=.80, p<.01, indicating these criteria are also
highly correlated. Similarly, for the symptom indices, within the control group designs,
the overall correlation between the 1 SEM and 0.5 SD criteria was r=.91, p<.01for the
treatment group and r=.91, p<.01 for the control group. Finally, for the single group
designs the correlation between 1 SEM and 0.5 SD was r=.97, p<.05, indicating these
criteria are highly correlated.
2.4 Discussion
The yoga and cancer literature is rapidly growing. This literature is characterized
by studies published with small sample sizes and variability in the type and length of
interventions, populations studied (e.g., cancer type, time during the treatment
continuum), and measures used to assess the patient-reported outcomes of interest. Thus,
to accurately characterize the results of the current literature in this area, reporting
multiple indicators of clinical significance is of great value (159). Studies evaluated for
clinical significance show some consistency across this research, demonstrating the
positive impact of yoga on quality of life, psychosocial and, to a lesser degree, physical
symptom indices.
Since this area of research is in its infancy, it is helpful to utilize multiple criteria
of clinical significance to explore the clinical efficacy of results from each study. The use
43
of not only ES and CIs, but also validated metrics such as the 1 SEM and 0.5 SD criteria
is recommended. A cursory look at the summary tables would indicate that to examine
any of these studies by one criterion alone, particularly in the case of P-values, much
information and insight into the results would be lost. Based on the current review, we
see emerging beneficial findings from yoga interventions in the areas of HRQL,
including overall HRQL and its mental and emotional domains, and in psychosocial
measures in the areas of anxiety, depression, negative affect and spiritual well-being.
There may also be some preliminary evidence for yoga’s role in fatigue. Findings for the
role of yoga in physical, functional and social domains of HRQL remain far more
inconclusive, as is the role of yoga in positive affect and sleep indices for cancer
survivors. Thus, considering clinical significance indicates stronger support than previous
reviews of the literature for the preliminary efficacy of yoga for improving overall HRQL
and its mental and emotional domains, in addition to psychosocial outcomes (2,33).
The present review highlights that yoga for cancer survivors results in a number
of clinically significant improvements in select patient-reported outcomes. Specifically,
multiple criteria, including a mean difference from pre- to post-intervention greater than
or equal to 1 SEM and/or 0.5 SD, and the respective ES and CI, were met for quality of
life and psychosocial outcomes (e.g., anxiety, depression, positive and negative affect,
spiritual well-being), and for some limited symptom outcomes (e.g., fatigue, sleep).
These results suggest that indices vary in their sensitivity / conservatism for reporting
clinical significance. For example, in cases where research participants patient-reported
outcome pre-scores are sub-clinical or it has been established that smaller changes in
patient-reported outcomes would be beneficial (e.g., palliation), using the mean
44
difference criterion of greater than 1 SEM may be a more pragmatic index of clinically
significant change. In addition, researchers may want to employ the respective smaller ES
with the SEM when smaller changes would be considered meaningful.
2.4.1 Baseline Cut-offs Reported in Literature
One consideration for the discrepancies between clinical significance and
statistical significance found for HRQL and psychosocial patient-reported outcomes may
be participant baseline characteristics. Baseline health status, as reflected in patientreported outcome scores, has been shown to affect the overall responsiveness of an
instrument (160). For example, if participants are several months post treatment, they
may report pre-intervention scores that are relatively high (positive). Thus, there is less
“room” for improvement and a ceiling effect is encountered on a given outcome. The
patient cannot score any higher and so a given level of change necessary for clinical
significance is not obtained (137). Therefore, improvements in patient-reported outcomes
must be contextualized within participant baselines scores.
While beyond the scope of this manuscript, the baseline values of all HRQL,
psychosocial, and symptom measures varied for each study in comparison to established
clinical cut-offs, where available. Clinical cut-off values are established to distinguish
between people who have a clinically important level of the construct and to enable
comparison between studies. Thus studies that had more room for improvement at
baseline, which can be judged in comparison to clinical cut-off values, may have resulted
in higher levels of clinical significance.
45
2.4.2 MIDs Reported in Literature
In addition to the 1 SEM, 0.5 SD and ES measures reported in the results, for
some instruments, MIDs currently exist in the literature. Thus it may also be of value to
compare current results to literature-reported MIDs. Utilizing a combination of
distribution-based methods and MIDs reported in the literature, as they are available,
allows for unique insights into both the interventions and study members. While
literature-based MIDs may allow for additional ease in comparing changes across studies,
they are developed based on a large cross-section of studies and may not always be
appropriate given any one study (159).
2.4.3 Limitations & Future Research
The current review has some limitations. First, it is important to note calculations
of clinical significance do not account for elements that can be controlled by the original
study design. Differences between the studies – type of yoga, duration, assessment
periods, number of assessments, type of control group (waitlist vs. active comparison
control), sample size, and cancer type – are all issues that should be considered in the
interpretation of the results. For example, some studies may show clinically significant
changes as a function of a longer duration (i.e., a time effect improvement in patientreported outcomes) rather than a direct benefit of yoga. In addition, small sample sizes
may also miss important clinical differences and mask significant improvement following
an intervention. In contrast, studies with large sample sizes may produce significant
findings that are clinically meaningless (161). Future studies will be strengthened by the
recruitment of larger, more homogenous samples to minimize this variability and lead to
more consistent results. In addition, the research clearly has relied primarily on
46
participation from breast cancer survivors, and minimal published work on yoga in other
groups of cancer survivors exists to date.
The relative dearth of symptom-focused patient-reported outcomes illustrates a
gap in the evaluation of the potential physical effects of yoga for cancer survivors.
Whereas it is certainly important that yoga consistently demonstrates a positive influence
on quality of life and psychosocial outcomes, research evaluation may be missing an
important component of the potential benefits that may be gained in symptom alleviation,
be it fatigue, sleep disturbance or pain. Future research that focuses more in-depth on
assessment of symptom patient-reported outcomes, in addition to fatigue and sleep, is
warranted.
Comparing yoga to other types of interventions, either physical activity or
psychosocial, may elucidate comparable effects or unique outcomes of yoga on patientreported outcomes in cancer survivors. Notably, yoga was not originally intended to
target specific outcomes. Thus, choosing which of the many yoga practices to implement
(e.g., movement, breath, meditation) and, based in part on these decisions, which
psychophysiological outcomes to assess, will be an ongoing challenge in the attempt to
explore this holistic ancient practice with contemporary methods of scientific exploration
(48).
Within this review paper a total of 18 different instruments were used to assess
patient-reported outcomes, including six instruments for HRQL alone. Recent research
has advocated for a more coordinated use of measures to increase validity of constructs
measured as well as to ensure the reliability of comparisons (162). In this capacity, recent
work by the Patient Reported Outcomes Measurement Information System (PROMIS)
47
cooperative group provides new direction in collectively and parsimoniously measuring
these phenomena (163). Independent of the PROMIS initiative, the literature on yoga for
cancer survivors is beginning to use some of the same measures within multiple
interventions. This practice of consistent use of measures is valuable for furthering the
evidence on yoga for cancer survivors. For example, recent studies have found clinically
significant changes using the FACT-G and subscales, FACT-F, FACIT-Sp (79,153,155).
Thus, we now have the benefit of inferring that it was not likely the measure alone, but
perhaps other intervention-dependent factors, that influenced the results. In this capacity,
we also strongly recommend the use of disease-specific instruments in future research.
Identifying mechanisms that explain how yoga leads to clinically significant
outcomes is another next step in understanding the ability of yoga to target specific
desired outcomes. Little research attention has been paid to the psychophysiological
mechanisms by which benefits are accrued via yoga practice. In this capacity, there have
been recommendations to integrate patient-reported outcomes with biomedical endpoints
as a means of better describing the complexity of these measures (62). Before yoga can
be broadly applied within oncology, both a strong theoretical understanding of how yoga
practice causes change and carefully designed and executed research that convincingly
evaluates not only the efficacy of yoga in clinical settings but also posits potential
mechanisms of action underlying these interventions are required (164).
Finally, since the timing of the assessment of a patient-reported outcome may
influence clinical significance, it is important for future studies of yoga to include
measures that look at change before and after a single yoga practice session. Assessing
48
these more acute effects of yoga would allow for documentation of the potential
immediate impact of yoga on outcomes of interest (165).
2.4.4 Conclusion
There is a need across the research literature to assess a variety of indices that
speak to a given study’s clinical significance, a metric relevant to researchers, clinicians,
and patients alike. The current analysis supports preliminary evidence of the clinical
significance of yoga for improving quality of life, psychosocial, and limited symptom
outcomes in cancer survivors. It behooves researchers to take these preliminary findings
as a starting point to begin to explore, report and revise these purported indices of clinical
significance further (139). Identifying the clinical significance in studies of yoga for
cancer survivors adds further description to the existing literature and highlights the
promising intervention benefits of a yoga intervention. While this study focused
specifically on one type of intervention in a defined group (i.e., cancer survivors), there is
the broader issue of the need to examine and present clinical significance data for other
types of intervention studies in order to ensure that “significant” findings are truly
meaningful to people, impacting various health outcomes and behaviors that are
important to them.
Acknowledgements
This work was supported by Alberta Health Services, University of Calgary
Research Grant funding, the Wake Forest School of Medicine Translational Science
Institute, and the National Cancer Institute (R25 CA122061-01A1). The content is solely
the responsibility of the authors and does not necessarily represent the official views of
the National Cancer Institute or the National Institutes of Health. Michael J. Mackenzie
49
acknowledges Social Sciences and Humanities Research Council of Canada (SSHRC)
and Alberta Innovates-Health Solutions (AIHS) for funding of his doctoral program of
research.
Author
Chandwani et
al., 2010 (157)
Cohen et al.,
2004 (158)
Culos-Reed et
al., 2006 (149)
Danhauer et
al., 2009 (79)
Sample
Cancer Type
and Treatment
Yoga Intervention
Control
Group
Outcome Measures
Reported in Current
Review
Quality of Life
-SF-36
Psychosocial
-CES-D, STAI
Symptom
-BFI, PSQI
Treatment: T1=30, T2=27
Control: T1=31, T2=31
Age range: 31-68
Mean age = 53
Breast (Stages 0-3)
Chemotherapy (n=47)
Radiation therapy (n=61)
60-minute classes twice
weekly for 6 weeks
Type: Vivekananda Yoga
Wait-list
control
Treatment: T1=19, T2=16
Control: T1=19, T2=14
Age range: NR
Mean age = 51
Lymphoma
(Stages I - IV)
Previous 12 months
receiving chemotherapy
Weekly classes for 7 weeks
(class length not reported)
Type: Tibetan Yoga
Wait-list
control
Treatment: T1=20, T2=18
Control: T1=18, T2=18
Age range: NR
Mean age = 51
Breast (n=32),
Other (n=6)
≥ 3 months post-treatment
75-minute weekly classes for
7 weeks
Type: Iyengar Yoga
Wait-list
control
Treatment: T1=22, T2=13
Control: T1=22, T2=14
Age range: 38-79
Mean age = 55
Breast (Stages I-IV)
Radiation therapy
(n=9)
Chemotherapy (n=11)
75-minute weekly classes for
10 weeks
Type: Integral Yoga
Wait-list
control
Psychosocial
-CES-D, STAI
Symptom
-BFI, PSQI
Quality of Life
-EORTC QLQ-C30
Psychosocial
-POMS
Quality of Life
-FACT-B, SF-12
Psychosocial
-CES-D, PANAS,
FACIT-Sp
Symptom
-FACT-F, PSQI
Table 2.1 Studies Reviewed from the Yoga-Cancer Literature – Control Design
50
50
Author
Littman et al.,
2011 (150)
Moadel et al.,
2007 (148)
Vadiraja et al.,
2009a & b
(80,147)
Sample
Treatment: T1=32, T2=27
Control: T1=31, T2=27
Age range: 33-74
Mean age = 60
Treatment: T1=45, T2=45
Control: T1=26, T2=26
Age range: 28-75
Mean age = 54
Treatment: T1=44, T2=42
Control: T1=44, T2=33
Age range: 30-70
Mean age = 46
Cancer Type
and Treatment
Yoga Intervention
Breast (Stages 0-III)
> 3 months postchemotherapy and/or postradiation
75-minute classes five times
per week for 26 weeks
(combination in-facility and
home-based practice)
Type: Viniyoga
Breast (Stages I-III)
Surgery (n=95)
Radiation Therapy
(n=10)
Chemotherapy
(n=27)
Anti-estrogen Therapy
(n=30)
90-minute weekly classes for
12 weeks. Home practice
encouraged
Type: Hatha Yoga
Breast (Stages II-III)
Radiation therapy (n=88)
Minimum of 3 in-person 60minute individual yoga
sessions per week for 6
weeks. Home practice
encouraged.
Type: Integrated Yoga
Program
Control
Group
Wait-list
control
Wait-list
control
3-4
sessions of
brief
supportive
therapy w/
education
Outcome Measures
Reported in Current
Review
Quality of Life
-FACT-G
Symptom
-FACT-F
Quality of Life
-FACT-G
Psychosocial
-FACIT-Sp
Symptom
-FACT-F
Quality of Life
-EORTC QoL C30
Psychosocial
-PANAS, HADS
Table 2.1 – continued
51
51
Author
Bower et al.,
2011 (151)
Sample
T1=12, T2=11
Age range: 46-65
Mean age = 54
Danhauer et al.,
2008 (152)
T1=51, T2=43
Age range: 34-82
Mean age = 59
Duncan et al.,
2008 (153)
T1=24, T2=22
Age range: NR
Mean age = 49
Galantino et al.,
2011 (154)
T1=10, T2=10
Age range: 50-71
Median age = 58
Cancer Type
and Treatment
Breast (Stages 0-2)
All had completed radiation therapy
or radiation therapy +
chemotherapy
Yoga Intervention
90-minute classes twice weekly for 12
weeks
Type: Iyengar Yoga
Outcome Measures
Reported in Current
Review
Quality of Life
-SF-36
Psychosocial
-BDI-II
Symptom
-PSQI, FSI
75-minute weekly classes for 10 weeks
Type: Integral Yoga
Quality of Life
-FACT-G, SF-12
Psychosocial
-CES-D, FACIT-Sp,
PANAS, STAI
Symptom
-FACT-F
Breast (42%)
Gynecologic (16.6%)
Lymphoma (12.5%)
Currently on or < 6 months post
treatment
90 minute weekly classes for 10 weeks
Type: Iyengar Yoga
Quality of Life
-FACT-G
Psychosocial
-FACIT-Sp
Breast (Stages I-III)
> 4 weeks post-chemotherapy
and/or post-radiation, all women
taking aromatase inhibitors
90-minute classes twice weekly for 8
weeks, home practice encouraged
Type: Iyengar-inspired
Quality of Life
-FACT-B
Ovarian (n=37)
Breast (n=14)
Radiation therapy (n=5)
Chemotherapy (n=29)
Table 2.2 Studies Reviewed from the Yoga-Cancer Literature – Single Group Design
52
52
Author
Sample
Cancer Type
and Treatment
Yoga Intervention
Outcome Measures
Reported in Current
Review
Speed-Andrews
et al., 2010 (155)
T1=17, T2=17
Age range: NR
Mean age = 54
Breast (Stages I-III)
Surgery (n = 23)
Radiation therapy (n = 19)
Chemotherapy (n = 18)
Hormone therapy (n = 10)
Two different sessions; 90 minute
classes twice weekly for 6 weeks and
90 minute classes 22 times over 12
weeks
Type: Iyengar Yoga
Quality of Life
-FACT-B, SF-36
Psychosocial
-CES-D, STAI
Ülger &Yağli,
2010 (156)
T1=20, T2=20
Age range: 30-50
Mean age = 41
Breast
> 6 months post-chemotherapy
60 minute weekly classes for 8 weeks
Type: Not specified
(“Classical Yoga”)
Quality of Life
-NHP
Psychosocial
-STAI
Abbreviations: T1, baseline; T2, follow-up; SF-36 & SF-12, Medical Outcomes Study Short-Form Health Survey; EORTC QLQ C30, European
Organization for Research and Treatment of Cancer- Quality of Life C30; FACT-B , Functional Assessment of Cancer Therapy –Breast; FACT-G,
Functional Assessment of Cancer Therapy –General; NHP, Nottingham Health Profile; BDI, Beck Depression Inventory; CES-D 20, Center for
Epidemiologic Studies Depression Scale – 20 Item; STAI, State/Trait Anxiety Inventory; POMS, Profile of Mood States; PANAS, Positive &
Negative Affect Schedule; FACIT-Sp, Functional Assessment of Chronic Illness Therapy-Spiritual Well-Being; HADS, Hospital Anxiety &
Depression Scale; BFI, Brief Fatigue Inventory; PSQI, Pittsburgh Sleep Quality Index; FSI, Fatigue Symptom Inventory; FACT-F, Functional
Assessment of Cancer Therapy-Fatigue. NR=Not Reported.
Table 2.2 – continued
53
53
Study
Outcome
measure
SF-36 PCS
Chandwani
(Physical)
et al., 2010
SF-36 MCS
(157)
(Mental)
EORTC-QoL
Culos-Reed
(Overall)
et al., 2006
EORTC-Emo
(149)
(Emotional)
FACT-B
(Overall)
FACT-SWB
(Social)
FACT-FWB
(Functional)
Danhauer et
FACT-EWB
al., 2009
(Emotional)
(79)
FACT-PWB
(Physical)
SF-12 PCS
(Physical)
SF-12 MCS
(Mental)
FACT-B
(Overall)
FACT-SWB
(Social)
Littman et
FACT-FWB
al., 2011
(Functional)
(150)
FACT-EWB
(Emotional)
FACT-PWB
(Physical)
SEM (Baseline
Internal
SD x √1 –
consistency
Cronbach α)
(Cronbach α)
Baseline 0.5 SD
Yoga
Control
Yoga
Control
.90
2.60
3.40
4.10
5.30
.90
3.80
3.50
6.00
5.60
.89
8.50
4.26
12.82
6.43
.80
9.01
6.64
10.07
7.43
.90
6.30
7.70
10.00
12.20
.69
2.60
3.00
2.60
2.70
.86
2.60
3.10
2.90
3.50
.69
2.00
2.70
2.00
2.60
.81
3.00
2.20
3.50
2.60
.86
4.50
3.80
6.1 0
5.10
.81
4.40
4.50
5.10
5.10
.90
2.97
4.49
4.70
7.10
.69
2.84
3.01
2.60
2.70
.86
1.23
1.46
1.70
2.60
.69
1.56
1.45
1.40
1.30
.81
1.00
1.70
1.20
2.00
Baseline Mean
(SD)
Yoga
41.50
(8.20)
47.80
(12.00)
64.58
(25.63)
79.58
(20.14)
104.90
(19.90)
23.30
(4.70)
18.70
(5.70)
18.10
(3.90)
19.70
(7.00)
42.70
(12.10)
43.40
(10.10)
89.00
(9.40)
21.70
(5.10)
22.70
(3.30)
19.80
(2.80)
24.70
(2.30)
Control
41.8
(10.6)
48.0
(11.1)
62.04
(12.85)
75.00
(14.85)
101.1
(24.4)
21.4
(5.4)
18.0
(6.9)
18.5
(5.2)
20.7
(5.1)
40.6
(10.1)
49.9
(10.2)
87.8
(14.2)
21.7
(5.4)
21.6
(5.1)
20.3
(2.6)
24.2
(3.9)
Within-Groups Mean
Difference (95% CI)
Yoga
1.60
[-1.74, 4.94]
1.90
[-2.16,5.96]
13.66
[2.85,24.47]ab
4.22
[-4.05,12.49]
9.90
[-0.71,20.51]a
-0.20
[-2.84,2.44]
3.20
[0.33,6.07]ab
2.70
[0.74,4.66]ab
2.80
[-1.18,6.78]
2.10
[-4.56,8.76]
8.80
[3.97,13.63]ab
1.30
[-2.58,5.18]
0.40
[-1.51,2.31]
-0.10
[-1.47,1.27]
0.50
[-0.84,1.84]
0.70
[-0.08,1.48]
Control
-2.80
[-6.35, 0.75]
1.90
[-1.76,5.56]
0.46
[-6.16,7.08]
-4.17
[-11.57,3.23]
-2.70
[-17.80,12.40]
-1.00
[-4.26,2.26]
-0.60
[-4.38,3.18]
-0.30
[-3.29,2.69]
0.40
[-2.44,3.24]
2.10
[-3.68,7.88]
-2.40
[-8.90, 4.10]
-0.10
[-5.61,5.41]
-0.80
[-2.96,1.40]
0.10
[-1.75,1.95]
0.50
[-0.59,1.59]
0.10
[-1.47,1.67]
Within-Groups
Effect Size - Hedges’ g
(95% CI)
Yoga
0.18
[-0.19,0.54]
0.17
[-0.19,0.54]
0.56
[0.08,1.04]*
0.23
[-0.22,0.67]
0.48
[-0.06,1.02]
-0.04
[-0.55,0.47]
0.57
[0.01,1.12]*
0.70
[0.13,1.28]*
0.36
[-0.17,0.89]
0.16
[-0.35,0.67]
0.93
[0.30,1.55]**
0.12
[-0.24,0.49]
0.08
[-0.29,0.44]
-0.03
[-0.39,0.34]
0.14
[-0.23,0.51]
0.33
[-0.05,0.71]
Control
-0.27
[-0.62,0.08]
0.18
[-0.17,0.52]
0.03
[-0.41,0.47]
-0.25
[-0.70,0.20]
-0.09
[-0.58,0.41]
-0.15
[-0.65, 0.35]
-0.08
[-0.57,0.42]
-0.05
[-0.54,0.44]
0.07
[-0.42,0.56]
0.18
[-0.32,0.68]
-0.18
[-0.68,0.31]
-0.01
[-0.37,0.36]
-0.14
[-0.50,0.23]
0.02
[-0.35,0.39]
0.17
[-0.20,0.54]
0.02
[-0.34,0.39]
Between Groups
Mean Difference
(95% CI)
4.40
[-0.52, 9.32]c
0.00
[-5.45,5.45]
13.20
[0.52,25.88]cd
8.39
[-2.71,19.49]cd
12.6
[-6.12,31.32]cd
0.80
[-3.43,5.03]
3.80
[-1.00,8.60]cd
3.00
[-0.63,6.63]cd
2.40
[-2.44,7.24]c
0.00
[-8.79,8.79]
11.20
[3.01,19.39]cd
1.40
[-5.34,8.14]
1.20
[-1.68,4.08]
0.20
[-2.11,2.51]
0.00
[-1.73,1.73]
0.60
[-1.16,2.36]
Between Groups
Effect size
– Hedges’ g
(95% CI)
0.46
[-0.06,0.97]
0.00
[-0.51,0.51]
0.67
[0.01,1.32]*
0.48
[-0.16,1.13]
0.49
[-0.25,1.24]
0.14
[-0.60,0.87]
0.58
[-0.17,1.34]
0.61
[-0.15,1.35]
0.36
[-0.38,1.10]
0.00
[-0.73.0.73]
1.00
[0.22,1.78]**
0.11
[-0.42,0.64]
0.22
[-0.31,0.75]
0.05
[-0.50,0.57]
0.00
[-0.53,-0.53]
0.18
[-0.35,0.71]
Table 2.3 Health-Related Quality of Life Measures – Control Design
54
54
Study
Moadel et
al., 2007
(148)
Vadiraja et
al., 2009
(80)
Outcome
measure
FACT-G
(Overall)
FACT-SWB
(Social)
FACT-FWB
(Functional)
FACT-EWB
(Emotional)
FACT-PWB
(Physical)
EORTC-Phys
(Physical)
EORTC-Role
(Role)
EORTC-Emo
(Emotional)
EORTC-Cog
(Cognitive)
EORTC-Soc
(Social)
SEM (Baseline
Internal
SD x √1 –
consistency
Cronbach α)
(Cronbach α)
Baseline 0.5 SD
Yoga
Control
Yoga
Control
.89
5.96
8.11
8.99
12.23
.69
3.47
3.46
3.12
3.11
.80
2.85
3.45
3.19
3.86
.74
2.43
3.03
2.39
2.97
.82
2.24
2.98
2.64
3.52
.71
12.49
16.68
11.60
15.49
.52
24.15
25.22
17.43
18.20
.80
8.84
7.80
9.89
8.72
.73
9.35
10.97
9.00
10.56
.77
12.73
11.72
13.28
12.22
Baseline Mean
(SD)
Yoga
76.53
(17.98)
20.98
(6.23)
18.33
(6.38)
16.36
(4.77)
20.87
(5.28)
73.20
(23.20)
72.72
(34.86)
56.45
(19.77)
85.29
(18.00)
52.82
(26.55)
Control
77.54
(24.45)
22.12
(6.22)
18.19
(7.72)
16.50
(5.94)
20.73
(7.03)
62.72
(30.98)
71.59
(36.40)
51.58
(17.44)
82.67
(21.12)
52.41
(24.43)
Within-Groups Mean
Difference (95% CI)
Yoga
1.54
[-3.60,6.68]
-0.30
[-1.99,1.39]
-0.16
[-2.08,1.76]
1.83
[0.51,3.14]**
0.16
[-1.51,1.83]
0.06
[-7.30,7.42]
7.16
[-3.32,17.64]
18.67
[12.47,24.87]ab
5.28
[0.13,10.43]
2.14
[-5.53,9.81]
Within-Groups
Effect Size - Hedges’ g
(95% CI)
Control
Yoga
Control
-7.16
0.09
-0.30
[-16.25,1.93]
[-0.20,0.37]
[-0.68,0.09]
-3.95
-0.05
-0.60
[-6.41,-1.49]ab [-0.34,0.24] [-1.00,0.19]**
-1.98
-0.02
-0.26
[-4.88,0.92]
[-0.31,0.26]
[-0.63,0.12]
-0.41
0.40
-0.07
[-2.65,1.83]
[0.10,0.70]** [0.44,0.31]
-0.82
0.03
-0.12
[-3.39,1.75]
[-0.26,0.32]
[-0.49,0.26]
6.24
0.00
0.20
[-4.19,16.66]
[-0.29,0.30]
[-0.14,0.54]
1.26
0.20
0.03
[-11.81,14.33] [-0.10,0.50]
[-0.30,0.37]
7.65
0.89
0.36
[0.48,14.82] [0.54,1.25]*** [-0.01,0.70]*
-1.90
0.30
-0.08
[-9.66,5.86]
[0.00,0.61]* [-0.42,0.25]
-2.48
0.08
-0.10
[-10.78, 5.82] [-0.22,0.38]
[-0.43,0.23]
Between Groups
Mean Difference
(95% CI)
8.70
[-0.96,18.36]c
3.65
[0.75,6.56]cd
1.82
[-1.53,5.17]
2.24
[-0.19,4.67]
0.98
[-1.96,3.91]
6.18
[-6.24,18.60]
5.90
[-10.64, 22.44]
11.02
[1.57,20.47]cd
7.18
[-1.83,16.19]
4.62
[-6.74,15.98]
Between Groups
Effect size
– Hedges’ g
(95% CI)
0.43
[-0.05,0.91]
0.60
[0.11,1.09]*
0.26
[-0.22,0.74]
0.44
[-0.05,0.92]
0.16
[-0.32,0.64]
0.23
[-0.23,0.68]
0.16
[-0.29,0.61]
0.53
[0.07,0.99]*
0.36
[-0.10,0.82]
0.18
[-0.27,0.64]
Table 2.3 – continued
55
55
Study
Chandwani
et al., 2010
(157)
Cohen
et al., 2004
(158)
Culos-Reed
et al., 2006
(149)
Outcome
measure
CES-D 20
(Depression)
STAI-State
(Anxiety)
CES-D 20
(Depression)
STAI – State
(Anxiety)
POMS-D
(Depression)
CES-D 20
(Depression)
PANAS-P
(Positive
Danhauer
Affect)
et al., 2009
PANAS-N
(Negative
(79)
Affect)
FACIT-Sp
(Spiritual
Well-Being)
FACIT-Sp
Moadel et al.,
(Spiritual
2007 (148)
Well-Being)
PANAS-P
(Positive
*Vadiraja
Affect)
et al., 2009
PANAS-N
(80)
(Negative
Affect)
HADS-A
*Vadiraja
(Anxiety)
et al., 2009b
HADS-D
(147)
(Depression)
SEM (Baseline
Internal
SD x √1 –
consistency
Cronbach α)
(Cronbach α)
Baseline 0.5 SD
Baseline Mean
(SD)
Yoga
11.60
(9.30)
36.30
(13.70)
10.20
(11.00)
34.30
(12.30)
Within-Groups Mean
Difference (95% CI)
Yoga
Control
Yoga
Control
Control
Yoga
Control
10.50
-5.00
-3.50
(6.70)
[-8.85,-1.15]ab
[-7.23,0.24]a
32.70
-8.30
-2.50
(10.00) [-13.09, -3.51]ab [-6.75,1.75]a
9.60
-1.20
.10
(8.57)
[-5.91,3.51]
[-3.79,4.00]
37.80
-0.20
-4.00
(14.60)
[-5.54,5.14]
[-10.65,2.65]a
.87
3.35
2.42
4.65
3.35
.94
3.36
2.45
6.85
5.00
.93
2.91
2.27
5.50
4.29
.95
2.75
3.26
6.15
7.30
.95
1.76
1.14
3.93
2.55
4.70
(7.86)
5.44
(5.10)
-2.48
[-5.68,0.72]a
.93
2.57
3.89
4.85
7.35
16.30
(9.70)
16.60
(14.70)
.87
3.17
3.53
4.40
4.90
32.70
(8.80)
.87
2.99
3.39
4.15
4.70
.87
2.38
2.60
3.30
.87
3.33
3.99
.87
2.62
.87
Within-Groups
Effect Size - Hedges’ g
(95% CI)
Yoga
-0.48
[-0.86,-0.09]*
-0.63
[-1.04,-0.23]**
-0.12
[-0.58,0.35]
-0.02
[-0.48,0.45]
Control
-0.32
[-0.67,0.03]
-0.20
[-0.55,0.15]
0.01
[-0.48,0.51]
-0.30
[-0.80,0.21]
0.06
[-2.54,2.66]
-0.34
[-0.80,0.11]
-8.20
[-13.27,-3.17]ab
1.20
[-7.14,9.54]
30.60
(9.80)
5.50
[1.16,9.84]ab
19.00
(8.30)
18.50
(9.40)
3.60
23.30
(6.60)
4.62
5.53
2.66
3.64
3.82
3.18
.79
1.77
.87
1.45
Between Groups
Mean Difference
(95% CI)
Between Groups
Effect size
– Hedges’ g
(95% CI)
-1.50
[-6.87,3.87]
-5.80
[-12.18,0.58]cd
-1.30
[-7.52,4.92]
-3.80
[-12.24,4.64]c
-0.14
[-0.65,0.37]
-0.46
[-0.98,0.05]
-0.15
[-0.85,0.55]
-0.31
[-1.02,0.39]
0.01
[-0.43,0.45]
-2.54
[-6.66,1.58]c
-0.39
[-1.04,0.25]
-0.82
[-1.42,-0.22]**
0.07
[-0.42,0.57]
-9.40
[-19.34,0.54]cd
-0.69
[-1.45,0.06]
1.20
[-4.21,6.61]
0.64
[0.08,1.21]*
0.11
[-0.39,0.60]
4.30
[-2.71,11.31]c
0.45
[-0.29,1.19]
-5.00
[-8.91, -1.09]ab
1.40
[-3.63,6.43]
-0.65
[-1.22,0.08]*
0.14
[-0.36,0.63]
-6.40
[-12.84,0.04]cd
-0.73
[-1.49,0.03]
23.20
(7.20)
2.70
[-0.92, 6.31]a
-1.70
[-6.16,2.76]
0.38
[-0.15,0.91]
-0.19
[-0.69,0.31]
4.40
[-1.40,10.20]cd
0.56
[-0.19,1.30]
37.87
(9.24)
34.58
(11.06)
-1.02
[-3.59,1.55]
-1.83
[-6.49,2.83]
-0.11
[-0.40,0.17]
-0.15
[-0.52,0.23]
0.81
[-4.08,5.70]
0.08
[-0.40,0.56]
3.69
24.05
(7.28)
21.81
(7.37)
3.80
[1.62,5.98]ab
1.52
[-1.17,4.21]
0.52
[0.20,0.84]***
0.19
[-0.15,0.53]
2.28
[-1.14,5.70]
0.30
[-0.15,0.76]
5.30
4.41
22.15
(10.60)
25.22
(8.82)
-9.24
[-12.41,-6.07]ab
-3.37
[-6.78,0.04]a
-0.86
[-1.21,-0.51]***
-0.33
[-0.67,0.01]
-5.87
[-10.56,-1.18]cd
-0.57
[-1.03,-0.11]*
1.82
1.94
1.99
1.25
2.01
1.74
8.05
(3.87)
7.57
(4.02)
9.35
(3.98)
8.00
(3.47)
-3.17
[-4.27,-2.07]ab
-3.43
[-4.57,-2.29]ab
-1.23
[-2.56,0.10]
-1.47
[-2.71,-0.23]a
-0.86
[-1.21,-0.51]***
-0.89
[-1.25,-0.54]***
-0.31
[-0.65,0.03]
-0.40
[-0.74,-0.05]*
-1.94
[-3.65,-0.23]c
-1.96
[-3.65.-0.27]cd
-0.51
[-0.97,-0.05]*
-0.52
[-0.98,-0.06]*
Table 2.4 Psychosocial Measures – Control Design
56
56
Study
Internal
Outcome measure consistency
(Cronbach α)
SEM (Baseline
SD x √1 –
Cronbach α)
Baseline 0.5 SD
Baseline Mean
(SD)
Within-Groups Mean
Difference (95% CI)
Yoga
Control
Yoga
Control
Yoga
Control
Yoga
Control
Within-Groups
Effect Size - Hedges’ g
(95% CI)
Yoga
Control
Between Groups
Mean Difference
(95% CI)
Between Groups
Effect size
– Hedges’ g
(95% CI)
BFI
(Fatigue)
.94
0.39
0.54
0.80
1.10
2.30
(1.60)
2.30
(2.20)
-0.40
[-1.58,0.78]a
0.20
[-1.17,1.57]
-0.12
0.05
[-0.49,0.24] [-0.29,0.39]
-0.60
[-2.44,1.24]c
-0.17
[-0.68,0.34]
PSQI
(Sleep)
.84
1.52
1.56
1.90
1.95
7.30
(3.80)
7.10
(3.90)
-0.70
[-2.46,1.06]
-0.40
[-2.15,1.35]
-0.15
-0.08
[-0.51,0.22] [-0.42,0.27]
-0.30
[-2.79,2.19]
-0.06
[-0.57,0.45]
BFI
(Fatigue)
.96
0.48
0.44
1.20
1.10
3.10
(2.40)
2.80
(2.20)
0.00
[-1.03,1.03]
0.30
[-0.72,1.32]
0.00
0.15
[-0.47,0.47] [-0.35,0.64]
-0.30
[-1.76,1.16]
-0.14
[-0.84,0.56]
PSQI
(Sleep)
.84
2.00
1.88
2.50
2.35
6.50
(5.00)
7.20
(4.70)
-0.70
[-2.82,1.42]
0.90
[-1.23,3.03]
-0.16
0.21
[-0.62,0.32] [-0.29,0.71]
-1.60
[-4.62,1.42]
-0.37
[-1.07,0.34]
FACT- F
(Fatigue)
.95
3.00
2.64
6.70
5.90
30.10
(13.40)
32.70
(11.80)
9.70
[2.87,16.53]ab
-0.10
[-7.45,7.25]
0.72
-0.01
[0.14,1.30]* [-0.50,0.49]
9.80
[-0.27,19.87]cd
0.71
[-0.04,1.47]
PSQI
(Sleep)
.84
1.88
2.12
2.35
2.65
8.30
(4.70)
8.60
(5.30)
-2.20
[-4.65,0.25]a
-1.60
[-4.14,0.94]
-0.46
-0.31
[-1.00,0.08] [-0.81,0.20]
-0.60
[-4.14,2.94]
-0.12
[-0.86,0.61]
Littman
et al., 2011
(150)
FACT-F
(Fatigue)
.95
1.30
1.90
2.90
4.25
43.10
(5.8)
43.20
(8.50)
1.9
[-0.20,4.00]a
-0.10
[-3.69,3.49]
0.33
-0.01
[-0.05,0.71] [-0.38,0.36]
2.00
[-2.16,6.16]c
0.25
[-0.27,0.78]
Moadel
et al., 2007
(148)
FACT-F
(Fatigue)
.95
2.75
3.22
6.16
7.21
34.27
(12.31)
35.88
(14.42)
2.29
[-1.11,5.69]
-1.11
[-6.50,4.30]
0.19
-0.08
[-0.10,0.48] [-0.45,0.30]
3.40
[-2.66,9.46]c
0.27
[-0.21,0.75]
Chandwani
et al., 2010
(157)
Cohen
et al., 2004
(158)
Danhauer
et al., 2009
(79)
Table 2.5 Symptom Measures – Control Design
57
57
Study
Outcome
Measure
Bower
et al., 2011
(151)
SF-36 Gen
Health (Overall)
Danhauer
et al., 2008
(152)
FACT-G
(Overall)
FACT-SWB
(Social)
FACT-FWB
(Functional)
FACT-EWB
(Emotional)
FACT-PWB
(Physical)
SF-12 PCS
(Physical)
SF-12 MCS
(Mental)
Duncan
et al., 2008
(153)
Galantino
et al., 2011
(154)
FACT-G
(Overall)
FACT-B
(Overall)
Internal
consistency
(Cronbach α)
SEM (Baseline
SD x √1 –
Cronbach’s α)
Baseline
0.5 SD
Within-Groups
Baseline Mean
Mean Difference
(SD)
(95% CI)
Within-Groups Effect
Size- Hedges’ g
(95% CI)
50.50
(22.10)
14.50
[1.44, 27.56]ab
0.61
[0.00,1.21]*
75.90
(10.70)
24.00
(4.30)
18.00
(4.30)
14.00
(2.10)
20.00
(5.70)
41.20
(11.40)
48.40
(8.60)
3.50
[0.33,6.67]
0.80
[-0.53,2.13]
0.80
[-0.44,2.04]
-0.20
[-0.81,0.41]
2.10
[0.47,3.73]
2.30
[-0.98,5.58]
3.60
[1.04,6.16]
0.32
[0.02,0.63]*
0.18
[-0.12,0.47]
0.19
[-0.11,0.49]
-0.10
[-0.39,0.20]
0.38
[0.07,0.68}*
0.21
[-0.09,0.50]
0.41
[0.11,0.72]**
7.30
64.6
(14.16)
10.63
[4.44,16.82]ab
0.69
[0.24,1.14]**
10.09
89.33
(20.18)
16.72
[-5.07,28.37]ab
0.81
[0.14,1.48]*
.81
9.63
11.05
.89
3.55
5.35
.69
2.39
2.15
.80
1.92
2.15
.74
1.07
1.05
.82
2.42
2.85
.86
4.27
5.70
.81
3.75
4.30
.89
4.70
.90
6.38
Table 2.6 Health-Related Quality of Life Measures – Single Group Design
58
58
Study
SpeedAndrews
et al., 2010
(155)
Ülger&Yağli,
2010 (156)
Outcome
Measure
FACT-B
(Overall)
FACT-SWB
(Social)
FACT-FWB
(Functional)
FACT-EWB
(Emotional)
FACT-PWB
(Physical)
SF-36 PCS
(Physical)
SF-36 MCS
(Mental)
NHP-Total
(Overall)
NHP-Social
(Social)
NHP-Emotion
(Emotional)
NHP-Physical
(Physical)
Internal
consistency
(Cronbach α)
SEM (Baseline
SD x √1 –
Cronbach’s α)
Baseline
0.5 SD
.90
6.67
10.55
.69
2.95
2.65
.86
1.83
2.45
.69
2.67
2.40
.81
2.31
2.65
.90
3.38
5.35
.90
2.85
4.50
.70
12.64
11.54
.56
20.01
15.08
.83
9.68
11.74
.72
11.86
11.21
Within-Groups
Baseline Mean
Mean Difference
(SD)
(95% CI)
108.20
8.90
(21.10)
[-1.51,19.31]a
20.70
0.70
(5.30)
[-1.84,3.24]
19.90
1.80
(4.90)
[-0.29,3.89]
15.80
2.10
(4.80)
[-0.37,4.57]
20.90
2.00
(5.30)
[-0.37,4.37]
45.50
3.70
(10.70)
[-1.41,8.81]a
44.40
4.20
(9.00)
[-0.01,8.41]a
103.23
-17.32
(23.08)
[-26.70,-7.94]ab
43.01
-20.56
(30.16)
[-32.24,-8.88]ab
61.37
-26.61
(23.47)
[-36.35,-16.87]ab
44.51
-20.41
(22.41)
[-29.77,-11.05]ab
Within-Groups Effect
Size - Hedges’ g
(95% CI)
0.39
[-0.08,0.86]
0.13
[-0.33,0.58]
0.39
[-0.08, 0.86]
0.39
[-0.09,0.86]
0.38
[-0.09,0.85]
0.32
[-0.14,0.79]
0.45
[-0.03,0.93]
-0.77
[-1.26,-0.29]**
-0.74
[-1.22,-0.26]**
-1.15
[-1.70,-0.60]***
-0.92
[-1.43,-0.41]***
Table 2.6 – continued
59
59
Study
Bower
et al.,2011(151)
Danhauer
et al., 2008
(152)
Duncan
et al., 2008
(153)
Speed-Andrews
et al., 2010
(155)
Ülger &Yağli,
2010 (156)
Outcome
Measure
BDI
(Depression)
CES-D 20
(Depression)
STAI-State
(Anxiety)
PANAS-P
(Positive Affect)
PANAS-N
(Negative
Affect)
FACIT-Sp
(Spiritual WellBeing)
FACIT-Sp
(Spiritual WellBeing)
CES-D 10
(Depression)
STAI – State
(Anxiety)
STAI-State
(Anxiety)
Internal
consistency
(Cronbach α)
SEM (T1 SD x√1
– Cronbach’s α)
Baseline
0.5 SD
Baseline Mean
(SD)
15.40
(8.00)
12.30
(8.60)
34.20
(10.70)
34.70
(8.60)
Within-Groups
Mean Difference
(95% CI)
-7.90
[-12.20,-3.60]ab
-3.10
[-5.66,-0.54]a
-2.40
[-5.57,0.77]a
1.50
[-1.06,4.06]
Within-Groups Effect
Size – Hedges’ g
(95% CI)
-1.00
[-1.69,-0.32]**
-0.36
[-0.66,-0.05]*
-0.22
[-0.52,0.08]
0.17
[-0.12,0.47]
.81
3.49
4.00
.93
2.27
4.30
.95
2.39
5.35
.87
3.10
4.30
.87
1.80
2.50
15.80
(5.00)
-1.70
[-3.14,-0.26]
-0.35
[-0.65,-0.04]*
.87
2.56
3.55
38.70
(7.10)
1.40
[-0.74,3.54]
0.19
[-0.10,0.49]
.87
3.65
5.07
30.83
(10.13)
5.88
[1.93,9.83]ab
0.60
[0.16,1.04]**
.85
0.23
0.30
.85
2.59
3.35
.95
1.75
3.92
31.00
(0.60)
45.20
(6.70)
55.05
(7.83)
-0.20
[-0.47,0.07]
1.80
[-1.48,5.08]
-11.70
[-14.96,-8.44]ab
-0.34
[-0.81,0.13]
0.25
[-0.21,0.71]
-1.51
[-2.14,-0.88]***
2.7 Psychosocial Measures – Single Group Design
60
60
Study
Outcome
Measure
Internal
consistency
(Cronbach α)
SEM (Baseline
SD x √1 –
Cronbach’s α)
Baseline
0.5 SD
Within-Groups
Baseline Mean
Mean Difference
(SD)
(95% CI)
Within-Groups Effect
Size - Hedges’ g
(95% CI)
FSI Average
(Fatigue)
.97
0.19
0.55
6.3
(1.1)
-3.6
[-4.44,-2.76]ab
-2.34
[-3.47,-1.22]***
PSQI (Sleep)
.84
1.64
2.05
7.1
(4.1)
-0.7
[-2.99,1.59]
-0.17
[-0.72,0.38]
Danhauer
et al., 2008 (152)
FACT-F (Fatigue)
.95
2.71
6.05
34.6
(12.1)
2.7
[-0.83,6.27]a
0.22
[-0.08,0.52]
Ülger&Yağli,
2010 (156)
NHP-S (Sleep)
.72
15.64
14.78
43.50
(29.55)
-23.47
[-34.69,-12.25]ab
-0.88
[-1.38,-0.38]***
Bower
et al., 2011 (151)
*p<.05, ** p<.01, ***p< .001 (Reported between group p-values)
a
Within-Group mean difference exceeds 1SEM
b
Within-Group mean difference exceeds Baseline 0.5 SD
c
Between-Group mean exceeds control group 1 SEM
d
Between-Group mean exceeds control group Baseline 0.5 SD
Abbreviations: α, alpha; SD, standard deviation; CI, confidence interval; SF-36 & SF-12, Medical Outcomes Study Short-Form Health Survey;
EORTC QLQ C30, European Organization for Research and Treatment of Cancer-Quality of Life C30; FACT-B, Functional Assessment of
Cancer Therapy–Breast; FACT-G, Functional Assessment of Cancer Therapy-General; NHP, Nottingham Health Profile; BDI, Beck Depression
Inventory; CES-D 20, Center for Epidemiologic Studies Depression Scale – 20 Item; STAI, State/Trait Anxiety Inventory; POMS, Profile of
Mood States; PANAS, Positive&Negative Affect Schedule; FACIT-Sp, Functional Assessment of Chronic Illness Therapy-Spiritual Well-Being;
HADS, Hospital Anxiety & Depression Scale ; BFI, Brief Fatigue Inventory; FSI, Fatigue Symptom Inventory; PSQI, Pittsburgh Sleep Quality
Index; FACT-F, Functional Assessment of Cancer Therapy-Fatigue
Table 2.8 Symptom Measures – Single Group Design
61
61
62
CHAPTER 3: MODELLING ASSOCIATIONS BETWEEN YOGA
PARTICIPATION, AFFECT, MINDFULNESS AND HEALTH OUTCOMES IN
CANCER SURVIVORS
63
3.1 Introduction
Receiving a cancer diagnosis, undergoing treatment and the subsequent recovery
takes a great toll on many cancer survivors. Research demonstrates psychosocial distress
stemming from the cancer experience is a significant problem for up to half of all cancer
patients, and many survivors experience lowered overall health-related quality of life
(HRQL) both during and following active treatment (7). In some cases, cancer survivors
exhibit signs of clinical psychological disorders and a host of more general mood, stress,
sleep, and fatigue symptoms (8-12).
In this capacity, regardless of intervention specifics, exercise has been shown to
improve almost all aspects of physical and psychological functioning in cancer survivors
(14,15). Specifically, exercise enhances a variety of HRQL, health, functional and
psychosocial outcomes in various cancer survivor groups both during and after cancer
treatment, and may also help to manage the long-term effects of treatment (20-22).
However, current data suggests only 22% of cancer survivors are active enough to
achieve health benefits (25) and only half of cancer patients offered an exercise program
will undertake and complete the intervention (26).
3.1.1 Yoga in Cancer Settings
Within the growing body of PA and cancer research, there have been calls to
examine modes of exercise from the area of integrative medicine (29). In this regard,
yoga is quickly emerging as an important integrative therapeutic approach in oncology.
Within the larger field of exercise and cancer, yoga is often considered a gentle, lowintensity form of exercise (22,32). Studies comparing the effects of yoga and exercise
indicate that, in both healthy individuals and those with various health conditions, yoga
64
may be as effective as more contemporary western forms of exercise, including walking,
jogging, cycling, and aerobics, at improving a variety of health-related outcome measures
(48). Within oncology, yoga groups compared to waitlist control groups or supportive
therapy groups show greater improvements in overall HRQL, psychological health,
stress-related symptoms, sleep and fatigue indices (1,2,33).
3.1.2 Theoretical Mechanisms of Action
Despite preliminary findings, little attention has been paid to the psychological
mechanisms by which benefits are accrued via yoga practice. Proposed mechanisms
tested in the current study include affect regulation and mindfulness.
3.1.2.1 Affect Regulation. Positive affect is an independent, adaptive pathway in
the cancer experience (68). Both baseline positive affect and enhancement of positive
affect are important components of symptom management and cancer recovery (166). In
general, exercise increases positive affect and reduces negative affect (71,72).
Specifically, positive affect tends to increase pre- to post-exercise following exercise
intensities that are not exhaustive (73). Individuals tend to choose and adhere to activities
associated with positive affective experiences (71,72,167). Fostering positive affective
experiences via a structured exercise program is an important target for interventions
designed to facilitate post-program exercise adherence in cancer settings (76).
In two recent nonrandomized trials, a single session of yoga was associated with
significant improvements in positive affect and reductions in negative affect, comparable
to changes seen with aerobic exercise (77,78). A pilot RCT examining the effects of a 10week restorative yoga in breast cancer survivors on or off treatment reported significant
benefits favouring the yoga group for positive affect, mental health, depression and
65
spirituality outcomes (79). A second pilot RCT examining the effects of a 12-week yoga
program in breast cancer outpatients undergoing adjuvant radiotherapy reported
significant improvement in positive affect, emotional function and cognitive function,
and decreased negative affect in the yoga group as compared to controls (80).
3.1.2.2 Circumplex Model of Affect. Given the proposed importance of positive
affect and yoga’s influence on this phenomena, the Circumplex Model of Affect (81)
offers a parsimonious explanatory approach. According to this model, affect is best
understood as an underlying state comprised of two orthogonal dimensions: feeling good
or bad (valence), and feeling energized or enervated (activation). These two dimensions
of valence and activation, termed core affect, underlie all affective states and specific
affective states are considered combinations of these two dimensions. Four quadrants can
be described as: energy (high activation pleasure), calmness (low activation pleasure),
tension (high activation displeasure), and tiredness (low activation displeasure) (83) (see
Figure 3.1).
3.1.2.3 Dual Mode Theory. This Circumplex approach can be further
contextualised within exercise settings utilising Dual Mode Theory (168) which suggests
exercise experiences, including yoga, are regulated by affect via the continuous interplay
of two mechanisms; 1) cognitive processes, and 2) interoceptive cues. The relative
importance of these two factors in the regulation of core affect is hypothesized to shift
systematically as a function of exercise intensity, with cognitive factors being dominant
at low and moderate intensities and interoceptive cues becoming more prominent at
higher intensities (85).
66
3.1.2.4 Effort-Related Attention Model. Individual attention during exercise can
influence changes in affective responses. Associative attention is defined as focusing
attention on present-moment physical sensations related to exercise, where exercisers
seek to monitor sensory input and adjust their effort accordingly. Dissociative attention is
defined as focusing on non-exercise related stimuli that divert attention away from
internal sensations and present moment exercise experience (90). The Effort-Related
Attention Model suggests these attention strategies during exercise vary based on
intensity (91). During low or moderate intensity workloads attention is voluntary and can
switch between dissociative and associative attentional strategies, as external stimuli do
not compete with internal stimuli, while at higher exercise intensities increased
physiological cues demand attention. As a result attention narrows, shifts internally and
becomes associative (92).
3.1.3 Mindfulness
Mindfulness is the systematic development of the ability to non-judgmentally
direct attention towards events in the field of consciousness in the present moment (169).
Research reveals both trait and state mindfulness, as specific forms of attention, are
related to affect regulation (170). The various extant measures of mindfulness have been
associated with higher positive affect and lower negative affect, perhaps through
engagement of attention upon immediate experience (94).
Research in the field of psychosocial oncology further indicates cancer survivors
who participate in mindfulness-based programs, of which yoga is a significant
component, with higher mindfulness scores report less mood disturbance and symptoms
of stress as well as report significantly improved self-regulated emotion and behaviour,
67
all of which contribute to enhanced HRQL (171-174). These changes in mindfulness
mediate changes in psychological outcomes including perceived stress and positive states
of mind (175). These psychological benefits have been maintained up to a year at followup (176-179).
It has been suggested yoga adds a contemplative element to exercise and can be
conceptualised as ‘mindfulness in motion’ (96). Specifically, the practice of yoga
provides an opportunity for sustained attention to the body, breath and mind through
progressive sequences of dynamic movements, restful postures, breathing exercises and
periods of meditative awareness, facilitating a strong mindfulness practice. However,
research evidence linking yoga in eliciting these changes in mindfulness remains
equivocal. Lengacher and colleagues (98) found minutes of yoga practice as part of the
overall MBSR program was not significantly related to positive changes in psychosocial
status and HRQL. These findings are in contrast to the work of Carmody and Baer (99),
whose research in a non-cancer heterogeneous medical population found the yoga
component of the mindfulness program to be most strongly related to improvements of
mindfulness measures, psychological symptoms and perceived stress. These findings
have more recently been corroborated in a study by Sauer-Zavala and colleagues (100),
which reports 105 minutes in yoga practice was associated with greater psychological
wellbeing independent of equivalent amounts of body scan or sitting meditation practices.
3.1.3.1 Facets of Mindfulness. The Five Facet Mindfulness Questionnaire was
chosen to examine mindfulness in yoga settings (180). These five facets of mindfulness
have been described as follows: the ability to1) observe and 2) describe experiences in
the present moment while 3) acting with awareness 4) without judgment or 5) prolonged
68
reaction (181). In cancer settings, work by Branstrom and colleagues (175) suggests prepost changes in facets of mindfulness mediate changes in psychological outcomes both
post-MBSR program and at a three- but not six-month follow-up (182). Recent work by
Garland and colleagues (183) is consistent with these findings, indicating significant
relationships between increased facets of mindfulness and reductions in mood and stress
symptoms. In a mediation analyses by Labelle and colleagues (184), while significant
changes in facets of mindfulness were observed, they did not mediate treatment change in
depression. It has also been suggested the five facet mindfulness questionnaire does not
measure mindfulness per se, but rather sequalae of mindfulness (185).
In sum, cancer-related distress and lower HRQL are prevalent among many
cancer survivors. Exercise has proven effective in ameliorating this distress and
improving HRQL. Despite these benefits, only a fraction of cancer survivors engage in
enough exercise to receive these benefits and over half of cancer survivors who engage in
an exercise program will not maintain it. Yoga is considered a low-intensity form of
physical activity that has been shown to independently improve HRQL, psychosocial
outcomes and symptom indices. Two mechanisms by which this may occur are via
increased positive affect and mindfulness.
3.2 Objectives
Given preliminary evidence that yoga practice is associated with, 1) positive affect, 2)
mindfulness, and 3) improved HRQL and mental health outcomes in cancer survivors, the
present study sought to examine:
69
3.2.1 Acute Class Effects
Objective 1a: To examine potential changes in valence, activation, and attention
pre-post each class of a seven week yoga intervention. Objective 1b: To examine
associations among changes in valence, activation and attention within the yoga classes
and whether these changes were moderated by proposed predictor variables, including
baseline affect and mindfulness. Covariates included baseline demographics, physical
activity, previous yoga experience, beliefs about yoga, affect, mindfulness, or
psychological functioning.
3.2.2 Longitudinal Program Effects
Objective 2a: To examine changes longitudinally pre-post yoga program and at
three- and six-month follow-ups, in measures of mood disturbance, stress symptoms and
HRQL as well as proposed predictor variables, including affect and mindfulness.
Objective 2b: To examine whether improvements in mood disturbance, stress symptoms
and HRQL were associated with concurrent changes in proposed predictor variables
including affect and mindfulness. Covariates included age, time since diagnosis, physical
activity, previous yoga experience and beliefs about yoga.
3.2.3 Yoga Practice Maintenance:
Objective 3a: To examine whether participants maintain their yoga practice
longitudinally, pre-post and three- and six-month follow-ups. Objective 3b: To examine
whether maintenance of yoga practice was associated with mood disturbance, stress
symptoms and HRQL as well as proposed predictor variables, including affect and
mindfulness. Covariates included age, time since diagnosis, physical activity, previous
yoga experience and beliefs about yoga.
70
3.3 Methods
3.3.1 Participants
Ethical approval was obtained from the Conjoint Health Research Ethics Board of
the University of Calgary/Alberta Health Services (Ethics ID # 23313). Program
participants were comprised of a heterogeneous group of cancer survivors enrolled in the
ongoing “Yoga Thrive: Therapeutic Yoga for Cancer Survivors” program. Participants
were eligible for study inclusion if they were: 1) age 18 years or older; and 2) had
received a cancer diagnosis at any time in the past. Previous participation in the Yoga
Thrive program was not an exclusion criterion, but was evaluated as part of the study.
Yoga Thrive is offered in several community locations throughout the city of Calgary,
Alberta, Canada. Participants were informed of the research study at the time of class
registration, either via telephone or online. Those that indicated interest in participating in
the research study were contacted by the study coordinator. Participants completed inclass measures before and after each yoga class in the seven-week program. Baseline,
post-program and the three- and six-month follow-up surveys were completed online.
3.3.2 Program
The Yoga Thrive program is a research-based, therapeutic yoga program for
cancer survivors and their support persons. This gentle, seven-week yoga program is
based on contemporary yoga practices modified for cancer survivors. Details of the
program have been previously described (149,186). A typical 75-minute class was broken
down as follows: 0–10 minutes - gentle breathing and movement, laying supine, with legs
flexed at the hip and supported by a wall; 10–60 minutes - a series of 6–10 modified yoga
postures / sequences comprised of gentle stretching and strengthening exercises with
71
attention to breath and bodily sensations; 60–75 minutes – guided supine meditation with
attention placed on both breathing and bodily sensations. The yoga classes became
progressively more challenging over the seven-week course, as participant flexibility and
strength improved.
3.3.3 Recruitment
Power calculations were derived using GPower 3.1. The sample size / power
calculation was based on a clinically significant (135) change in score pre-post program
on the Profile of Mood States (POMS) Total Mood Disturbance (TMD) score as the
primary outcome (Cohen’s d = .50). With an α of .05 and power of .80, a minimum of 34
participants were to be recruited from the Yoga Thrive program. During the eight-month
recruitment period, 125 people enrolled in the program over 233 times in 23 separate
seven-week sessions over four waves (Fall, Winter, Spring, Summer). Of these 125
participants, 100 expressed an interest in participating in the research component of the
program at the time of enrollment. Of these 100 interested participants, 18 had not
received a cancer diagnosis and were attending the program as support persons, eight
subsequently declined study participation when contacted by the study coordinator, three
consented but their seven-week session was later cancelled due to lack of enrollment and
the study coordinator was unable to contact one interested participant. Subsequently, 70
participants were eligible for the current study, 66 of who completed baseline measures
used in the current analyses (see Figure 3.2).
72
3.3.4 Instruments
3.3.4.1 Baseline
Demographics: Demographic information included age, education, marital status
and current employment status. Medical history included self-reported cancer diagnosis,
date of diagnosis and type(s) of cancer treatment.
Beliefs about Yoga Scale. The BAYS (187) is an 11-item self-report measure
developed to examine common positive and negative beliefs about yoga in order to: 1)
inform promotion of yoga to diverse populations; 2) frame yoga programs to retain
participants; and 3) help yoga instructors understand participant expectations related to
yoga (baseline α = .78).
3.3.4.2 Acute Effects (objectives 1a & 1b).
Feeling Scale. The FS (188) is an 11-point, single-item, bipolar measure of
affective valence (pleasure-displeasure). The scale ranges from -5 to +5. Anchors are
provided at zero (‘Neutral’) and all odd integers, ranging from ‘Very Good’ (+5) to ‘Very
Bad’ (-5). Instructions were to, “estimate how good or bad you feel right now.” The FS is
commonly used for the assessment of affective responses during exercise (189).
Felt Arousal Scale. The FAS (190) is a 6-point, single-item measure of perceived
activation. The scale ranges from 1 to 6, with anchors at 1 (‘Low Arousal’) and 6 (‘High
Arousal’). Instructions were to, “estimate how aroused you feel right now (low arousal =
calm or fatigued, high arousal = anxious or energized).
Association / Dissociation Scale. The A/D S (191) is a single-item 10 point
bipolar scale designed to measure to what extent attention is primarily associative or
dissociative. The scale ranges from 1 to 10, with anchors at 1 (very associative) and 10
73
(very dissociative) (192). Instructions were to, “estimate your present state of attention
right now (associative = focused in present moment, dissociative = unfocused in past or
future).
3.3.4.3 Longitudinal Effects (objectives 2a & 2b).
Godin Leisure Time Exercise Questionnaire . The GLTEQ (193) was used to
assess physical activity levels. The LSI contains three questions that assess the frequency
of mild, moderate, and strenuous physical activity performed for at least 15 minutes
duration during free time in a typical week within the past month. A weekly total of
moderate-vigorous physical activity can also be computed from this scale (baseline α:
LSI = .88; moderate + vigorous physical activity = .86).
Activation-Deactivation Adjective Check List. The AD ACL (194) is a 20-item
measure of the four quadrants of circumplex affective space. The Energy pole is
theorized to map the high activation pleasure quadrant of the circumplex, Tension maps
the high-activation displeasure quadrant, Tiredness maps the low-activation displeasure
quadrant, and Calmness maps the low-activation pleasure quadrant (83) (baseline α:
energy = .88; tired = .84; tension = .79; calm = .80).
Five Facet Mindfulness Questionnaire. The FFMQ (180) is a 39-item five-point
Likert scale designed to measure five factors that represent elements of trait mindfulness
as it is currently conceptualized. The five facets are observing, describing, acting with
awareness, non-judging of and non-reaction to inner experience. Higher scores indicate
higher levels of mindfulness (baseline α: observing = .85; describing = .87; acting with
awareness = .89; non-judging = .91; non-reacting = .88).
74
Profile of Mood States - Short Form. The POMS-SF (195) is a widely used
instrument in the study of psychological aspects of cancer. It is a 37-item five point
Likert scale designed to assess fluctuating affective states over a one-week period. The
instrument provides a total mood disturbance score as well as six factor-based subscale
scores including tension-anxiety, depression-dejection, anger-hostility, vigour-activity,
fatigue-inertia, and confusion-bewilderment. Higher scores indicate more mood
disturbance, except on the vigour-activity subscale, where higher scores indicate more
vigour. Due to high multicollinearity (r >.7) between the POMS-SF total score and six
subscales at all time points, only the POMS-SF total score was used (baseline α = .93).
Calgary Symptoms of Stress Inventory. The C-SOSI (196) was designed to
measure physical, psychological, and behavioural responses to stressful situations. It is a
56-item five-point Likert scale inventory. The instrument provides a total stress score as
well as 10 subscale scores including peripheral manifestations, cardiopulmonary
symptoms, central-neurological symptoms, gastrointestinal symptoms, muscle tension,
habitual patterns, depression, anxiety/fear, emotional irritability, and cognitive
disorganisation. Higher scores indicate higher reported levels of stress symptoms. Only
the SOSI total stress score was used (baseline α = .91).
Functional Assessment of Cancer Therapy – General Version. FACT-G (version
4) (197) is a 27-item questionnaire that measures four domains of HRQL (physical wellbeing, functional well-being, social well-being, emotional well-being) in patients
undergoing cancer treatment. The FACT-G employs a five-point Likert scale for all
questions. Due to high multicollinearity (r >.7) between the FACT-G total score and four
subscales at all time points, only the FACT-G total score was used (baseline α = .93).
75
3.3.4.4 Yoga Practice Maintenance (objectives 3a & 3b). Participants were asked
to report weekly frequency of ongoing yoga practice via the Yoga Thrive program,
community-based yoga programs, home yoga practice or combinations thereof.
3.3.5 Data Analysis
All data analyses were conducted using IBM SPSS version 19. Demographics and
medical history were described using frequency and descriptive statistics to characterize
study participants. This investigation used a repeated measures design in which
participants served as their own control. Before beginning primary analyses, all
continuous variables were checked for normality (skewness, kurtosis) and corrected
where necessary. Statistical analyses were conducted on the entire sample (n = 66).
3.3.5.1 Multilevel Models. In examining both the acute pre-post effects of each
yoga session (objectives 1a & 1b) and the longitudinal effects of yoga practice
(objectives 2a & 2b), two sets of multilevel (mixed model) analyses were conducted: 1)
estimated marginal means (piecewise) models to examine effects over time; and 2)
multilevel regression analyses to examine associations between outcomes and predictor
variables. Multilevel models were appropriate for analyzing data with dependent
observations (such as within-subject repeated measures). That is, each time measurement
of a given variable (level 1) was nested within each participant (level 2). These models
take into account this hierarchical structure by modelling at both levels (198). In addition,
multilevel models retain cases for which missing data is present and provide a valid
analysis when data are assumed missing at random (199). Correlations among
observations from the same individual were modeled using an unstructured covariance
matrix across all time points.
76
In each multilevel regression model time was measured continuously and
included linear, quadratic and cubic terms (200). All models were tested step by step. An
initial unconditional model was developed for each outcome variable (data not shown)
followed by unconditional growth models (data not shown). Based on these growth
models, predictor variables were tested individually for main effects and for interaction
effects with each time term (data not shown). Significant predictors and their time
interactions, if also significant, were then tested together as part of their respective
overall scale (data not shown). A final trimmed model was developed by entering all
significant predictors and their interactions to test overall prediction of outcome variables
across time (exclusion p >.1). This method of model development has proven robust with
smaller sample sizes and ensures these models do not tax the “carrying-capacity” of the
dataset (201).
If interaction terms were not significant, it could be assumed the relationships
between predictor and outcome variables were constant throughout the study. In each
model, the time variable was centered at initial status; therefore the intercept of the
regression model can be interpreted as participant reports of the outcome variable at
baseline. To enhance interpretability of the model intercept parameters, all predictor
variables were grand-mean centered.
3.3.5.2 Clinical Significance. Clinical significance, or meaningful clinically
important difference, a marker of the effectiveness of programs, was assessed using
multiple criteria including: 1) An effect size using Cohen’s interpretation of >.20 as a
small effect >.50 as a moderate effect and >.80 as a large effect (142). Effect sizes were
calculated using Cohen’s d for each outcome variable between baseline and a) post
77
program (8 weeks), b) three-month, and c) six-month follow-up (T1-Tx / T1 SE * √N). 2)
A standard mean difference between baseline and follow-up time points ≥ 1 standard
error of the measure (SEM) (138). SEM was calculated as baseline (T1 SE * √n) * (√1 –
Cronbach’s α). 3) Comparing changes from pre to post to meaningful clinically important
difference in the research literature (135). Pseudo R2 statistics for each model were
calculated as unconditional model residual variance – trimmed model residual variance /
unconditional model residual variance. These statistics indicate the reduction in residual
variance between the unconditional and trimmed model and provide an estimate of effect
size similar to traditional OLS regression (202). These effects can be interpreted using
Cohen’s criteria of >.02 as a small effect >.13 as a moderate effect and >.26 as a large
effect (142). All statistical significance levels were set at p = .05 with a two-tailed test.
3.3.5.3 Acute Effects (objectives 1a & 1b). Estimated marginal means models
were created to assess overall changes in valence, activation and attention pre-post each
yoga class (objective 1a). Multilevel regression analyses were then conducted to assess
concurrent associations between valence, activation and attention and the moderating
effects of baseline demographics, physical activity, previous yoga experience, beliefs
about yoga, affect, mindfulness, and / or psychological functioning (objective 1b).
3.3.5.4 Longitudinal Effects (objectives 2a & 2b). Estimated marginal means
models were created to assess overall change in outcome, predictor, and continuous
covariates longitudinally, pre-post and at three-and six-month follow-ups (objective 2a).
Multilevel regression analyses were conducted to assess concurrent associations between
mood disturbance, stress symptoms and HRQL, proposed predictor variables including
78
affect and mindfulness, and covariates including demographics, time since diagnosis,
physical activity, previous yoga experience and beliefs about yoga (objective 2b).
3.3.5.5 Yoga Practice Maintenance (objectives 3a & 3b). Associations of
predictor and outcome variables with program maintenance (ongoing yoga participation
in either the yoga for cancer survivors program, community-based yoga programming or
engaging in home practice) was assessed via logistic generalized estimating equations
(GEE) (203). In this context, GEE take into consideration the within-subject relationships
between predictor and outcome variables. Yoga maintenance was dichotomized as either
0 – no yoga or 1 – ongoing yoga. To determine whether there was a difference between
those who practiced yoga and those who did not at each time point, an initial estimated
marginal means model with no predictors was run (objective 3a). Logistic GEE were
then run to determine associations between yoga practice and predictor variables
(objective 3b). Individual covariates for age, time since diagnosis and previous yoga
experience were entered individually as main effects, followed by physical activity,
affect, mindfulness and health outcomes (data not shown). A final trimmed model was
developed by entering all significant predictors and their interactions to test overall
prediction of yoga practice maintenance across time (exclusion p >.1).
3.4 Results
Results are reported in the following order: demographics: brief overview of
descriptive participant demographics; acute effects: estimated marginal means models
examining program-averaged pre-post class change in measures of valence, activation
and attention (objective 1a); multilevel regression analyses examining predictors of prepost class change in valence, activation, and attention (objective 1b); longitudinal effects:
79
estimated marginal means models examining longitudinal changes in mood disturbance,
stress symptoms and HRQL, predictor variables including affect and mindfulness, and
covariates (objective 2a), multilevel regression analyses examining predictors of
longitudinal change in mood disturbance, stress symptoms and quality of life (objective
2b); yoga practice maintenance: longitudinal GEE logistic estimated marginal means
models of yoga practice pre-post seven-week yoga program and at three-and six-month
follow-ups (objective 3a), longitudinal GEE logistic regression analyses examining
predictors of yoga practice maintenance over time (objective 3b)
3.4.1 Demographics
The average participant was approximately 53 years of age. The study sample was
90% female and 62.1% of participants had received a breast cancer diagnosis, stage II-III
approximately two years prior to study enrollment. Most participants were married
(66.7%), highly educated (54.5% had completed university / college) and affluent (60.6%
had a combined household income > $80,000 per annum). Many participants (39.4%) had
returned to work fulltime. Participants attended an average five of seven yoga sessions
(see Table 3.1).
3.4.2 Acute Class Effects (objectives 1a and 1b)
Estimated marginal means models were calculated pre-post each class for selfreported measures of valence, activation and attention based on a combined average of all
seven-weeks of the Yoga Thrive program (objective 1a). Participants reported large
increases in valence and attention and small increases in activation over the seven-week
program (see Table 3.2, Figure 3.3). Multilevel regression analyses were then computed
for valence, activation and attention (objective 1b).
80
3.4.2.1 Valence. Increases in valence pre-post yoga session over the seven-week
program were significantly associated with the linear passage of time. Those with higher
self-reported tiredness and mood disturbance at baseline reported lower valence pre-post
class. In addition, those who reported higher dissociative attention reported lower
valence, while those who reported higher activation reported higher valence. The Pseudo
R2 value suggests inclusion of these variables reduced error in predicting valence pre-post
class by 83% (see Table 3.3).
3.4.2.2 Activation. Increased activation was also associated with the acute passage
of time pre-post class. However, the significant linear effect of time was eclipsed by the
inclusion of valence in the model (p= .499, unconditional growth model not shown).
Those with higher valence concurrently reported higher activation pre-post class.
Conversely, those with higher scores in tiredness and the ability to be non-judgemental of
inner experience at baseline reported lower activation pre-post class. The Pseudo R2 value
suggests inclusion of these variables reduced error in predicting activation pre-post class
by 34% (see Table 3.3).
3.4.2.3 Attention. An acute decrease in dissociative attention was observed prepost class. Higher reported tension at baseline was associated with increased dissociative
attention pre-post yoga class. Higher reported baseline valence, and ability to describe
events and act with awareness were associated with decreased dissociative attention prepost yoga class. Interestingly, higher mood disturbance at baseline was also associated
with lower dissociative attention pre-post class. The Pseudo R2 value suggests inclusion
of these variables reduced error in predicting attention by 68% (see Table 3.3).
81
3.4.3 Longitudinal Program Effects (objectives 2a and 2b)
Estimated marginal means models were computed for covariates including age,
time since diagnosis, physical activity, and previous yoga experience, predictor variables
including affect and mindfulness, and outcome variables including mood disturbance,
stress symptoms, and HRQL (objective 2a).
3.4.3.1 Covariates. Participants had already completed Yoga Thrive 1.46 times
prior to study enrollment. Most participants reported completing the Yoga Thrive
program at eight-weeks, and had completed the program an additional time by the six
month follow-up. Small significant increases from baseline in moderate-vigorous PA
were observed at the three- and six-month follow-ups. Participants increased their
moderate- to vigorous PA from 93 to 121 minutes per week over the course of the study,
still less than the recommended 150 minutes / week of vigorous to moderate PA (23), and
these increases did not exceed the SEM of the measure. A small significant increase in
total physical activity (LSI) was also observed at the six-month follow-up (see Table 3.4,
Figure 3.4).
3.4.3.2 Affect. Small increases in energy were observed post-program that
became more pronounced at the three- and six-month follow-up. Small decreases in
tiredness post-program were not maintained at the 3-month follow-up but again improved
at the 6-month follow-up. Small decreases in tension were observed post-program, were
not maintained at the three-month follow-up, improved at the six-month follow-up, but
did not exceed the SEM. Small increases in self-reported calmness were observed prepost intervention that approached significance but this increase abated over time (see
Table 3.4, Figure 3.5).
82
3.4.3.3 Mindfulness. Small significant increases in observational skills were
observed only at the six-month follow-up. However they did not exceed the SEM. No
changes in participants’ ability to describe were observed throughout the program.
Moderate increases in participants’ ability to act with awareness were observed postprogram that were maintained at the three- and six-month follow-ups. Small increases in
participants’ ability to be non-judgmental of inner experience were observed postprogram that were maintained at the three- and six-month follow-up. However, the
estimated difference at the 3-month follow-up did not exceed the SEM. No significant
changes in non-reaction were observed post program but small significant improvements
were observed at the three- and six-month follow-ups. However, neither change exceeded
the SEM (see Table 3.4, Figure 3.6).
3.4.3.4 Mood Disturbance, Stress Symptoms and Quality of Life. Decreases in
mood disturbance were moderate post program. Reductions in mood disturbance were not
maintained at the three-month follow-up but improved to post-program values at the sixmonth follow-up. A moderate decrease in stress symptoms was observed from baseline to
post program that was maintained at both the three- and six-month follow-ups. A small
statistically significant improvement in HRQL was observed post program that did not
exceed the SEM. However, HRQL significantly improved at the three-month follow-up
and was maintained at the six-month follow-up (see Table 3.4, Figure 3.7).
Follow-up multilevel regression analyses were conducted to explore the
associations between mood disturbance, symptoms of stress, and HRQL, predictor
variables including affect and mindfulness and moderating variables including age, time
83
since diagnosis, physical activity, previous yoga experience and yoga beliefs (objective
2b).
3.4.3.5 Mood Disturbance. Significant linear, quadratic, and cubic effects were
observed for time. Those with higher beliefs in yoga at baseline reported lower mood
disturbance at all time points. Those who reported higher levels of overall PA, higher
energy, higher levels of the ability to act with awareness and be non-reactive to inner
experience also reported lower mood disturbance. Those who reported greater tiredness
reported greater mood disturbance. The Pseudo R2 value suggests inclusion of these
variables reduced error in predicting mood disturbance by 31% (see Table 3.5).
3.4.3.6 Stress Symptoms. There were significant linear and quadratic effects of
time, indicating an initial decline in stress symptoms that slowed over time. Higher
positive beliefs about yoga at baseline were associated with lower symptoms of stress.
Those who reported higher moderate to vigorous PA and a greater ability to act with
awareness and be non-judgmental of inner experience concurrently reported lower stress
symptom scores. Higher tension was significantly associated with higher stress
symptoms. The Pseudo R2 value suggests inclusion of these variables reduced error in
predicting stress symptoms by 40% (see Table 3.5).
3.4.3.7 Quality of Life. A significant linear increase in HRQL was observed prepost yoga intervention and at three- and six-month follow-ups. Higher baseline beliefs
about yoga were associated with higher HRQL. Those who reported higher overall PA,
energy and the ability to be non-judgmental of inner experience reported higher HRQL.
Those who reported high levels of tension reported lower levels of HRQL. The Pseudo
84
R2 value suggests inclusion of these predictor variables reduced error in predicting HRQL
by 29% (see Table 3.5).
3.4.4 Yoga Practice Maintenance (objectives 3a and 3b)
Longitudinal GEE logistic estimated marginal means models were conducted
examining yoga practice pre-post yoga program and at three-and six-month follow-ups
(objective 3a). Upon initiation of the current study, 48% (32 participants) reported
previous yoga practice experience either through the Yoga Thrive program, community
classes, home practice or combinations thereof. At the end of the seven-week yoga
program, 96% were still practicing yoga in various settings, a significant increase of 48%
from baseline. At the three-month follow-up yoga practice had dropped, with 69% of
participants were still practicing yoga in various settings, a significant 20% increase from
baseline. At the six-month follow-up there was a slight increase in participation rates,
with 76% of participants reporting continued yoga practice in various settings, a
significant 27% increase from baseline (see Table 3.6, Figure 3.8).
Follow-up GEE logistic regression analyses were performed to examine what
participant characteristics were related to ongoing yoga practice (objective 3b). In
examining associations between yoga practice at each time point and predictor variables,
there was an initial positive linear effect for time, reflecting the increased rate of
participants reporting yoga practice pre-post program. The quadratic trend reflects a
significant decline post-intervention completion in yoga practice at three- and sixmonths. Those who reported more frequent participation in the Yoga Thrive program at
each time had an 88% greater chance of continuing their yoga practice for each time they
completed the Yoga Thrive program. In addition, those who reported higher energy at
85
each time point were 14% more likely to maintain a yoga practice for each one unit
increase in energy, as were those who reported a greater ability to be non-reactive to
inner experience. These individuals were 8% more likely to continue yoga practice for
each one unit increase in non-reactive ability. Finally, those with higher self-reported
HRQL at each time point had a 3% greater chance for each one unit increase in HRQL to
continue yoga practice (see Table 3.7).
3.5 Discussion
Despite an emerging body of evidence highlighting the benefits of yoga for cancer
survivors, little work has been done to bridge these clinical findings with theoretical
mechanisms of action. The current research seeks to bridge this gap between determining
clinical benefits and describing the mechanisms underlying these improvements in mood
disturbance, stress symptoms and HRQL. Using the Circumplex Model of Affect as it has
been applied in exercise settings (73) and emerging work examining the mechanisms of
mindfulness (94,181,185,204-207) and yoga (38,45,109) a series of multilevel models
were developed to examine: 1) acute changes pre-post yoga class in measures of valence,
activation and attention in a seven week intervention and potential moderators of change
(objectives 1a and 1b); 2) longitudinal changes in mood disturbance, stress symptoms and
HRQL, predictor variables including affect and mindfulness, and whether these predictor
variables moderated change in psychological outcomes (objectives 2a and 2b); and 3)
maintenance of yoga practice and whether maintenance could be predicted by moderating
variables (objectives 3a and 3b).
86
3.5.1 Acute Class Effects (objectives 1a and 1b):
Based on average scores compiled over seven sessions, acute increases in valence,
activation and associative attention were observed. These findings are consistent with
studies demonstrating that a single exercise session elicits positive core affect (73) as
well as increases in associative attention (89,95) (See Figure 3.3). This high-activation
state of pleasant affect has been variously referred to as energy, pleasant activation,
positive activation and positive activated effect (72). Findings in the field of aerobic
exercise suggest this state is facilitated immediately post-exercise for intensity-duration
doses considered low to moderate (72). The 75-minute duration of the Yoga Thrive
sessions exceeded this definition of moderate duration (up to 35 minutes). However, it
must be kept in mind that yoga is not steady state exercise per se, but rather varies in
intensity. Specifically, during active yoga sequences, heart rate and respiration may
increase consistent with exercise (114), while during the restful supine sequences and
meditation that conclude most yoga classes, activity is consistent with relaxation and
reduced physiological arousal (115).
Findings suggest focus of attention shifts pre-post yoga class to a primarily
associative state, characterised by inward attention on kinesthetic and interoceptive cues
(89). This associative attentional state has been described as a rough correlate of states of
mindfulness at the acute level (95). A more specific corollary may be what Lutz and
colleagues (208) have called “focused attention,” a state characterized by voluntary focus
of attention on a chosen object. In the case of yoga, this often consists of awareness of the
coordination of breath and movement (97).
87
In examining the multilevel regression analyses of these acute increases in
valence, activation and attention, baseline measures of mood disturbance are associated
with poorer rating of valence and higher associative attention both pre-post class. This
association suggests higher baseline mood disturbance makes one increasingly aware of
not feeling good. This association fits within the rubric of the notion of “mood
congruency” (81) whereby based on our existing mood we are cognitively-primed to
attend to stimuli that reiterate these feelings. Higher tiredness (fatigue) at baseline is
associated with decreased valence and activation pre-and post yoga class, while those
who report higher tension at baseline report higher dissociative attention pre-post class.
Taken together, these findings highlight how underlying mood and affective states can
impact acute measures of core affect and attention.
Those who reported a greater ability to describe events and act with awareness at
baseline reported higher associative attention. The ability to accurately describe feelings
and thoughts may be related to more associative attentional states while acting with
awareness, which focuses on attending to activities in the present moment, is highly
congruent with the construct of associative attention (180). Their association likely
reflects this overlapping meaning. Those who reported higher non-judgemental
awareness at baseline reported lower activation pre-post yoga session. Non-judging of
inner experience is associated with acceptance of present-moment experience (181). This
openness to present moment experience has been associated with reduced arousal in
experienced meditators (209).
88
3.5.2 Longitudinal Program Effects (objectives 2a and 2b):
This research confirms previous work suggesting yoga is clinically efficacious in
reducing psychological symptoms and improving HRQL (1,2,33). Moderate clinicallysignificant improvements in both mood disturbance and stress symptoms and small
improvements in HRQL were seen post-program completion. Continued significant
changes in all three measured health outcomes were observed, despite the fact that 48%
of participants at baseline were already yoga practitioners and had arguably derived
initial program benefits. However, without a comparison group, it is impossible to
determine whether these positive effects were directly caused by the yoga program, or
may have been due to the simple passage of time.
Affect: While the ADACL as a measure of Circumplex Affect has been used in
primarily acute exercise settings (73), findings in the area of regular aerobic training (71)
suggest positive improvements in areas of energy (high-activation positive affect) over
time, provided the exercise is of lower intensity and duration. Current study results
suggest a linear increase in energy over time, consistent with acute increases in valence
and activation pre-post each yoga class. Given the burden of fatigue in cancer survivors,
these improvements are important and corroborate corollary improvements in fatigue
indices in both the cancer exercise (9,19,210), mindfulness (179,211) and yoga literature
(1). Reductions in tension were most prominent post-program completion. Interestingly
no changes in levels of calmness longitudinally were observed. This is of interest, as yoga
is often associated with states of increased calmness versus increased energy (33).
Mindfulness: Scores on the FFMQ at baseline were similar to those reported by
Baer for a non-meditating community sample (212), with the exception of the ability to
89
act with awareness and non-judgment of internal experience, which were lower in the
current sample. This finding was somewhat surprising, given associations made between
mindfulness and yoga (96) and the fact 48% of the current sample had an existing yoga
practice. In comparison to Branstrom’s study examining mindfulness as a mediator in
cancer patients (175,182), in the current study baseline mindfulness scores were
comparatively lower at all time points, with the exception of a higher score in nonreaction to inner experience. Baseline mindfulness scores were comparable to those of
Garland’s (183) study examining relations between mindfulness, stress and mood in those
completing an eight-week MBSR course in cancer settings, with the exception of a lower
reported ability to act with awareness and be non-judgmental of inner experience.
In the present study, overall mindfulness effect sizes pre-post intervention
indicate moderate clinically significant increases in the ability to act with awareness and
small increases in the ability to be non-judgmental of inner experience. Similar
improvements in these mindfulness facets were reported by Garland (183) and Branstrom
(175,182) though these effects were of a larger magnitude and also included
improvements in the other facets of mindfulness. While improvements in facets of
mindfulness in the current study were associated with mood disturbance, stress and
HRQL, these changes were to a lesser degree than mindfulness-based programs in which
intensive mindfulness practice is central.
Mood Disturbance, Stress Symptoms and Quality of Life: Baseline positive yoga
beliefs were consistently associated with lower mood disturbance, stress symptoms and
higher HRQL at all time points. Sohl and colleagues (187) suggest these beliefs are
intrinsic to both initial program engagement and reported health outcomes. In addition,
90
those who reported higher baseline physical activity reported lower mood disturbance,
stress symptoms and higher HRQL at all time points. These findings reflect the large
body of research suggesting the effects of exercise on mental health and HRQL in cancer
survivors (210,213). Higher energy at all time points was consistently associated with
lower mood disturbance and higher quality of life. Higher tension was associated with
higher stress symptoms and lower HRQL, while higher tiredness was associated with
increased mood disturbance. These findings are consistent other research suggesting
dimensional measures of affective states may mediate more categorical mood states
(68,214,215).
Higher reported ability to act with awareness was associated with decreased stress
and mood disturbance. As mentioned previously, this ability to attend to events in the
present moment is most strongly correlated with associative attentional strategies and
may have been related to participant yoga practice. Higher non-judgment of inner
experiences relation to lower stress symptoms and higher HRQL may reflect an accepting
attentional strategy in which participants have found alternate ways to respond to
stressful situations (94). Higher non-reactivity to inner experience was related to lower
mood disturbance and has been implicated as a primary mediator in the improvement of
overall psychological wellbeing in a mixed sample of university students and experienced
meditators (216).
In sum, observed improvements in mood disturbance, stress and HRQL are
encouraging given participants had already completed the program 1.5 times on average
at baseline and thus while they had potentially already derived some benefit, they were
still experiencing program-related improvements. Findings suggest improvements in
91
health outcomes are associated with beliefs about yoga, physical activity, the circumplex
model of affect, particularly the dimensions of energy, tension and tiredness, and facets
of mindfulness, including the ability to act with awareness, be nonjudgmental of and nonreactive to inner experience. Associations between positive affective states and
mindfulness are becoming clearer in the current research literature as state and trait
measures of both constructs have been found to be highly related (170). Moreover, this
examination of distinctive elements of mindfulness may ultimately provide novel views
concerning the role of positive affect in exercise and other behavioral interventions (217).
3.5.3 Yoga Practice Maintenance (objectives 3a & 3b):
Participants reported a high level of adherence to the yoga program, with a 96%
adherence rate post-program. It should be emphasised that at baseline, 48% had already
self-identified as having some form of regular yoga practice, either as a previous Yoga
Thrive program participant, enrolled in community classes, engaged in home practice or
combinations thereof. This rate dropped from 96% to 69% at the three-month follow-up.
However, 76% of participants were engaged in yoga practice at the subsequent six-month
follow-up. Previous participation in the Yoga Thrive program was a significant predictor
of yoga practice maintenance at all time points. This finding is similar to Speed-Andrews
and colleagues’ finding that yoga experience over the past year was significantly related
to current yoga adherence (218). Positive experiences and increased wellbeing derived
from yoga classes are suggested mechanisms for this increased program adherence (219).
Those reporting higher energy at each time point were also more likely to
continue yoga practice. This relationship between positive affect and adherence has been
widely reported in the literature (71,76,167). It is hypothesised the positive appraisal of
92
physical activity as pleasurable is likely to bolster participant self-efficacy and lead to
increased adherence and maintenance (66). Recent ACSM guidelines also suggest
positive affect may be an important determinant of exercise adherence (74).
Results suggesting those higher in non-reactivity were more likely to adhere to
yoga practice are supported in the literature. Research by Ulmer and colleagues (220)
suggests those with higher mindfulness scores are less reactive and better able to
accurately appraise and respond to stressors that may impact routine activities, including
yoga practice. Non-reaction refers specifically to the ability to remain calm in the face of
distressing thoughts and emotions. This ability to be non-reactive to inner experience has
salutary effects and seems to be a key facet of mindfulness as well as one of the facets
most effected by contemplative training (216).
Finally higher HRQL has been associated with increasing individual self-efficacy
to adhere and maintain exercise programs (66). Taken together these results suggest
previous yoga experience coupled with higher energy, a greater reported ability to be
non-reactive to inner experience and higher HRQL interact to improve the likelihood
participants will maintain yoga practice. If affect and mindfulness are both associated
with improved HRQL and mental health outcomes and cited as leading to increased
intervention adherence, it behooves clinicians to consider developing approaches that
focus more exclusively on the inherent psychological benefits of exercise (73,74).
3.5.4 Limitations:
A significant limitation of the current study is the lack of a measure of home
practice. Current work (221) suggests the addition of home practice predicts mindfulness,
subjective wellbeing, fatigue, sleep and a variety of behavioral outcomes. In addition
93
frequency of yoga home practice was a better predictor of health than either total years of
practice or class frequency. Given these findings, the thorough assessment of yoga
practice outside of class time is imperative. Additional limitations include the self-report
nature of all outcomes. While markers of clinical significance have been introduced for
these outcomes, including effect sizes and the SEM, future research would be better
served by including more objective anchor-based methodology as well as supportive
biomarkers of change (62). In addition, including participants who had already completed
the program added complexity to the current analyses. However this complexity is
inherent to community-based research. Multilevel analyses allowed for the integration of
theory to develop novel research designs able to evaluate program outcomes, contextual
factors and interactions thereof within existing clinical treatment programs (63,222).
3.5.5 Conclusion
Before yoga can be broadly applied within psychosocial oncology, carefully
designed and executed research that convincingly evaluates not only the efficacy of yoga
in clinical settings but also posits potential mechanisms of action underlying these
interventions is required (51). The current research provides a foundation to more fully
explore the applications and benefits of yoga for cancer survivors. These objectives have
been examined within the structure of the existing Yoga Thrive program to explore
whether these associations are observed in routine, everyday practice. Examining these
proposed theoretically-based mechanisms for yoga’s salutary effects enables a broader
understanding not only of “if” yoga works, but also “how.” This combined knowledge
can then be transferred directly back into the community to further develop innovative
94
yoga programs and best practices with the express aim of improving psychosocial health
and HRQL in cancer survivors.
95
Table 3.1 Demographics
Baseline (n = 66)
Age, years (SD)
Cancer Diagnosis – Breast
Cancer Stage
II
III
Time Since Diagnosis, months
Gender – Female
Marital Status – Married / Common-law
Education Level – completed university / college
Annual Household Income - > $80,000
Employment Status – Fulltime
Program attendance, sessions
n (%)
52.88 (8.11)
41 (62.1)
22 (33.3)
17 (25.8)
23.93 (21.93)
60 (90.9)
44 (66.7)
36 (54.5)
40 (60.6)
26 (39.4)
5.08 (1.94)
96
Table 3.2 Pre-Post Class Estimated Marginal Means Models
Pre Class
T1 (n=66)
Outcome Variable
Measure
Post Class
T2 (n=66)
T2 –T1 SMD
Mean
(SE)
Mean
(SE)
Mean
Difference
(SE)
p
d
1.16
(0.18)
3.07
(0.18)
1.91
(0.15)
.000
1.59
3.29
(0.12)
3.69
(0.12)
0.40
(0.14)
.006
0.36
5.51
(0.18)
3.96
(0.18)
-1.55
(0.21)
.000
0.95
Valence
-Feeling Scale
Activation
-Felt Arousal Scale
Attention
-AssociationDissociation
Scale
SMD = standard mean difference, SE = standard error, p = significance, d = Cohen’s d
Predictor
Est. (SE)
1.69 (0.33)
Valence
Activation
(Feeling Scale)
(Felt Arousal Scale)
Est. (SE)
df
t
p
df
107.64 5.10 .000 3.37 (0.09) 56.95
t
35.63
p
.000
Attention
(Association-Dissociation Scale)
Est. (SE)
df
t
p
5.15 (0.15) 55.50 33.85 .000
Intercept
Time
Linear Time
1.31 (0.17) 81.49 7.50 .000 0.12 (0.18) 93.08 0.68 .499 -0.48 (0.25) 98.83 -1.89
Baseline Outcomes
Mood
-0.02 (0.004) 55.24 -3.63 .001
-0.02 (0.01) 61.95 -2.27
Disturbance
Baseline Affect
Tension
0.08 (0.04) 55.78 1.92
Tired
-0.08 (0.02) 57.88 -3.43 .001 -0.04 (0.02) 62.10 -1.98 .052
Baseline Mindfulness
Describe
-0.06 (0.02) 55.54 -2.85
Act w/
-0.04 (0.02) 55.60 -1.85
Awareness
Non-judgement
-0.04 (0.01) 54.82 -2.73 .009
Concurrent Measures
Valence
NA
NA
NA
NA 0.13 (0.05) 72.65 2.30 .024 -0.54 (0.08) 86.24 -6.09
Activation
0.29 (0.08) 72.74 3.52 .001
NA
NA
NA
NA
Attention
-0.32 (0.06) 92.83 -5.46 .000
NA
NA
NA
Pseudo R2
Valence = 0.83
Activation = 0.34
Attention = 0.68
Variables Excluded (p > .1): age, time since diagnosis, previous Yoga Thrive experience, yoga beliefs (BAYS) baseline physical
activity (GLTEQ), energy and calmness (ADACL), observe &non-reaction (FFMQ), stress symptoms (SOSI), quality of life
(FACTG). Est. = Estimate, SE = standard error, df = degrees of freedom; p = significance.
.062
.026
.059
.006
.070
.000
NA
Table 3.3 Acute Associations – Pre-Post Class Multilevel Regression Model
97
97
Baseline
T1 (N=66)
8-weeks post
T2 (n=53)
Measures
3-month follow-up
T3 (n=48)
T2 –T1 SMD
6-month follow-up
T4 (n=45)
T3 – T1 SMD
T4 – T1 SMD
Cronbach
α
SEM
Mean
(SE)
p
d
Mean
(SE)
p
d
Mean
(SE)
p
d
20.20
(2.46)
.93
5.25
11.58
(2.61)
.000
-0.47
15.01
(2.67)
.036
-0.27
11.63
(2.74)
.001
-0.44
59.08
(3.10)
.91
7.49
47.41
(3.26)
.000
-0.63
46.78
(3.31)
.000
-0.64
49.11
(3.35)
.000
0.51
73.51
(1.98)
.93
4.15
76.71
(2.06)
.017
0.30
78.97
(2.09)
.000
0.50
79.66
(2.12)
.000
0.54
0.61
2.45
(0.29)
.000
1.00
3.02
(0.30)
.000
1.37
3.49
(0.30)
.000
1.73
Mean
(SE)
Outcome Variables
Mood Disturbance
-POMS Total
Stress Symptoms
-SOSI Total
Quality of Life
-FACT G Total
Predictor Variables
Program Experience (Yoga Thrive)
1.46
.93
- Enrolment
(0.29)
(ICC)
Physical Activity
-LSI
21.70
(2.03)
.88
5.66
23.39
(2.18)
.388
0.11
25.39
(2.22)
.066
0.23
27.83
(2.26)
.003
0.38
-GLTEQ
92.94
(13.64)
.86
41.13
96.50
(14.52)
.788
0.03
121.74
(15.27)
.042
0.26
120.48
(15.48)
.055
0.24
Table 3.4 Longitudinal Estimated Marginal Means Model
98
98
Baseline
T1 (N=66)
8-weeks post
T2 (n=53)
Measures
T2 –T1 SMD
Mean
(SE)
Cronbach
α
SEM
.88
1.45
.84
1.74
.79
1.59
.80
1.54
Predictor Variables
Affect
-Energy
-Tired
-Tension
-Calm
Mindfulness
3-month follow-up
T3 (n=48)
9.53
(0.52)
11.60
(0.54)
9.35
(0.43)
12.77
(0.43)
Mean
(SE)
10.88
(0.57)
10.16
(0.59)
7.66
(0.47)
13.77
(0.47)
p
d
.022
0.29
.023
-0.29
.001
-0.43
.054
0.24
T3 – T1 SMD
Mean
(SE)
11.86
(0.59)
10.62
(0.62)
8.47
(0.49)
13.27
(0.49)
p
d
.000
0.48
.136
-0.19
.084
-0.23
.350
0.12
24.92
25.81
26.08
.85
2.47
.244
0.15
.139
(0.79)
(0.85)
(0.87)
23.04
23.86
23.20
-Describe
.87
2.46
.243
0.16
.821
(0.85)
(0.90)
(0.92)
-Act w/
18.54
20.40
20.31
.89
1.85
.000
0.46
.001
Awareness
(0.69)
(0.73)
(0.74)
20.35
22.47
21.96
-Non-Judging
.91
2.02
.006
0.35
.041
(0.84)
(0.90)
(0.92)
19.92
20.82
21.61
-Non-Reacting
.88
1.76
.082
0.22
.002
(0.63)
(0.67)
(0.68)
T = time, SMD = standard mean difference, SE = Standard error, p = significance, d = Cohen’s d.
-Observe
6-month follow-up
T4 (n=45)
0.18
0.03
0.42
0.34
0.40
T4 – T1 SMD
Mean
(SE)
13.11
(0.60)
8.82
(0.63)
7.99
(0.50)
13.23
(0.50)
27.13
(0.90)
24.12
(0.93)
21.23
(0.75)
22.65
(0.94)
21.51
(0.69)
p
d
.000
0.72
.000
-0.52
.009
-0.33
.399
0.10
.007
0.34
.151
0.18
.000
0.62
.005
0.36
.004
0.36
Table 3.4 – continued
99
99
Predictor
Intercept
Time
Linear Time
Time2
Time3
Moderators
Beliefs about Yoga
Physical Activity
Leisure Score Index
Moderate + Vigorous
PA
Affect
Mood Disturbance
(POMS Total)
Est. (SE)
df
16.98
76.35
(2.10)
-5.36
(2.22)
1.59
(0.74)
-0.12
(0.06)
t
8.09
Stress Symptoms
(C-SOSI Total)
Est. (SE)
p
df
54.89
.000
69.19
(2.40)
91.15
-2.41 .018
86.47
2.16
85.68
-1.96 .053
-0.57
(0.22)
52.55
-2.63 .011
-0.17
(0.08)
161.15 -2.17 .032
-
-
-
.033
-
Energy
-0.67
(0.34)
Tension
-
-
-
-
0.75
(0.31)
176.96
2.45
.015
Tired
178.62 -1.96 .052
-3.90
(0.97)
0.41
(0.11)
Quality of Life
(FACTG Total)
p
Est. (SE)
t
df
76.95
22.87 .000
60.94
(1.49)
t
p
51.53 .000
108.75 -4.04
.000
0.27
(0.16)
60.05
1.65
.104
90.10
3.62
.000
-
-
-
-
-
-
-
-
-
-
-
-0.79
(0.27)
50.62
-2.87
.006
0.42
(0.19)
57.19
2.23
.030
-
-
-
-
0.14
(0.05)
173.61
2.86
.005
.037
-
-
-
-
158.34
2.62
.010
150.36 -4.27
.000
-0.03
(0.01)
-
174.50 -2.10
-
1.35 (0.33) 158.74
-
-
-
-
4.07
.000
-
-
0.44
(0.17)
-0.81
(0.19)
-
-
-
-
Table 3.5 Longitudinal Multilevel Regression Model
100
100
Predictor
Mindfulness
Act w/ Awareness
Non-judgement
Non-Reaction
Mood Disturbance
(POMS Total)
Est.
df
(SE)
-0.58
(0.25)
-0.56
(0.26)
t
p
144.80 -2.33 .021
-
-
-
140.04 -2.13 .035
Stress Symptoms
(C-SOSI Total)
Est.
df
(SE)
-0.67
(0.29)
-0.76
(0.24)
-
t
p
Quality of Life
(FACTG Total)
Est.
df
(SE)
t
p
174.16 -2.29
.023
-
-
-
-
185.89 -3.20
.002
0.56
(0.12)
180.77
4.71
.000
-
-
-
-
-
-
-
Pseudo R2
Mood Disturbance = 0.31
Stress Symptoms = 0.40
Quality of Life = 0.29
Variables Excluded (p>.1): age, time since diagnosis, previous Yoga Thrive program experience, calm (ADACL), observe & describe
(FFMQ). Est. = estimate, df = degrees of freedom, p = significance.
Table 3.5 – continued
101
101
Baseline
T1 (N=66)
8-weeks post
T2 (n=53)
T2 –T1 SMD
Measures
Adherence
3-month follow-up
T3 (n=48)
6-month follow-up
T4 (n=45)
T3 – T1 SMD
T4 – T1 SMD
Mean
(SE)
Mean
(SE)
Mean
Difference
(SE)
p
d
Mean
(SE)
Mean
Difference
(SE)
p
d
Mean
(SE)
Mean
Difference
(SE)
0.48
(0.06)
0.96
(0.03)
0.48
(0.07)
.000
0.98
0.69
(0.07)
0.20
(0.08)
.008
0.41
0.76
(0.06)
0.27
(0.09)
p
d
.002 0.55
T = time, SMD = standard mean difference, SE = standard error, p = significance, d = Cohen’s d.
Table 3.6 Yoga Practice Maintenance Estimated Marginal Means Model
102
102
103
Table 3.7 Predictors of Yoga Practice Maintenance.
B
Wald
Odds Ratio
Parameter
p
2
(SE)
χ
(95% CI)
1.50
4.50
(Intercept)
15.77
.000
(0.38)
(2.14, 9.45)
0.63
1.87
Time
7.11
.008
(0.23)
(1.18. 2.96)
-0.09
0.91
Time2
10.83
.001
(0.03)
(0.86, 0.96)
Previous Participation in Yoga
0.63
1.88
8.42
.004
Thrive
(0.22)
(1.23, 2.87)
0.13
1.14
Energy
4.94
.026
(0.06)
(1.02, 1.27)
0.08
1.08
Non-Reaction to Inner Experience
5.86
.015
(0.03)
(1.02, 1.16)
0.03
1.03
Quality of Life
4.23
.040
(0.01)
(1.00, 1.06)
Excluded variables (p>.1): age, time since diagnosis, yoga beliefs, attendance, physical
activity, tension, tired, calm, observe, describe , act with awareness, non-judgment, mood
disturbance, stress symptoms. B = estimate, SE = standard error, p = significance.
Figure 3.1 Circumplex Model of Affect
104
104
Figure 3.2 Participant Recruitment Flow Diagram
105
105
Figure 3.3 Affect Circumplex 2D Model (valence, activation)
106
106
Figure 3.4 Affect Circumplex 3D Model (valence, activation, affect, attention)
107
107
Figure 3.5 Continuous Moderators: total physical activity (LSI), moderate-vigorous physical activity (GLTEQ) and previous Yoga
Thrive experience († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
108
108
Figure 3.5 Continuous Moderators (continued): total physical activity (LSI), moderate-vigorous physical activity (GLTEQ) and
previous Yoga Thrive experience – continued († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
109
109
Figure 3.6 Longitudinal Changes in Affect (AD ACL) († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
110
110
Figure 3.6 Longitudinal Changes in Affect (AD ACL) – continued († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
111
111
Figure 3.7 Longitudinal Changes in Mindfulness (FFMQ) († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
112
112
Figure 3.7 Longitudinal Changes in Mindfulness (FFMQ) – continued († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
113
113
Figure 3.7 Longitudinal Changes in Mindfulness (FFMQ) – continued († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
114
114
Figure 3.8 Longitudinal Changes in Mood Disturbance (POMS), Stress Symptoms (C-SOSI), and Quality of Life (FACT-G)
(† p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
115
115
Figure 3.8 Longitudinal Changes in Mood Disturbance (POMS), Stress Symptoms (C-SOSI), and Quality of Life (FACT-G) –
continued
(† p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
116
116
Figure 3.9 Yoga Practice Maintenance († p<0.10, * p<0.05, ** p<0.01, *** p<0.001).
117
117
118
CHAPTER 4: MODELLING ASSOCIATIONS BETWEEN AFFECT,
ATTENTION AND HEART RATE VARIABILITY IN A SINGLE YOGA
SESSION FOR CANCER SURVIVORS: A NEUROPHENOMENOLOGICAL
APPROACH
119
4.1 Introduction
Contemporary yoga practice consists of lifestyle prescriptions, postures, breath
regulation, and meditative techniques modifiable based on desired outcomes as well as
participant health status (51). Yoga practices are said to have numerous applications,
ranging from liberation or kaivalya, as per the Yoga Sutras (YS IV.34), to ordinary
physical and mental training (55).
Emerging research suggests yoga shares many of the benefits of exercise in
addition to yoga-specific psychophysiological control (38,223). Studies comparing the
effects of yoga and exercise indicate that in healthy individuals and those with various
health conditions, yoga may be as effective as more contemporary western forms of
exercise including walking, jogging, cycling and aerobics at improving a variety of
health-related outcome measures (48). Within cancer settings yoga groups compared to
waitlist control groups or supportive therapy groups show greater improvements in
overall QOL, psychological health, stress-related symptoms, and fatigue indices (1,2,33).
The root aim of yoga can be summarised in Patanjali’s definition of Yoga as a
path towards, “stilling the fluctuations of the mind (YS I.2).” Importantly, meditative
states experienced within yoga practice are correlated to neurophysiological systems that
mediate both positive affect and internalised attention (58,59). Based on this definition of
yoga, the aims of the current study were threefold: 1) to investigate whether yoga practice
elicits change in affect, attention, perceived exertion and cardiac activity including heart
rate variability (HRV), a non-invasive metric of cardiac autonomic function, during a
single yoga session; 2) to determine the mechanisms by which these changes were
elicited. Specifically, whether changes in affect, attention, and perceived exertion were
120
associated with one another and cardiac activity; and 3) whether participants own
subjective experience of the yoga session corroborated these quantitative findings.
4.1.1 Affect
A current area of clinical research interest is the study of positive affect as an
independent, adaptive pathway in the cancer experience (68). Engagement in behaviours
associated with state positive affect may cultivate trait-like changes that correspond with
enhanced QOL and mental health indices (69). In general, positive affect tends to
increase pre- to post-exercise following exercise intensities that are not exhaustive (73).
In two recent studies, a single session of yoga was associated with significant
improvements in positive affect and reductions in negative affect comparable to changes
seen with aerobic exercise (77,78). In addition, a recent pilot RCT reported significant
improvement in positive affect, emotional function and cognitive function, and decreased
negative affect in the yoga group as compared to supportive therapy controls (80). These
findings suggest positive affect is an important health outcome and may be correlated
with positive changes in HRQL and mental health indices.
4.1.2 Attention
Previous research has shown an individual’s focus of attention during exercise can
influence changes in affective responses (89). Associative and dissociative attentional
strategies indicate the extent to which exercisers focus their attention on exercise-related
feedback. Associative attention is defined as focusing attention on present-moment
physical sensations related to exercise, in which exercisers seek to monitor sensory input
and adjust their effort accordingly. Dissociative attention is defined as focusing on nonexercise related stimuli that divert attention away from internal sensations and present
121
moment exercise experience, such as listening to music, watching television while
exercising, or making plans for post-exercise activities (90).
The practice of mindfulness, or the ability to direct attention towards events in the
field of consciousness (thoughts, emotions, body sensations) in the present moment
without judgement (94), is conceptually similar to associative attentional processing as
described via a dimensional model, with the exception that this associative state does not
necessarily carry with it the attitude on non-judgement. Mindfulness in exercise settings
is believed to enhance accurate observation and acceptance of both one’s internal bodily
cues, the content of thoughts and the external environment (95). Higher levels of both
trait and state mindfulness have been associated with higher positive affect and lower
negative affect via engagement of sustained non-reactive attention upon immediate
experience and disengagement of habitual cognitive processing (224). Within exercise
settings, mindfulness builds upon associative attention by refining perceptions of
perceived exertion on a moment-by-moment basis via increased sensitivity to a host of
cognitive and interoceptive cues, and concurrently limits emotional reactivity to these
cues via non-reactivity to these experiences (95).
It is hypothesised the strong mind-body interaction within contemporary yoga
practice adds a unique contemplative dimension to exercise and has been referred to as
‘mindfulness in motion’ (96). The practice of yoga provides an opportunity for sustained
attention to the body, breath and mind through a progressive sequence of dynamic
movements, restful postures, breathing exercises and periods of meditative awareness.
Making these sensations the focus of attention facilitates a strong mindfulness practice.
Froeliger and colleagues (98) suggest the object of mindfulness in contemporary yoga
122
practice comes from both physical sensations of the body moving and awareness of
breath.
4.1.3 Heart Rate Variability
A commonly reported neurophysiological mechanism potentially mediating
relations between yoga practice, improved attention regulation and positive affect lies in
the Autonomic Nervous System (ANS). The ANS is the most prominent physiological
factor in determining heart functions and is comprised of both parasympathetic and
sympathetic input, and the amount of input the body is receiving from each can be
examined through the study of heart rate variability (HRV), the modulation of the heart
rate about its mean value (101).
HRV offers a non-invasive quantitative method of investigating autonomic effects
on the heart. HRV is also increasingly being used as an index of affect regulation (104106). HRV represents the sequences of complex beat-to-beat variation in heart rate
produced by the interplay of sympathetic and parasympathetic neural activity as recorded
via an electrocardiogram (ECG), known as the R-R interval. HRV measurements are
classified under three broad domains: time domains, spectral analyses and non-linear
methods. Time-domain measures plot HRV as the normal R-R ratio (N-N) over time
(225). Commonly reported time domain measures include SDNN, the standard deviation
of all N-N intervals thought to reflect overall HRV, and RMSSD, the root mean square of
successive N-N interval differences, thought to reflect vagal activity. Spectral analyses of
HRV distinguish among sources of HRV by plotting the frequency at which the length of
the N-N interval changes (225), as these rhythms occur at different frequencies (101).
Low frequency HRV (LF) occurs at a range of 0.04-0.15 Hz and reflects mixed
123
sympathetic and parasympathetic modulation of HRV and is often used as an index of
cardiac sympathetic nervous system activity (226). High frequency HRV (HF) occurs at a
range of 0.15-0.4 Hz and is widely used indicator of cardiac parasympathetic nervous
system activity (vagal tone) (227). The ratio of low-frequency to high-frequency HRV
(LF/HF) is a purported measure of sympathovagal balance (101). Non-linear methods are
purported to better measure biological mechanisms such as heart rate. Entropy measures
(approximate entropy, sample entropy) represent the total complexity of the time domain.
The larger the entropy value the more complex and healthier the system is (107). In
general, higher SDNN, RMSSD, HF HRV and sample entropy are indicative of positive
health outcomes, while higher LF HRV and LF/HF ratio are associated with negative
health outcomes (107).
In cancer settings, impaired autonomic nervous system function, as evidenced by
lower HRV, has been implicated in an increased risk of depression in metastatic breast
cancer patients (228), poorer prognosis in colorectal cancer (229), cancer-related fatigue
in a breast cancer sample (230), increased nausea post-chemotherapy (231), increased
cancer cachexia (wasting) (232) and shortened survival time in cancer patients with
advanced cancer (233-235). Conversely, exercise has been associated with higher HRV
in cancer survivors and improved autonomic function and quality of life and has been
implicated in longer survival times (236).
Yoga programs are reported to result in increased parasympathetic nervous
system (PNS) activity and significantly increases cardiac vagal modulation (108).
Streeter and colleagues (2012) hypothesise yoga practice corrects PNS under-activity via
stimulation of the vagus nerves to ameliorate stress-related illness symptoms (109). The
124
conscious regulation of breathing practiced in yoga may influence autonomic function by
regulating the balance of sympathetic and parasympathetic activation, the latter
increasing as the breath deepens and slows (110,111). Yoga practice also appears to
result in increased vagal modulation (113) and control over autonomic responses such as
heart rate (112). During active poses, heart rate and respiration may increase consistent
with exercise (114). However, during and following the restful supine and seated
meditative postures that conclude most yoga classes, parasympathetic activity, as
measured via HRV frequency domains, consistent with relaxation and a reduction in
physiological arousal is apparent (115). In a cross-sectional study examining resting HRV
in yoga and non yoga practitioners, LF HRV appeared higher in non-yoga practitioners
while HF HRV was higher in those who practice yoga versus those who did not (237).
4.1.4 Neuro-Affective Theories and Approaches
The current research uses several supportive theories from the psychological
sciences to further frame these phenomena. These theories include: 1) the Circumplex
Model of Affect, 2) the Dual-Mode Theory of Affective Responses to Exercise, 3) the
Effort-related Attention Model, 4) the Neurovisceral Integration Model, all utilised within
the confines of an overarching 4) Neurophenomenological approach.
4.1.4.1 Circumplex Model of Affect. Given the proposed importance of positive
affect and yoga’s influence on affect, the Circumplex Model of Affect (81) offers an
important parsimonious explanatory approach. According to this model, affect is best
understood as an underlying state comprised of two orthogonal dimensions: feeling good
or bad (valence), and energized or enervated (activation-arousal). These two dimensions
of valence and activation-arousal, termed core affect, underlie all affective states. Upon
125
this affective basis are layered various cognitive processes that interpret and refine
emotions and moods and influence reflexes, perception, cognition, and behaviour (82).
Specific affective states are considered combinations of these two dimensions. The four
quadrants within this model can be described as: high activation pleasure (energy), low
activation pleasure (calmness), high activation displeasure (tension), and low activation
displeasure (tiredness) (see Figure 4.1).
4.1.4.2 Dual Mode Theory. The Circumplex approach can be further
contextualised to exercise research settings utilising Dual Mode Theory (168), which
suggests exercise experiences are regulated by affect via the continuous interplay of two
mechanisms; 1) cognitive processes, and 2) interoceptive cues. The relative importance
of these two factors in the regulation of core affect is hypothesized to shift systematically
as a function of exercise intensity, with cognitive factors being dominant at low and
moderate intensities and interoceptive cues becoming more prominent at higher
intensities (85). To date, Dual Mode Theory has been used primarily in affective
responses at various exercise intensities (73) but has not been utilised in the exploration
of lower-intensity activity such as yoga.
4.1.4.3 Effort-Related Attention Model. The Effort-Related Attention Model
suggests these attention strategies during exercise vary based on intensity (91). During
low or moderate intensity workloads attention is voluntary and can switch between
dissociative and associative attentional strategies, as external stimuli do not compete with
internal stimuli, while at higher exercise intensities increased physiological cues demand
attention. As a result attention narrows, shifts internally and becomes associative (92).
126
4.1.4.4 Neurovisceral Integration Model. The Neurovisceral Integration Model
(116) links these dimensions of valence, activation and attention via cardiac autonomic
function and specifically PNS function, as measured by heart rate and HRV. There is
abundant evidence that regular exercise training induces greater parasympathetic
modulation of cardiac function as measured via HRV in physically fit individuals (238).
Daily exercise also improves positive affect by shifting the autonomic balance to
parasympathetic predominance as measured via spectral HRV (119). Conversely,
evidence suggests sedentary lifestyles result in negative mood symptoms, especially
feelings of fatigue and depressive symptoms, which are related to decreases in
parasympathetic function as measured via LF/HF ratio (118).
4.1.4.5 Neurophenomenology. The present research is concerned with the
phenomenological lived experience of contemporary yoga practice. The phenomena are,
in part, explored via the lens of neurophenomenology, a convergent mixed methods
research approach that combines qualitative subjective first-person narratives,
quantitative self-reports and physiological data to develop a layered approach to
conscious experience (123,239,240). The strategy of neurophenomenology is to provide
mutually informative insights between first-person and neurophysiological accounts as a
means of generating new data in the exploration of consciousness (121). In addition, the
importance of contemplative mental training for research on consciousness is central to
neurophenomenology (123). This novel research approach has been used previously in
examining contemplative practices, particularly with advanced meditation practitioners
(208), using both EEG and fMRI methodologies in neuroscientific research. However,
127
this same approach has never been used in a contemporary yoga setting using HRV as a
psychophysiological correlate of affect and attention responses.
4.2 Objectives
The present study explores participants’ experience of a single yoga session in an
effort to understand how yoga practice influences participants’ immediate experience
both during and immediately post-yoga practice. Based on the Circumplex Model of
Affect (81), the Dual-Mode Theory of Affective Responses to Exercise (85), the EffortRelated Attention Model (91), the Neurovisceral Integration Model (106,116,117), and
Neurophenomenology (123,239,240), this lab-based protocol examines the relations
between valence, activation, attention, perceived exertion and cardiac activity both
during and before and after a single yoga session in cancer survivors.
Objective 1a: To examine changes in valence, activation, attention and perceived
exertion during a single yoga session as well as associated measures of cardiac activity,
including heart rate and HRV. Objective 1b: To examine whether changes in valence,
activation, attention and perceived exertion were related to one another and/or changes in
cardiac activity. Objective 1c: To examine whether changes in valence, activation,
attention, perceived exertion and cardiac activity were related to subjective first-person
descriptions of the yoga session. Objective 2a: To examine whether significant changes
in dimensions of affect, including measures or energy, tension, tiredness and calmness,
occurred pre-post yoga class. Objective 2b: To examine whether changes in dimensions
of affect were related to one another and/or to concurrent changes in valence, activation,
attention, perceived exertion, and cardiac activity. Objective 2c: To examine whether
128
changes in affect or cardiac activity were related to subjective first-person descriptions of
the yoga session.
4.3 Methods
4.3.1 Subject Recruitment
Ethical approval was obtained from the Conjoint Health Research Ethics Board
of the University of Calgary/Alberta Health Services (Ethics ID # 23313). Purposeful
sampling was used with the goal of identifying information-rich participants in order to
study each participant in-depth during a single yoga session. The sample was thus drawn
from participants who had previously completed the 7-week Yoga Thrive intervention for
cancer survivors (149,186). For the quantitative component of this study, a preliminary
power calculation using GPower 3.1 (241) using an a priori clinically meaningful effect
size of 0.50 (139) on the Feeling Scale (188) as the primary outcome with an α of .05 and
power of .80, indicated a minimum of 16 participants were required.
4.3.2 Procedures
Upon arrival in the lab, participants completed a questionnaire package assessing
baseline demographics, weekly physical activity, previous yoga experience, yoga beliefs,
and general affect. Participants then completed an individually-taught, single
standardized 80-minute yoga session. In order to assess objective 1a, using a research
protocol similar to Ekkekakis and colleagues (242-244), psychological measures
including valence, activation, attention and perceived exertion were completed eight
times during the yoga session. The assessments were completed in 10-minute intervals at
the end of each of the following sequences: 1) supine resting baseline, 2) supine yoga, 3)
seated yoga, 4) kneeling yoga, 5) standing yoga, 6) second supine yoga, 7) supine
129
meditation, and 8) supine resting post yoga session. In order to assess objective 2a, a
measure of general affect was completed at the end of all supine resting sequences: 1)
supine resting baseline, 2) supine meditation, and 3) supine resting post yoga session.
Participants had prior training in filling out all self-reported measures, having participated
in a yoga intervention study previously that used these measures.
Cardiac activity was assessed throughout the lab study via a wireless chest-band
heart rate monitor (RS800CX, Polar Electro, Kempele, Finland) to obtain continuous
ambulatory measures of heart rate and HRV as indicants of psychophysiological function
(104,106,245). Measures included heart rate (bpm) and HRV (time domains: SDNN,
RMSSD; frequency domains: HF HRV nu; non-linear: Sample Entropy) (101). Because
HRV is regulated differently throughout the day, experiments were conducted during the
morning to ensure similar cardiac activity for all participants (246). In addition,
participants were asked not to exercise or consume caffeine or alcoholic beverages 12
hours prior to the experiment and to refrain from eating and drinking 3 hours prior to the
experiment, as these agents have been shown to influence HRV (101,247). All measures
were subsequently divided into 5-minute epochs derived from the midpoint of each yoga
sequence for data analysis. Both resting and ambulatory measures of heart rate and HRV
are included in the study. While the validity of non-resting measures of HRV has been
called into question (248), given the low-intensity (peak 52% of age-predicted maximum
heart rate) of the yoga session and previous research in yoga settings using HRV as an
ambulatory measure (115), this data was examined throughout.
Following completion of all self-report scales, each participant completed a single
10-minute interview after the post-yoga supine resting condition (objective 1c and 2c).
130
The aim of the interview was to provide a structure that tapped into specific areas of
participants’ subjective experience of the yoga session (249). Participants were asked to
describe their experience of each component of the session via a series of open-ended
questions (243). All self-report scale and questionnaire data was then reviewed with
participants as well as heart rate range at the beginning and end of each sequence (HRV
data required additional processing and analysis and could not be reviewed at this time).
Participants were invited to reflect and comment on all available data. This interview
approach permitted discussion and allowed for data to enter the interview not directly
elicited, allowing patients to provide information they believed was important and
relevant to them (250). At the end of each interview, the main points were summarized
by the researcher for the individual for the purpose of member-checking and credibility
(153). All interviews were audio taped with consent.
4.3.3 Instruments – Baseline
Demographics: Demographic information included age, education, marital status
and current employment status. Medical history included cancer diagnosis, date of
diagnosis and type(s) of cancer treatment. Participant height and weight was also
measured to calculate Body Mass Index (BMI) (mass (kg) / height (m)2).
Godin Leisure Time Exercise Questionnaire . The GLTEQ (193) has been used to
assess previous physical activity levels. The LSI contains three questions that assess
frequency of mild, moderate, and strenuous physical activity performed for at least 15
minutes duration in a typical week within the past month. A weekly total of minutes of
moderate-vigorous physical activity can also be computed from this scale (baseline
Cronbach’s α: LSI = .88; moderate + vigorous physical activity = .86).
131
Beliefs about Yoga Scale. The BAYS (187) is an 11-item self-report measure
developed to examine common positive and negative beliefs about yoga in order to help
yoga instructors understand participant expectations related to yoga (baseline Cronbach’s
α = .78).
4.3.4 Instruments – In-Task Measures (objective 1a):
Feeling Scale. The FS (188) is an 11-point, single-item, bipolar measure of
affective valence (pleasure-displeasure). The scale ranges from -5 to +5. Anchors are
provided at zero (‘Neutral’) and at all odd integers, ranging from ‘Very Good’ (+5) to
‘Very Bad’ (-5). Instructions were to “estimate how good or bad you feel right now.” The
FS is commonly used for the assessment of affective responses during exercise (189)
(ICC = 0.90).
Felt Arousal Scale. The FAS (190) is a 6-point, single-item measure of perceived
activation. The scale ranges from 1 to 6, with anchors at 1 (‘Low Arousal’) and 6 (‘High
Arousal’) (ICC = 0.89). Instructions were to “estimate how aroused you feel right now
(low arousal = calm or fatigued, high arousal = anxious or energized).
Association / Dissociation Scale. The A/D S (191) is a single-item 10 point
bipolar scale designed to measure to what extent attention is primarily associative or
dissociative. The scale ranges from 1 to 10, with anchors at 1 (very associative) and 10
(very dissociative) (192) (ICC = 0.74). Instructions were to “estimate your present state
of attention right now (associative = focused in present moment, dissociative = unfocused
in past or future).
Rating of Perceived Exertion. The RPE (251) is a single item measure used to
assess perceptions of effort. The RPE is a 15-point scale ranging from 6 (Extremely
132
Light) to 20 (Extremely Hard) (ICC = 0.83). Instructions were to “rate your perception of
exertion (i.e., how strenuous the exercise felt to you / how tired you feel as a result).”
Heart Rate Variability. Given the high multicollinearity (r >.7) between various
time and spectral estimates of HRV (101), a stepwise multivariable approach (252) was
used to choose appropriate indices of HRV. Measures included in the HRV analysis were
HF HRV nu, SDNN, RMSSD, and Sample Entropy. Both SDNN and Sample Entropy are
thought to represent total complexity of the heart rate signal including both sympathetic
and parasympathetic components, while both RMSSD and HF HRV nu are thought to
reflect parasympathetic activation of the cardiac autonomic nervous system (107).
4.3.5 Instruments – Pre, Post-Supine Meditation & Post-Yoga Supine Resting (objective
2a)
Activation Deactivation Adjective Check List. The AD ACL (194) is a 20-item
measure of the four quadrants of circumplex affective space. Energy is theorized to map
the high activation pleasure quadrant of the circumplex, Tension maps the high-activation
displeasure quadrant, Tiredness maps the low-activation displeasure quadrant, and
Calmness maps the low-activation pleasure quadrant (83) (baseline α: energy = .88; tired
= .84; tension = .79; calm = .80).
4.3.6 Data Analysis
All data analyses were conducted using IBM SPSS version 19. Demographics and
medical history were described using frequency and descriptive statistics to characterize
study participants. This investigation used a repeated measures design in which
participants served as their own control. Before beginning primary analyses, all
continuous variables were checked for normality (skewness, kurtosis) and corrected
133
where necessary. Statistical analyses were conducted on the entire sample (n = 18). HRV
indices were analysed via five-minute epochs derived from the mid-point of each yoga
sequence (115).
4.3.6.1 Multilevel Modelling. In the initial analyses, multilevel models (linear
mixed models) were used to create estimated marginal means models to sequentially
examine change between: 1) baseline measures and all sequences of the yoga intervention
(objective 1a), and 2) between baseline measures, the end of the yoga session and the end
of the post yoga session supine-resting condition (objective 2a). The purpose of these two
separate analyses is to allow comparisons between the conclusions that were drawn from
pre-to-post assessments as opposed to in-task assessments (242). Multilevel models are
appropriate for analyzing data with dependent observations (such as within-subject
repeated measures) due to the hierarchical structure of the data. That is, each time
measurement of a given variable (level 1) was nested within each participant (level 2)
(198). Correlations among observations from the same individual were modeled using an
unstructured covariance matrix across all time points.
Objective 1b: In the exploratory analyses, multilevel regression analyses were
conducted to assess concurrent associations between valence, activation, attention, and
perceived exertion with indices of cardiac function. Age, BMI, time since diagnosis,
physical activity, previous yoga experience, and baseline affect were modelled as timeinvariant covariates. Time was measured continuously and included linear and quadratic
terms (200). All models were tested step by step (202,253). First, an initial unconditional
model was developed for each outcome variable (data not shown) followed by an
unconditional growth model that included only time variables (data not shown). Predictor
134
variables were then added individually to the unconditional growth model and tested for
main and interaction effects with each time term (data not shown). Significant predictors
and their time interactions, if significant, were then tested together as part of their
respective overall scale (data not shown). A final trimmed model was developed by
entering all significant predictors and their interactions to test overall associations across
time (exclusion p >.1). This method of model development has proven robust with
smaller sample sizes and ensures these models do not tax the “carrying-capacity” of the
dataset (201).
If interaction terms were not significant it could be assumed relations between
predictor and outcome variables were constant throughout the study. In each model the
time variable was centered at initial status; therefore the intercept of the regression model
can be interpreted as participant reports of the outcome variable at baseline. To enhance
interpretability of the model intercept parameters, all predictor variables were grandmean centered. Follow-up multilevel regression analyses were not computed for cardiac
activity indices but instead were used as continuous predictors in models predicting,
valence, activation, attention and perceived exertion.
Objective 2b: These same steps were followed to examine resting measures on the
ADACL at three time points 1) post-resting baseline, 2) post-supine meditation condition,
3) post-yoga session supine resting condition. Associations between the ADACL and all
other time invariant and time-varying predictors were then examined via additional
multilevel regression analyses.
4.3.6.2 Clinical Significance. Clinical significance, or meaningful clinically
important difference, a marker of program effectiveness, was assessed using multiple
135
criteria including: 1) an effect size using Cohen’s interpretation of >.20 as a small effect
>.50 as a moderate effect and >.80 as a large effect (142), 2) a standard mean difference
between baseline and follow-up time points ≥ 1 standard error of the measure (SEM),
and by comparing changes from pre to post to meaningful clinically important difference
in the research literature (135). Effect sizes were calculated using Cohen’s d for each
outcome variable, representing change from baseline to the end of each yoga sequence
and the post-yoga session resting condition (T1-Tx / T1 SE * √N). SEM was calculated
as baseline (SE * √n) * (√1 – ICC) (254). Pseudo R2 statistics were calculated as
unconditional model residual variance – trimmed model residual variance / unconditional
model residual variance. These statistics indicate the reduction in residual variance
between the unconditional and trimmed model and provide an estimate of effect size
similar to traditional OLS regression (202) and can be interpreted using Cohen’s criteria
of >.02 as a small effect >.13 as a moderate effect and >.26 as a large effect (142). All
statistical significance levels were set at p = .05 with a two-tailed test.
4.3.6.3 Qualitative Analysis (objectives 1c and 2c). All interviews were conducted
by a single researcher (MM), and coded independently by a research assistant (AW) and
the interviewer (MM). Interviews were transcribed verbatim and analysed using the
NVivo 9 computer program. Using a process of inductive thematic analysis (249), raw
data quotes were identified, labelled and organized into key themes. Common themes
were then combined into higher-order categories. Themes and categories were compared
and contrasted to assure consistency of coding and that all data was accounted for by the
core categories. Direct quotes were provided from participants to ensure the
136
transferability of categorical assertions made by the researchers and so that readers can
experience for themselves participant perspectives (153,243,255).
4.3.6.4 Mixed Methods. Finally, a convergent mixed methods
neurophenomenological approach was used to analyse and integrate self-report
questionnaire / scale, physiological and qualitative datasets. This multilayered analysis
and interpretation provided mutual insights between phenomenological first-person
accounts and physiological data. In coding with triangulation in mind, the emphasis was
placed on seeking corroboration between quantitative and qualitative data (256).
Specifically, after initial quantitative analysis and independent coding of the qualitative
data, the latter was used to further guide analysis and interpretation of the quantitative
data. This led in several cases to re-examining and remodelling data from the multilevel
regression analyses. Theoretically, integrating quantitative and qualitative methods and
applying an iterative analytic process (257) permits more accurate quantitative
descriptors of phenomena and a more parsimonious understanding of measured
constructs (258).
4.4 Results:
Results were examined in the following order: demographics: a brief overview of
participant demographics. In-task effects: estimated marginal means models examined
change in valence, activation, attention, perceived exertion and cardiac activity during the
single yoga session (objective 1a); exploratory multilevel regression analyses examined
the associations between valence, activation, attention, perceived exertion and cardiac
activity (objective 1b); qualitative descriptions of both the yoga class and participantproposed explanatory mechanisms for the effects of yoga (objective 1c). Pre-post yoga
137
session effects: estimated marginal means models examined changes in energy, tension,
tiredness and calmness before and after the single yoga session (objective 2a);
exploratory multilevel regression analyses examined the relationships between these
affective states, valence, activation, attention, perceived exertion and cardiac activity
(objective 2b); qualitative descriptions of the four affective states as experienced before
and after the single yoga session (objective 2c).
4.4.1 Demographics
Forty participants from a concurrent yoga intervention study were eligible to
participate in this research. Of those 40 participants, 18 chose to participate in the singlesession yoga study. The mean participant age was 54.0 years. All participants were
female and 61.1% of participants had received a breast cancer diagnosis, stage II-III,
approximately 36.0 months prior to study enrollment. Participants were commonly
married (55.6%), highly educated (61.1%), affluent (44%) and many participants had
returned to work fulltime (61.1%). Participants had completed the Yoga Thrive program
a mean of 4.9 times previously. The average participant BMI was 24.8, considered to be
in the high normal to overweight range. Participants reported an average of 144.9 minutes
of moderate-vigorous physical activity per week, just short of CSEP / ACSM
recommended levels of 150 minutes per week (23). In addition, participants reported high
positive beliefs concerning the outcomes of engaging in yoga practice (see Table 4.1).
4.4.2 Valence, activation, attention, perceived exertion and cardiac activity during yoga
session (objectives 1a and 1b):
4.4.2.1 Valence. A marginally significant increase (p = .100) in valence began
during the kneeling sequence, exceeding 1 SEM from baseline criteria. This increase in
138
valence became more pronounced during the standing sequence, second supine sequence
and supine meditation conditions and was maintained post yoga intervention [F(7,119) =
19.37, p < .001] (see Table 4.2, Figure 4.3). To examine what may have elicited these
changes, exploratory multilevel regression analyses were computed. Time was a
significant predictor of a linear increase in valence over the course of the intervention. In
addition, those with higher overall physical activity and energy at baseline reported
higher valence during all subsequent yoga sequences. Increased associative attention and
lower RPE were also related to higher reported valence consistently over time. Finally, a
marginally significant effect for SDNN was observed, suggesting the possibility those
with higher recorded HRV reported higher affective valence throughout the yoga session.
The Pseudo R2 value suggests inclusion of these predictor variables reduced error in
predicting valence during the yoga session by 60% (see Table 4.4).
4.4.2.2 Activation. A marginal statistically significant increase in activation was
observed that exceeded the 1 SEM during the seated yoga condition and was maintained
during the kneeling yoga condition. Activation peaked in the standing yoga condition and
a moderate effect was maintained in the second supine yoga condition, before returning
to baseline values for both the supine meditation and post-yoga supine resting conditions
[F(7, 119) = 2.12, p = .045] (see Table 4.2, Figure 4.3). In further exploring what may
have predicted change in activation, a quadratic growth pattern was evident, indicating an
initial linear increase in activation followed by a significant decline. Those with a higher
BMI at baseline reported lower activation throughout the yoga session. In addition, those
with higher dissociative attention also consistently reported higher activation. Finally, a
marginally significant effect for SDNN was observed, suggesting the possibility those
139
with lower recorded HRV consistently reported higher activation throughout the yoga
session. The Pseudo R2 value suggests inclusion of these predictor variables reduced
error in predicting activation during the yoga session by 42% (see Table 4.4).
4.4.2.3 Attention. Attention began to significantly shift during the kneeling yoga
sequence and became progressively more associative, culminating during the supine
meditation condition. This state of associative attention was maintained during the postyoga supine resting period [F(7,119) = 9.02, p < .001] (see Table 4.2, Figure 4.3). In
further examining what predicted changes in attention, a marginally significant effect for
linear time was found when time was remodelled as a linear continuous variable. In
addition, higher reported affective valence was associated with increased associative
attention at all time points, while increased activation was associated with increased
dissociative attention. A complex interaction was observed between sample entropy and
time in relation to attention. Follow-up simple slopes analyses (259) illustrate those with
low HRV had the lowest dissociative attention scores at baseline but experienced nonsignificant reductions in dissociative attention over time, while those with average HRV
experience significant decreases in dissociative attention over time. Those with high
HRV had the highest dissociative attention scores at baseline but experienced the greatest
decrease in dissociative attention (see Figure 4.2). The Pseudo R2 value suggests
inclusion of these predictor variables reduced error in predicting attention during the yoga
session by 57% (see Table 4.4).
4.4.2.4 Rating of Perceived Exertion. RPE significantly increased from baseline to
the standing yoga sequence, then dropping in the second supine yoga condition before
returning to baseline values in the supine meditation and post yoga session supine resting
140
conditions [F(7,119) = 21.30, p <.001] (see Table 4.2, Figure 4.3). In multilevel
regression analyses time exhibited a quadratic growth curve: an initial linear increase was
observed followed by a significant decrease in growth. Higher reported physical activity
and energy at baseline consistently predicted increased RPE. Those who reported lower
valence also reported increased RPE and those that reported higher activation also
reported higher RPE. Finally, higher heart rate was associated with higher RPE. The
Pseudo R2 value suggests inclusion of these predictor variables reduced error in
predicting perceived exertion during the yoga session by 60% (see Table 4.4).
4.4.2.5 Cardiac Activity. Estimated marginal means models suggest heart rate was
responsive to each yoga sequence [F(7,119) = 74.84, p< .001], with significant increases
in heart rate from baseline in the seated, kneeling, and standing conditions, followed by a
significant decrease during the supine meditation that was maintained during the postyoga session supine resting condition. HF HRV nu exhibited a consistent pattern of vagal
withdrawal [F(7,119) = 9.02, p < .001], showing a consistent decline up until the standing
condition, followed by an increase during the second supine yoga condition and then a
return to baseline values during supine meditation and the post- yoga session supine
resting condition. SDNN exhibited initial increases during the first supine and seated
conditions before returning to baseline values during the kneeling, standing and second
supine conditions. An additional increase in SDNN was observed during supine
meditation before a return to baseline values during the post yoga supine resting
condition [F(7,119) = 4.16, p < .001]. RMSSD remained stable from baseline throughout
the first supine, seated and kneeling conditions. Significant vagal withdrawal was
observed during the standing yoga sequence that returned to baseline values during the
141
second supine sequence, followed by a subsequent increase in vagal activity during
supine meditation prior to RMSSD returning to baseline values during post-yoga supine
resting condition [F(7,119) = 4.02, p = .001]. Finally, sample entropy showed a pattern of
lessening complexity beginning in the initial supine sequence, culminating in the
kneeling sequence. Sample entropy remained reduced in both the standing and second
supine conditions before returning to approximate baseline values in the supine
meditation and post-yoga resting conditions [F(7,119) = 17.27, p < .001] (see Table 4.3,
Figure 4.4).
4.4.3 In-Task Yoga Session Qualitative Findings (objective 1c):
Several themes emerged in analysing the qualitative interview data that supported
findings from quantitative data analyses. Themes were grouped into two overarching
categories: 1) description of yoga session; and 2) explanatory mechanisms including, a)
attention and b) breath awareness.
4.4.3.1 Description of Events . This category consisted of participant descriptions
of the yoga sequence.
Baseline: The majority of participants drew comparisons between the baseline
resting condition and all subsequent sequences. Many participants characterised baseline
as a time where dissociative attention was predominant, activation was alternately
characterised as either low or high, and valence was low. In examining the high degree of
dissociative attention reported at baseline, many participants suggested they lacked focus
at the beginning: “I found my head was doing all sorts of busy brain stuff.” and, “I was
feeling agitated – my thoughts were all over the place.” Contrary to the quantitative
142
findings and the majority of participant comments, others found the resting baseline very
relaxing: “I felt good and relaxed – looking forward to what we were going to do.”
Supine I, Seated, Kneeling, and Standing Yoga Sequences: The first supine
sequence was characterised by many participants as a preparatory phase of rest and
relaxation. Contrary to the quantitative findings, several participants reported noticing an
immediate shift in attention during the first supine yoga sequence: “The first sequence
got my mind focused. Focusing on breathing really clears your mind.” Other participants
reported, in concordance with the quantitative findings, that shifts in valence, activation
and attention occurred later in the yoga session: “It takes a bit for me to get there. Then
suddenly I am there in the moment. It was about the 2nd (seated) or 3rd (kneeling)
sequence when I felt I’d really gotten there.”
Participants remarked that increased exertion (RPE) and physical sensation as the
yoga sequences became physically more active facilitated attention: “The lunge part
(kneeling sequence) is where I actually felt I was doing more physical yoga for the first
time in the series. I like doing physical yoga so that was a good feeling for me.” and, “I
was feeling a bit more focused because my body was doing something, so the mind was
not able to go grocery shopping.” The standing sequence was described as more
physically demanding than the previous yoga sequences, which is corroborated by the
quantitative data. “This was just more physically demanding, more muscles working”
and, “I definitely felt like that was the most energetic of the sequences.” Another
participant commented the standing sequence elicited, “More wakefulness and more
energy.... it felt like it was all coming together better.” Several participants suggested
during the active poses that heart rate was a good indicator of their level of exertion:
143
“Your heart tells you what you are doing.” These supine, seated, kneeling and standing
yoga sequences reflect changes in valence and associative attention, including awareness
of heart rate, activation and rating of perceived exertion.
Supine Yoga Sequence II and Supine Meditation: In the second supine sequence,
quantitative analyses showed that self-reported activation and perceived exertion
remained high, valence increased somewhat and attention was similar to the standing
sequence. Despite participants continuing to report high activation and RPE, their
subjective descriptions characterised the second supine sequence as more relaxing: “I can
feel that deep rest and let go. I can feel my hips and how things are relaxing the whole
way down.” During the supine meditation sequence valence and associative attention
peaked and there were moderate increases in RMSSD and SDNN and moderate decreases
in heart rate. One participant described this sequence as follows: “I was really in the yoga
moment... I was really experiencing my breath and the sensation through my whole
body.” Breath awareness was a strong point of associative attention in the supine
meditation sequence: “Breath awareness allows you to realise your breath goes through
your body... It was great to realise how much your breath affects every part of you.”
Despite generally positive self-reports of increased valence and associative
attention and accompanying shifts in decreased heart rate and increased SDNN and
RMSSD, not all participants had a pleasant experience in the supine meditation sequence.
One participant reported, “Not so good. I completely zoned out.” When asked why this
was the case the participant replied: “I think it’s just because I am very dissociative
normally, it’s my default. I have a hard time focusing and am very easily distracted.” For
others, the supine meditation was associated with sleep states: “It was nice and relaxing.
144
I probably could have fallen asleep. Nothing was really going through my mind.” Several
participants indicated the concluding supine meditation left them feeling more wakeful.
“When I came in I was drowsy and sleepy. Savasana (supine meditation) is just like
having a power nap... at the end of it I was feeling much more energized.” Concurrent
increases in energy and calmness and reductions in tension during the supine meditation
sequence seem to corroborate these descriptions.
Post Yoga Session Supine Resting: Several participants commented on the return
of dissociative attentional patterns immediately following the supine meditation. These
descriptions reflect shifts in attention as well as decreases in SDNN and RMSSD: “My
mind started to wake up and think again because you don’t have anything to focus on.”
and, “I still felt very relaxed, but there is a difference between just lying around and
savasana (supine meditation).” Others characterised this time as a continuation of the
previous supine meditation condition: “I was still pretty relaxed from the carry-over from
savasana (supine meditation).” Differences highlight these experiences are not uniform
and give weight to multilevel regression models that help to tease apart associations
between self-reported outcomes and other moderating variables.
Comparisons between Pre and Post Yoga Session: When participants were asked
to reflect on what they felt changed from the initial supine resting condition until the final
post yoga supine resting condition, the following insights were offered in relation to
focus of attention: “When I came in my mind was jumping everywhere. At the end I was
very focused... whereas at the start there was just no focus.” Participants reported these
shifts regardless of baseline valence: “At the end you just feel better than when you
started, no matter where you started. Even if you start feeling really poor, you feel better
145
even if it’s not great at the end.” and, “I feel much better than when I came in even
though I felt good when I came in. I thought I was relaxed, but I am a bit more relaxed
now.” At the end of the session participants still reported higher valence and associative
attention. Physiologically post yoga session participants had lower resting heart rate than
baseline, while measures of HRV had returned to baseline values.
4.4.3.2 Explanatory Mechanisms. Many participants offered their own
explanations for mechanisms that may have played a role in their experience of the single
yoga session. Two common reported mechanisms were 1) attention, and 2) breathing.
Attention: Participant references to shifts in attention were common. Several
participants described yoga as a different state of relaxed consciousness: “It focuses you
and totally gets rid of all the noise: the stuff going on in your brain.” and, “I am very
focused. I am right here, right now.” As participants continued their practice, they
noticed a shift in attention concomitant with positive valence: “I was very focused on
what we were doing and really enjoy how it feels... I didn’t find the busy brain was taking
over.” Despite a linear decrease in dissociative attention over the course of the
intervention, many participants reported this state was not uniform: “I was really relaxed
but also had some monkey brain moments… I would pay attention for a few breaths and
then my mind would go off… Before you know it I am busy running errands.” Taken
together, these participant statements suggest a strong emphasis is placed on the role of
attention in the experience of a single yoga session. Supporting the quantitative data,
focus of attention was implicated in changes in both valence and activation in the
multilevel regression analyses, with those indicating greater associative attention
concurrently reporting greater valence and lower activation.
146
Breathing: Many participants referenced the breathing aspect of yoga practice and
suggested a functional link between the breath and associative attention: “The more you
have to focus on your breathing, the more it clears all the other stuff out of your mind...
You are so focused on breathing, quieting, being calm… (Yoga) really quiets your mind.”
Attention to breathing and the physical postures was reported to help keep participants
focused and present: “When I finally start concentrating on my breath I realise I am right
here, right now, in this moment.” and, “When you are just concentrating on the breathing
and on doing the poses, everything else just disappears... It is so nice to just be in the
moment.” Several participants shared changes in valence, activation and attention prepost yoga were linked to breath regulation: “It was much easier for me to bring my
breathing back down. I got more relaxed and more feeling, so better focus.” While the
present research study did not measure breathing specifically, examining the changes in
cardiac activity associated with valence, activation, attention and perceived exertion in
the regression analyses suggest this connection, given these autonomic changes can be
associated with breath awareness (260).
4.4.4 Pre-Post Yoga Session Affect Regulation (objectives 2a & 2b)
4.4.4.1 Energy. Energy increased from baseline values to the end of the yoga
session and was maintained during the post yoga supine resting condition [F(2,34) =
6.91, p = .003] (see Table 4.5, Figure 4.5). When follow-up multilevel regression models
were built, the effect of valence eclipsed the previously significant linear effect of time
(p=.998; unconditional growth model not shown), due to a high degree of
multicolinearity between the two variables (r >.7). Those who reported higher valence
consistently reported higher energy levels. Higher energy was also related to decreased
147
tiredness. Those with higher energy also exhibited higher HF HRV nu and SDNN and
lower RMSSD. The Pseudo R2 value suggests inclusion of these predictor variables
reduced error in predicting energy pre-post yoga session by 86% (see Table 4.6).
4.4.4.2 Tension. Tension decreased from baseline values to the end of the yoga
session and was maintained during the post yoga supine resting condition [F(2,34) =
3.67, p = .036] (see Table 4.5, Figure 4.5). In examining predictors of change in tension,
those with lower resting HF HRV nu reported higher levels of tension. The Pseudo R2
value suggests inclusion of these predictor variables reduced error in predicting tension
pre-post yoga session by 72% (see Table 4.6).
4.4.4.3 Tiredness. A marginally significant decline in Tiredness was observed prepost intervention that did not exceed the SEM. However this difference became
significant during the post yoga supine resting condition [F(2,34) = 3.35, p = .047] (Table
4.5, Figure 4.5). In examining predictors of change, increased time since diagnosis
predicted increased tiredness. Increased activation and increased heart rate were also
associated with decreased tiredness. The Pseudo R2 value suggests inclusion of these
predictor variables reduced error in predicting tiredness pre-post yoga session by 63%
(see Table 4.6).
4.4.4.4 Calmness. Calmness increased from baseline to the end of the yoga
program. An additional increase in calmness was observed during the post yoga supine
resting condition [F(2,34) = 9.34, p = .001] (see Table 4.5, Figure 4.5). Increased tension
was associated with decreased calmness while higher resting heart rate was associated
with increased calmness. The Pseudo R2 value suggests inclusion of these predictor
148
variables reduced error in predicting calmness pre-post yoga session by 49% (see Table
4.6).
4.4.5 Pre-Post Yoga Session Qualitative Findings (objective 2c)
In examining participants’ interview data, categories emerged reflecting the four
quadrants of the circumplex model: energy, tension, tiredness and calmness.
4.4.5.1 Energy. In agreement with regression analyses indicating a relationship
between increased energy, valence and decreased tiredness, participants generally
reported feeling they had more energy at the end of the yoga session: “I felt really
relaxed and actually felt more energized than when I came in. I was ready to go off and
conquer the world.” and, “I feel way more energized. I feel like I had a great nap
(laughs)... If I practice yoga I have more energy, not less.” These findings suggest
increased feelings of energy are related to in–task feelings of rest and relaxation.
4.4.5.2 Tension. Generally participants reported feelings of tension when they
first arrived: “At the beginning I was probably more tense and not for anything other
than I hadn’t done the yoga yet.” Post yoga session, they reported tension to have
decreased: “It makes me relax more so I don’t feel as tense, which is good because you
should be relaxing more if you are doing these things (yoga).” States of tension were
associated with a pre-yogic state while engaging in the yoga session was associated with
reduced tension and heightened relaxation.
4.4.5.3 Tiredness. Many people reported decreased tiredness at the end of the
yoga session: “I was tired at the start. My mind was just thinking a bunch of different
things: totally unfocused.” Participants also spoke about tiredness in relationship to
energy levels: “I was very tired coming in. I feel a lot more energetic now and rested.”
149
and, “You have that battery charge... You are rested but not tired.” The multilevel
regression analyses suggest participants reported higher arousal levels and had higher
recorded heart rate in relation to lower tiredness. While participants suggest a relationship
between tiredness and dissociative attention, this relationship was not borne out in the
multilevel regression analyses.
4.4.5.4 Calmness. Discussion of feelings of calmness was a prevalent theme. “I
am calm. I am ready to go now. I feel in a unit. I feel good. I’d like to give you some
professional (sounding) words but I can’t...” In describing pre-post yoga session
increases in calmness, one participant related this increased calmness to decreases in
tension and this connection was also evident in the regression analyses: “I think over the
course of the practice, the calmness level really increased... You can almost feel yourself
exhaling, letting the tension go.” Many participants noted that although they were
calmer, they felt more energy. “I’m calmer, but more energized, as opposed to being
tired and lethargic.” However, the connection between calmness and energy was not
evident in the regression analyses. Several participants also described calm in conjunction
with focus of attention: “It really focuses you and calms you. It stops your mind from
going crazy all over the place.” and, “Maybe it’s just the experience of yoga that makes
me feel calmer. As I feel calmer I tend to be, I guess, in the present. It’s when I’m not
calm that I’m in the future.” Despite these subjective first-person descriptions, attention
was not associated with calmness in the multilevel regression analyses.
4.5 Discussion
This is the first study which has integrated the Circumplex Model of Affect, DualMode Theory, Effort-Related Attention Model, Neurovisceral Integration Model,
150
Neurophenomenology and accompanying methodologies (patient reported outcomes,
physiological indices, and first-person phenomenological reports) in the exploration of
the subjective experience of a single yoga session in a group of cancer survivors.
4.5.1 In Task Effects (objectives 1a and 1b):
In reviewing results in relation to the circumplex model of affect, the majority of
change in the single yoga session occurred in the low-activation pleasure quadrant,
touching into the high-activation pleasure quadrant in yoga sequences of higher intensity.
Specifically, participants report increasingly improved valence throughout the yoga
session while reported activation and perceived exertion rose and fell largely with the
intensity of each yoga sequence (see Figure 4.6). The addition of a dimensional measure
of associative attention leads to a 3-dimensional spherical model of eight quadrants. The
majority of change in the single yoga session occurs in the low-activation / pleasure /
associative attention quadrant, touching into the high-activation / pleasure / associative
attention quadrant in more strenuous yoga sequences (see Figure 4.7). Specifically, both
valence and attention improve over the course of the yoga session and activation rises and
falls with the difficulty of the yoga sequence.
4.5.1.1 Valence. The linear trend in valence is in concordance with previous
research utilising Dual-Mode Theory examining steady state exercise above or below the
lactate or ventilatory threshold (73), which suggests pleasant changes in valence in most
individuals at sub-threshold intensities. Given participants in the current experiment
worked at a maximum of 52% of their age-predicted maximal heart rate (208 - 0.7 * age)
(261) during the standing yoga sequence, the overall yoga session can be considered lowintensity exercise and changes in valence would be indicative of these sub-threshold
151
states. The correspondence between SDNN and valence corroborates emerging research
utilising the Neurovisceral Integration Model linking HRV and affective responses
(104,262,263).
4.5.1.2 Activation. Activation was related to an overall quadratic time effect,
evidence of the measure’s responsiveness to differing yoga sequences based on intensity.
A negative relationship between body mass and activation was also observed, with those
with higher BMIs reporting lower activation during each yoga sequence. These findings
are contrary to work by Ekkekakis and colleagues (264), whose findings did not indicate
a negative relationship between BMI and activation but rather BMI and valence. Those
who reported higher activation reported higher dissociative attention and reduced total
HRV. Participant-reported activation also reflected these sub-threshold intensities, with
only moderate activation expressed by participants throughout more active sequences
(i.e., standing yoga). Participant ratings of perceived exertion also reflect these changes,
with a maximum reported score of approximately 11 of 20 in the standing yoga sequence,
indicative of a “light” and “somewhat hard” exertion. The association between HRV and
the activation dimension of the circumplex model of affect is consistent with previous
research (265).
4.5.1.3 Attention. Attention shifted linearly from a dissociative to associative
state. This shift in associative attention does not seem to be as intensity-based as other
forms of exercise (89). In the case of concurrently monitored HRV (sample entropy),
those with higher recorded sample entropy reported greater increases in associative
attention than those with lower recorded sample entropy, for whom there were no
significant changes in focus of attention. These findings suggest potential HRV
152
modulation of attention (266,267). We concur with a recent review by Salmon and
colleague’s (95) that associative attentional states may be somewhat analogous to
mindful awareness and, in the case of yoga, may heighten kinesthetic awareness of yoga
practice. However, this highly-valenced, pleasant associative attentional state may be
more closely aligned with what Lutz and colleagues have termed “focused attention”
meditation states (208).
4.5.2 Pre-Post Yoga Session (objectives 2a and 2b):
Pre-post affective states captured during supine resting conditions (baseline
resting, supine meditation, post yoga sessions resting) by the ADACL suggest significant
improvements in energy and calmness, and corresponding reductions in tension and
tiredness. These findings are consistent with increases in energy and decreases in
tiredness and tension seen pre-post other acute exercise sessions, regardless of exercise
intensity (86). However, we also see increases in calmness not apparent in other research
(242), perhaps indicating a unique feature of yoga practice. Moreover, participants
corroborated that changes in affect pre-post class were related to yoga practice. This
participant confirmation adds credibility to the research (257).
4.5.3 Integration (all objectives)
Building on experimental frameworks previously developed by Telles and
colleagues (113,115,260) and Ekkekakis and colleagues (73), this research suggests
measures of affective valence and associative attention increase linearly throughout a
single yoga session, while measure of activation and perceived exertion are contingent on
the intensity of each yoga sequence. Specifically, participants denote higher activation
153
and exertion in more vigorous sequences and lower activation-exertion in more gentle
sequences.
Recorded cardiac indices suggest a similar pattern, with participants reaching subbaseline resting heart rates in the final supine meditation and post-yoga resting
sequences. The relationship of HRV to these conditions is more complex, with time
domains showing a pattern of vagal withdrawal and lowered HRV in the active yoga
sequences but increased HRV in the supine meditation sequence. These findings are
consistent with previous research (115) examining HRV in a single yoga session. These
increases in vagal activity may also be in part due to the inherent benefits of supine
posture in post-exercise parasympathetic reactivation (268). However, frequency and
non-linear measures show a consistent pattern of vagal withdrawal relative to the
intensity of the respective yoga sequence.
Participants’ first-person phenomenological accounts largely corroborate these
growth models and provide interesting insight into the multilevel regression analyses as
to differences in the experience of the yoga session. While participants’ collective
subjective descriptions say little about the physiological measures per se, their descriptive
emphasis on the attentional aspect of yoga and awareness of breath specifically is
illuminating given the relationship between HRV, respiration and attention
(112,260,269). Given these associations between all three levels of evidence, additional
research devoted to respiratory influences on core affect, attention and autonomic
function are of great interest.
The current research was designed to model neurophenomenological processes
and outcomes, resulting in a detailed exploration of the lived experience of yoga in the
154
context of valence, activation, attention, perceived exertion, cardiac activity and
phenomenological first-person narratives. The experience of yoga had not been studied
using an integrated theoretical and methodological approach to date. The intention was to
bridge the “explanatory gap” (121) between the phenomenal character of experience with
the quantitative measurement of that experience and to further contextualise this research
via subjective description. From this theoretical and methodological integration a more
coherent picture of the experience of yoga emerged.
4.5.4 Limitations
Many participants remarked that compared to their usual yoga classes and
routines the single session was lighter intensity. This discrepancy between the acute
single yoga session and participants regular yoga practice must be kept in mind. In
addition, given the primacy participants placed on awareness of breathing in the yoga
session, consideration of the use of an ambulatory measure of breaths per minute is
warranted. While several exercise studies support the validity of using ambulatory heart
rate monitors to assess HRV (270,271) there has been some controversy in their use with
women over age 60 (272,273). Given the age and health status of participants in the
current study these limitations must be kept in mind. Another area of concern is the use
of multilevel modelling on such a small sample size. While the study was powered on
detecting a 0.50 difference on the feeling scale it was not powered for the multilevel
regression analyses. While research suggests samples of 50 or more are preferable,
numbers of as small as 50 observations may be sufficient and will show little bias in the
regression coefficients (274).
155
4.5.5 Conclusion
Identifying potential mechanisms that explain how yoga leads to clinically
significant outcomes is an important next step as yoga programs are increasingly used in
oncology settings. Recent recommendations from the UK Medical Research Council (63)
suggest time spent delineating the mechanisms behind complex interventions, both
theoretically and practically, helps to strengthen the causal chain of evidence and further
refine both interventions and research design. Taking the opportunity to explore
participants’ lived experience of yoga provides useful information that strongly suggests
future directions for research. The richness of this data provides a source from which
questions and theories can emerge concerning the self-perceived experiences of those
who practice yoga and have had a cancer experience.
156
Table 4.1 Demographics
Baseline (n = 18)
Age, years (SD)
n (%)
53.97 (8.49)
Gender – Female
Time since diagnosis, months (SD)
Times previously completed Yoga Thrive
18 (100)
35.95 (19.00)
4.94 (3.00)
Cancer Diagnosis – Breast
Cancer Stage
II
III
Marital Status – Married / Common-law
Education Level – completed university / college
Annual Household Income – > $80,000
Employment Status – Fulltime
11 (61.1)
Body Mass Index (BMI)
LSI
GLTEQ, minutes per week
24.80 (4.49)
31.17 (17.40)
144.89 (112.51)
Beliefs about Yoga Scale (BAYS)
54.59 (8.65)
6 (33.3)
5 (27.8)
10 (55.6)
11 (61.1)
8 (44)
11 (61.1)
Valence
Activation
Attention
Exertion
(Feeling Scale;
(Felt Arousal Scale;
(Assoc. – Dissoc. Scale;
(Borg RPE Scale;
ICC = 0.90, SEM = 0.36) ICC = 0.89, SEM = 0.45) ICC = 0.74, SEM = 1.10) ICC = 0.83, SEM = 0.82)
Mean
Mean
Mean
Mean
P
d
p
d
p
d
p
d
(SE)
(SE)
(SE)
(SE)
2.06
2.78
5.06
6.33
Pre- Resting
(0.27)
(0.32)
(0.51)
(0.47)
2.28
2.83
4.56
7.89
Supine 1
.409
0.20
.864
0.04
.147
-0.23
.005
0.79
(0.27)
(0.32)
(0.51)
(0.47)
2.22
3.28
4.50
9.28
Seated
.536
0.15
.124
0.37
.107
-0.26
.000
1.49
(0.27)
(0.32)
(0.51)
(0.47)
2.50
3.22
4.06
10.44
Kneeling
.100
0.39
.171
0.32
.004
-0.46
.000
2.08
(0.27)
(0.32)
(0.51)
(0.47)
2.78
3.56
3.83
10.78
Standing
.008
0.64
.018
0.57
.001
-0.56
.000
2.25
(0.27)
(0.32)
(0.51)
(0.47)
3.50
3.61
3.78
9.28
Supine 2
.000
1.28
.011
0.61
.000
-0.59
.000
1.49
(0.27)
(0.32)
(0.51)
(0.47)
Supine
4.11
2.83
2.72
7.11
.000
1.83
.864
0.04
.000
-1.08
.151
0.39
Meditation
(0.27)
(0.32)
(0.51)
(0.47)
4.06
2.94
3.44
6.39
Post-Resting
.000
1.78
.607
0.12
.000
-0.75
.918
0.03
(0.27)
(0.32)
(0.51)
(0.47)
ICC = Intraclass correlation coefficient, SEM = Standard Error of Measurement; Cohen’s = Tx – T1 / (T1SE*(√N))
Table 4.2 Estimated Marginal Means Models – Valence, Activation, Attention and Rate of Perceived Exertion.
157
157
Heart Rate
(ICC = 0.97,
SEM = 1.68)
Pre- Resting
Supine 1
Seated
Kneeling
Standing
Supine 2
Supine
Meditation
Post-Resting
Mean
(SE)
67.94
(2.29)
69.48
(2,29)
75.07
(2.29)
81.52
(2,29)
88.19
(2,29)
70.04
(2.29)
62.08
(2.29)
62.53
(2.29)
HF HRVnu
(ICC = 0.83,
SEM = 7.48)
p
D
-
-
.302
0.16
.000
0.73
.000
1.40
.000
2.09
.159
0.22
.000
-0.60
.000
-0.56
Mean
(SE)
41.41
(4.28)
22.90
(4.28)
18.77
(4.28)
21.72
(4.28)
22.51
(4.28)
27.84
(4,28)
35.59
(4,28)
41.76
(4.28)
SDNN
(ICC = 0.91,
SEM = 7.91)
p
d
-
-
.000
-1.02
.000
-1.25
.000
-1.09
.000
-1.04
.002
-0.75
.185
-0.32
.935
0.02
Mean
(SE)
35.96
(6.21)
46.56
(6.21)
44.56
(6.21)
40.18
(6.21)
29.88
(6.21)
41.28
(6.21)
49.03
(6.21)
38.84
(6.21)
Sample Entropy
(ICC = 0.82,
SEM = 0.14)
RMSSD
(ICC = 0.90,
SEM = 8.82)
p
d
-
-
.014
0.40
.045
0.33
.321
0.16
.155
-0.23
.212
0.20
.003
0.50
.497
0.11
Mean
(SE)
35.87
(6.57)
38.08
(6.57)
34.59
(6.57)
32.34
(6.57)
22.34
(6.57)
35.57
(6.57)
47.26
(6.57)
40.15
(6.57)
p
d
-
-
.658
0.08
.798
-0.05
.481
-0.13
.008
-0.49
.952
-0.01
.024
0.41
.393
0.15
Mean
(SE)
1.43
(0.08)
1.20
(0.08)
1.04
(0.08)
0.66
(0.08)
0.88
(0.08)
1.13
(0.08)
1.36
(0.08)
1.30
(0.08)
p
d
-
-
.010
-0.68
.000
-1.15
.000
-2.29
.000
-1.62
.001
-0.89
.429
-0.21
.127
-0.40
ICC = Intraclass correlation coefficient, SEM = Standard Error of Measurement; SE = Standard Error; P = significance, d = Cohen’s d
- Tx – T1 / (T1SE*(√N)); HF HRV nu = High Frequency Heart Rate Variability (normalized units); SDNN = standard deviation of all
normal-normal r-r intervals; RMSSD = square root of the mean of squares of successive r-r interval differences.
Table 4.3 Estimated Marginal Means Models – Heart Rate, HF HRVnu, SDNN, RMSSD, Sample Entropy.
158
158
Predictor
Intercept
Linear Time
Time2
Valence
(Feeling Scale)
B
(SE)
1.95
(0.21)
Activation
(Felt Arousal Scale)
df
T
p
15.96
9.36
.000
0.28
(0.04)
21.74
7.97
.000
-
-
-
-
Body Mass
Index
B
(SE)
2.42
(0.24)
Attention
(Association-Dissociation
Scale)
B
df
t
p
(SE)
df
t
p
25.70
10.21
.000
4.49
(0.50)
19.15
8.96
.000
.000
-0.14
(0.07)
28.70
-1.97
.059
.000
-
-
-
-
0.48
108.07 4.09
(0.12)
-0.06
106.36 -3.98
(0.01)
Perceived Exertion
(Borg RPE Scale)
B
(SE)
6.05
(0.49)
1.96
(0.25)
-0.26
(0.04)
df
t
p
28.39
12.29
.000
133.99
7.71
.000
124.45
-7.13
.000
-0.08
(0.04)
15.90
-2.06
.056
-
-
-
-
-
-
-
-
Physical
Activity
(LSI)
0.03
(0.01)
17.02
2.76
.013
-
-
-
-
-
-
-
-
0.05
(0.01)
13.32
3.35
.005
Baseline
Energy
0.11
(0.04)
17.46
3.12
.006
-
-
-
-
-
-
-
-
0.16
(0.05)
12.48
3.02
.010
Valence
-
-
-
-
-
-
-
-
-0.27
118.33 -2.89
(0.09)
.005
-0.56
(0.16)
122.72
-3.60
.000
Activation
-
-
-
-
-
-
-
-
0.29
128.90
(0.09)
3.36
.001
0.32
(0.11)
42.78
2.84
.007
Attention
-0.13
(0.05)
65.38
-2.55
.013
0.22
(0.06)
77.46
3.65
.000
-
-
-
-
-
-
-
-
RPE
-0.12
120.02 -4.48
(0.03)
.000
-
-
-
-
-
-
-
-
-
-
-
-
Table 4.4 Multilevel Regression Analyses – Valence, Activation, Attention and Rate of Perceived Exertion
159
159
Valence
(Feeling Scale)
Predictor
Heart Rate
SDNN
Attention
(Association-Dissociation
Scale)
Activation
(Felt Arousal Scale)
Perceived Exertion
(Borg RPE Scale)
B
(SE)
df
T
p
B
(SE)
df
t
p
B
(SE)
df
t
p
B
(SE)
df
t
p
-
-
-
-
-
-
-
-
-
-
-
-
0.04
(0.02)
73.55
2.27
.026
1.76
.082
.065
-
-
-
-
-
-
-
-
0.008
124.11
(0.004)
-0.008
103.28 -1.86
(0.004)
Sample
Entropy
-
-
-
-
-
-
-
-
1.05
122.49
(0.45)
2.30
.022
-
-
-
-
Sample
Entropy
*Time
-
-
-
-
-
-
-
-
-0.30
118.96 -2.75
(0.11)
.007
-
-
-
-
Pseudo R2
Valence = 0.60
Activation = 0.42
Attention =0.57
Perceived Exertion = 0.60
Excluded (p > .1): age, time since diagnosis, previous yoga thrive experience, yoga beliefs (BAYS), baseline tension, tiredness
calmness (ADACL), HF HRV nu, RMSSD. B = estimate; SE = standard error; df = degrees of freedom; p = significance.
Table 4.4 – continued
160
160
Predictor
Energy
(ICC = 0.77, SEM = 1.98)
Tension
Tired
(ICC = 0.50, SEM = 1.28) (ICC = 0.83, SEM = 1.59)
Calm
(ICC = 0.55, SEM = 1.43)
Mean
(SE)
P
d
Mean
(SE)
p
d
Mean
(SE)
p
d
Mean
(SE)
p
d
Pre- Resting
10.17
(0.97)
-
-
6.78
(0.43)
-
-
10.39
(0.91)
-
-
17.00
(0.50)
-
-
Savasana
13.39
(0.97)
.002
0.78
5.44
(0.43)
.020
-0.73
9.00
(0.91)
.089
-0.36
18.50
(0.50)
.004
0.71
Post-Resting
13.00
(0.97)
.005
0.69
5.56
(0.43)
.032
-0.67
8.39
(0.91)
.016
-0.52
19.00
(0.50)
.000
0.94
ICC = intraclass correlation coefficient; SEM = Standard Error of Measurement; SE = Standard Error; p = significance; d = Cohen’s d
- Tx-T1 / (T1*(√N)).
Table 4.5 Estimated Marginal Means Model – AD ACL.
161
161
Predictor
Intercept
Linear Time
Time Since
Diagnosis
Tension
Energy
B
(SE)
Df
12.18
25.99
(0.92)
0.0003
38.51
(0.18)
T
p
13.29
.000
0.002
.998
Valence
6.78
(0.65)
-0.20
(0.09)
df
t
p
16.97
10.44
.000
17.02
-2.15
.046
Tired
B
(SE)
df
t
10.87
16.40 13.08
(0.83)
-0.37
21.76 -2.79
(0.13)
0.14
15.66 2.68
(0.05)
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
.000
-
-
-
-
-
.045
-
-
-
-
-
-0.61
403.72 -5.54
(0.11)
0.92
44.49 2.06
(0.45)
Tired
Tension
B
(SE)
p
.000
.011
Calm
B
(SE)
16.86
(0.56)
0.30
(0.08)
df
t
p
18.08
29.90
.000
24.73
3.97
.001
.017
-
-
-
-
-
-
-0.28
(0.14)
31.40
-2.03
.051
-
-
-
-
-
-
-
-
-
-
-
-
-
-
.015
-
-
-
-
.035
0.10
(0.04)
29.02
2.614
.014
-0.70
37.24 -2.56
(0.27)
-0.14
29.01 -2.21
(0.06)
Activation
-
-
-
-
-
-
-
-
Heart Rate
-
-
-
-
-
-
-
-
2.56
.017
-0.02
(.009)
35.62
-1.90
.066
-
-
-
-
-
-
-
-
2.34
.026
-
-
-
-
-
-
-
-
-
-
-
-
-2.60
.014
-
-
-
-
-
-
-
-
-
-
-
-
HF HRV
SDNN
RMSSD
Pseudo R
2
0.06
25.25
(0.02)
0.15
29.78
(0.07)
-0.14
33.43
(0.06)
Energy = 0.86
Tension = 0.72
Tired = 0.63
Calm = 0.49
Excluded (p > .1): age, previous yoga thrive experience, baseline physical activity (GLTEQ), yoga beliefs (BAYS). BMI, energy and
calm (ADACL); attention (ADS); perceived exertion (RPE); sample entropy. B = estimate; SE = Standard Error; df = degrees of
freedom, p = significance.
Table 4.6 Multilevel Regression Analyses – AD ACL.
162
162
Figure 4.1 Circumplex Model of Affect.
163
163
164
Figure 4.2 Simple slopes of the regression for dissociative attention on condition at low, mean, and high levels of sample entropy.
Note: low, medium, and high values of sample entropy are defined as plus and minus 1 sd above the mean (-0.40, 0, 0.40). Slopes of
the simple regression for high and mean sample entropy significantly differ from zero, but the simple slope at low sample entropy
does not.
164
Figure 4.3 Estimated Marginal Means Models for Valence, Activation, Attention and Perceived Exertion.
165
165
Figure 4.3 Estimated Marginal Means Models for Valence, Activation, Attention and Perceived Exertion – continued.
166
166
Figure 4.4 Estimated Marginal Means Models for Heart Rate, HF HRV nu, SDNN, RMSSD & Sample Entropy.
167
167
Figure 4.4 Estimated Marginal Means Models for Heart Rate, HF HRV nu, SDNN, RMSSD & Sample Entropy – continued.
168
168
Figure 4.4 Estimated Marginal Means Models for Heart Rate, HF HRV nu, SDNN, RMSSD & Sample Entropy – continued.
169
169
Figure 4.5 Estimated Marginal Means Models for Activation-Deactivation Adjective Checklist.
170
170
Figure 4.5 Estimated Marginal Means Models for Activation-Deactivation Adjective Checklist – continued.
171
171
Figure 4.6 Affective responses (Mean FS and FAS) to a single yoga session plotted on a two-dimensional circumplex model. Labels
indicate time points for affective responses: Baseline, Supine 1, Seated, Kneeling, Standing, Supine 2, Supine Meditation, Post.
172
172
Figure 4.7 Affective responses (Mean FS, FAS and the addition of ADTS) to a single yoga session plotted on a three-dimensional
circumplex model (Valence, Activation & Attention). Labels indicate time points for affective responses: Baseline, Supine 1, Seated,
Kneeling, Standing, Supine 2, Supine Meditation, Post.
173
173
174
CHAPTER 5: GENERAL DISCUSSION
175
5.1 The Present Studies
The impetus for the present studies was to explore yoga’s ability to, “still the
fluctuations of the mind” in contemporary cancer settings. In these efforts, at least two
streams of research were combined: 1) the use and benefits of yoga practice in cancer
settings (51); and 2) how meditative states experienced within yoga practice are
correlated to psychophysiological systems that regulate positive affect and internalised
attention (58,59). The purpose was both to examine “if” and “to what extent” benefits
occur from practicing yoga for cancer survivors and explore by what mechanisms these
benefits accrue.
Research evidence suggests: 1) yoga practice is associated with improved HRQL
and psychological functioning; 2) a key definition of what yoga “is” or “does” entails
“stilling the fluctuations of the mind;” and 3) psychophysiological states elicited during
yoga practice are said to be associated with both positive affect and states of internalised
attention. Based on these statements, three studies were constructed that: 1)
systematically reviewed the current literature on yoga in cancer settings; 2) examined
longitudinal changes in HRQL and psychological outcomes, and their relations to affect
and attention-mindfulness in a community based yoga for cancer survivors program; and
3) examined the acute effects of a single yoga session on measures of core affect and
attention and concurrent associations with cardiac activity.
5.1.1 Integration of Theoretical Models, Constructs and Approaches
To aid in this examination of relations between yoga practice, affect, attention and
health outcomes, several existing theoretical models, constructs and approaches were
made use of, including: the Circumplex Model of Affect, Dual Mode Theory, the Effort-
176
Related Attention Model, the Neurovisceral Integration Model and a mixed methods
research approach, Neurophenomenology. The Circumplex Model provided the
overarching affective framework for both Chapters 3 and 4. The Dual Mode Theory
assisted in the investigation and explanation of acute changes in affect pre-post yoga
class in Chapter 3 and changes in affect during a single yoga session in Chapter 4. The
Effort-Related Attention Model described the use of associative and dissociative attention
at varying intensities in both Chapters 3 and 4. The construct of mindfulness facilitated
the exploration of several different “skills of attention” in Chapter 3. The Neurovisceral
Integration Model provided a theoretical rationale for including the use of HRV as a
neurophysiological correlate of autonomic activity, affect and attention regulation in
Chapter 4. Neurophenomenology allowed the entire scope of these affective, attentional
and neurophysiological processes to be subsumed under one large model and was used in
Chapter 4 to provide an overarching mixed methods approach to the examination of the
quantitative and qualitative data.
5.2 Study Findings
Having developed both a theoretical background and study objectives, research
began in Summer, 2010 on the systematic review which was completed in Fall, 2011 and
accepted for publication in Spring, 2012 by Evidence-Based Complementary and
Alternative Medicine. Recruitment for the longitudinal study began in the Fall, 2010 and
was completed in Summer, 2011. All data from the final six-month follow-up for the
longitudinal study was available by Winter, 2011. Recruitment for the lab study began in
Fall, 2011 and was completed in Spring, 2012. A summary of all three study findings
follows.
177
5.2.1 Study 1(Clinical Significance Review, Chapter 2)
Objective 1a: Yoga programs were offered primarily to breast cancer patients.
Classes were 60-90 minutes in length, met from one to five times a week and lasted from
6-26 weeks. Objective 1b: There were clinically significant effects of yoga practice on
HRQL, psychological indices and, to a lesser extent, fatigue and sleep indices (1).
5.2.2 Study 2 (Longitudinal Study, Chapter 3)
Objective 1a: Yoga practice was associated with increased valence, activation and
attention pre-post yoga class in the seven-week program. Objective 1b: These
improvements were related to baseline measures of affect and mindfulness. Objective 2a:
Significant decreases in mood disturbance and stress symptoms were observed post- yoga
program that were maintained at the three- and six month follow-ups, while HRQL
improved linearly across all time points. Energy improved linearly across all time points,
while tension and tiredness decreased post intervention, regressed towards baseline
values at the three-month follow-up and returned to post-program completion values at
the six-month follow-up. Mindfulness based skills in observation increased significantly
from baseline only at the six-month follow-up, while act with awareness and nonjudgement of inner experience both showed small-moderate increases post yoga program
that were maintained at the three- and six- month follow-ups. Small improvements in
non-reaction were seen at the three and six-month follow-ups. Objective 2b: Longitudinal
improvements in mood disturbance, stress symptoms and HRQL were related to
concomitant changes in affect and mindfulness. Objective 3a: 48% of the participants had
an existing yoga practice at baseline and 96% reported continued practice upon
completion of the seven-week program. 69% and 76% were still practicing yoga at the
178
three-and six-month follow-ups respectively. Objective 3b: Yoga practice maintenance
was related to previous yoga experience, energy and non-reaction to inner experience.
5.2.3 Study 3 (Lab Study, Chapter 4)
Objective 1a: Participation in a single yoga session was related to a linear
increase in valence and attention during class, while shifts in activation, perceived
exertion and cardiac activity were related to each yoga sequence. Objective 1b: Changes
in valence, activation, attention and perceived exertion were related to cardiac activity.
Objective 1c: First-person interviews largely confirmed quantitative findings. Participants
suggested increased associative attention and awareness of breath were potential
mechanisms for the effects of yoga during the single session. Objective 2a: Pre-post class
significant increases in energy and calmness were observed and reductions in tension and
tiredness. Objective 2b: Changes were related to one another and concomitant changes in
valence, activation, and cardiac activity. Objective 2c: First-person interviews confirmed
quantitative findings and participants attributed improvements in affect pre-post the
single yoga session specifically to yoga practice.
5.3 Theoretical Integration
5.3.1 Acute Effects
The present research highlights associations between yoga practice as a means of
“stilling the fluctuations of the mind,” affect and attention regulation, HRQL and mental
health outcomes. In both the longitudinal and lab studies, significant increases in valence,
activation and associative attention were observed pre-post class. Within the lab session,
increases in valence and associative attention were related to one another and moderated
by activation, perceived exertion and cardiac activity. The relationship between valence
179
and activation is corroborated by Ekkekakis and colleagues (73), whose research suggests
exercise below lactate threshold is generally reported as pleasant both during the actual
exercise session and upon completion. This lower-intensity level may have also led to a
continued primacy of positive cognitive attentional cues versus negatively-valenced
interoceptive cues seen at higher exercise intensities, as evidenced in the Effort-Based
Attention Model (91).
Research concerning the peak-end rule (275) suggests affect experienced at both
the highest intensity of exercise and conclusion of exercise predicts exercise adherence.
Given yoga is experienced at its affective and attentional peak in supine meditation, it is
not surprising findings indicate yoga is experienced positively both throughout and at the
end of the yoga session. Peak positive affect pre-post class may in part explain the high
rate of yoga practice maintenance, with 96% of participants reporting continued yoga
practice upon program completion and 69% and 76% of participants reporting ongoing
participation at three- and six-month follow-ups, though these relations were not assessed
directly. This is in contrast to commonly reported exercise maintenance rates in cancer
settings which average approximately 50% (26). This facilitation of positive affect in
cancer survivors may well lead to a high expectancy of program success for cancer
survivors and correspondingly increased exercise adherence (76).
5.3.2 Longitudinal Measures of Affect
The use of the Activation-Deactivation Checklist to measure changes in affect
longitudinally is uncommon. We ascribed to Russell’s (2003) thesis that when affective
dimensions are examined longitudinally, they are no longer tied to a specific context as in
acute exercise settings, where the temporal proximity to an exercise session is evident.
180
Rather, these measures become more reflective of general mood, or what Russell (2003)
has termed prolonged, objectless core affect (81). Higher energy at each time period was
related to reduced mood disturbance and improved HRQL. Conversely, higher tension
was related to increased stress symptoms and reduced HRQL, while tiredness was related
to increased mood disturbance. It could be argued including both the Profile of Mood
States and Activation-Deactivation Checklist created measurement redundancies, as both
measure moods. However, each measure is categorically different, as the POMS serves as
a categorical indicator of mood disturbance using discrete largely pathological categories
(with the exception of vigour), while the ADACL uses a dimensional model to describe a
range of negatively and positively valenced, high and low activation states. Given yoga
conceptually contains a range of affective states, including both low-activation pleasure
(calm) and high-activation pleasure (energy) and that individuals, particularly cancer
survivors for whom fatigue can figure so prominently, may also experience both lowactivation displeasure (tired) and high activation displeasure (tension), it behooves
researchers to utilize measures that capture the full range of affective experiences.
Furthermore, dimensional measures of affective states may mediate more categorical
mood states (68,214,215).
5.3.3 The Role of Attention in Affective Experience
While the role of valence and activation has been thoroughly studied in the
literature, the role of attention in these experiences is less clear (89,95,276). Yoga
facilitated a highly associative state related to positive affect in both the longitudinal and
lab studies. This associative attentional strategy has been correlated with measures of
mindfulness (96) as well as focused attention states (58,208).
181
These acute focused attention states are purported to develop three skills of
regulative attention (208): 1) the ability to monitor distractions without destabilizing the
intended focus; 2) the ability to disengage from distractions without further involvement;
and 3) the ability to redirect focus promptly on a chosen object. It is suggested these three
skills are related to significant decreases in emotional reactivity (208).
Longitudinal findings suggest participants reported a greater ability to act with
awareness, a mindfulness facet closely related to one’s ability to concentrate and a
corollary of associative and focused attention. Participants also reported increased nonjudging of inner experience, the taking of an impartial stance to thoughts and feelings,
and non-reaction to inner experience, the ability to let thoughts and emotions come and
go. Research suggests act with awareness, non judgment and non-reaction are most tied
to decreased psychological distress (180). The ability to act with awareness was
associated with lower mood disturbance and stress symptoms at all times, while nonjudging of inner experience was related to reduced stress symptoms and improved
HRQL. Non-reaction was related to reduced mood disturbance. These findings are
similar to Garland (183), in which both the ability to act with awareness and be nonjudgmental of inner experience was correlated with reduced stress and mood disturbance.
While non-reaction to inner experience was not related to any health outcomes in the
Garland study, research with experienced meditators and non-meditators suggests the
ability to be non-reactive to inner experience is highly related to meditation experience
and is most likely to mediate psychological health outcomes (216). The lack of change in
both the observe and describe facets in the longitudinal study may have been due to
182
participant unfamiliarity with the specific mindfulness constructs described in these
subscales (212).
5.3.4 Autonomic Concomitants of Affect and Attention
Both the neurovisceral integration model and neurophenomenological approach
were of primary theoretical interest in the lab study. These models highlight the dynamic
interactions between physiological, cognitive and behavioral systems in the regulation of
affect and attention (104,106,240). Overall, heart rate and HRV indices indicate a period
of vagal withdrawal up to the standing poses, reflecting the intensity of the yoga session.
These values begin to return to resting values during the supine sequence. During the
supine meditation increased HRV in the time domains and decreased heart rate are
evident. These findings seem to corroborate research that suggests yoga’s active poses
create physiological stress similar to exercise (114) while the resting poses heighten
autonomic activity, however briefly (115). Valence and activation were moderated by
high and low SDNN respectively, a correlate of total heart rate complexity. Attention was
related to sample entropy, such that higher sample entropy predicted a significant
increase in associative attention while lower sample entropy resulted in no significant
change in attention throughout the intervention. As expected, perceived exertion was
linked to participant heart rate (277). However, in contrast to findings reported using
Tenenbaum’s Effort-Related Attention model (91,93), attention was not related to
perceived exertion but rather core affect, specifically higher valence and lower activation.
This suggests the regulation of attention may not be intensity-dependent in the case of
yoga. Higher energy was related to higher HF HRV and SDNN and lower RMSSD, while
tension was related to lower HRV. Lower tiredness was related to higher resting heart
183
rate, and calmness was related to higher resting heart rate. In the case of both calmness
and tiredness, associations with higher resting heart rate may be related to higher states of
awareness, and concomitant increases in heart rate that affect low-activation pleasant and
unpleasant affective states (58). This increased heart rate may be an example of the socalled “meditation paradox” (278) in which a given meditative technique may elicit a
higher mean heart rate in states that are objectively relaxing. However, in examining
mean heart rates in the estimated marginal means models, heart rate during supine
meditation and post yoga session are significantly lower than baseline. Nor was attention
associated with calmness or tension. Overall, findings support the neurovisceral
integration models main thesis that both attention and affect are linked via underlying
autonomic function (106). Specifically, associations between valence, activation,
attention and perceived exertion and HRV were all in the expected directions during the
yoga sequence, as were both energy and tension during baseline, supine meditation and
post-resting states.
5.3.5 Integration of the Neurophenomenological Approach
Utilising the neurophenomenological approach allowed us to further compare data
at all levels, including patient-reported measures of valence, activation, attention and
perceived exertion during the yoga sequences, continuously monitored heart rate and
HRV data, pre-post yoga session measures of affect, and participant’s subjective
experiences of these conditions. These detailed descriptions informed both the analysis
and its interpretation, highlighting that participants themselves viewed changes in
attention both during class and pre-post class as a primary mechanisms of engagement in
yoga practice and in deriving affective benefits. Participants further suggested awareness
184
of breath was a primary technique to achieve these aims. These subjective reports
highlighted elements of the data that were not observable by strict quantitative analysis
alone and provided much needed context for the quantitative research findings (122).
Taken together, findings suggest yoga practice facilitates an increasingly
associative, positively-valenced affective state both during and after a single yoga
session. Longitudinally, measures of affect and mindfulness were related to changes in
mood disturbance, stress symptoms and HRQL and maintenance of yoga practice.
Findings from the lab-based study suggest both valence and attention have autonomic
correlates within yoga practice and preliminary neurophenomenological findings suggest
these changes may be related to increased associative attention, particularly to sensations
of the breath and body during yoga practice.
5.4 The Role of Yoga in Exercise Settings
5.4.1 Yoga as Cardiovascular Exercise
Given the regulation of valence, activation and attention in cancer survivors both
acutely during yoga practice and over time with practice, an additional inquiry might be
what is the role of yoga in exercise settings in general and within the scope of cancer
treatment specifically?
A common comment in the literature is that the intensity of most yoga practices is
not sufficient enough to elicit the cardiovascular benefits of regular exercise training
(279). Research by Hagins and colleagues (280) suggests in a typical yoga session,
healthy participants’ heart rate averaged only 49.4% of maximum heart rate, peaking at
54.8% of maximum heart during a sun salutation sequence. These findings are similar to
the current lab study, in which participants reached 52% of their maximum heart rate
185
during the standing sequence and this intensity was maintained for approximately 10
minutes of a 75 minute yoga practice. The authors conclude yoga is a low-intensity form
of exercise that does not meet the recommendations for maintaining or improving health
or cardiovascular fitness (280).
However, several studies have found yoga does lead to reductions in age-related
deterioration in cardiovascular function. Specifically, yoga practitioners with five or more
years of experience report reduced heart rate, systolic and diastolic blood pressure in
comparison to age-matched controls (281). These findings were further corroborated in
an intervention study in which reductions in heart rate, systolic and diastolic blood
pressure were seen over a period of six months in which participants engaged in an hour
of daily yoga training (282). Yoga was also linked to improved exercise tolerance after
45 minutes of daily yoga training for two months (283) and improved pulmonary
function after a minimum of ten weeks of training (284). A clinical review further
suggests yoga can be beneficial in both primary and secondary prevention of
cardiovascular disease (285). Mechanisms for yoga specific improvement in
cardiovascular health are related primarily to increased breathing efficiency and
improved cardiac autonomic function (111,286). Future research directions in this area
include appropriately matched comparative RCTs in which both exercise and yoga are
tested concurrently. Given the laboratory findings, both a cross-sectional study and
longitudinal RCT could be designed in which HRV is measured daily in matched yoga
practitioners and exercisers to examine the cardiovascular benefits in both populations.
186
5.4.2 Yoga as Relaxation
Perhaps one of the strongest psychophysiological arguments for what yoga “does”
is the elicitation of the relaxation response. The relaxation response is a set of integrated
physiological changes elicited when a subject assumes a relaxed position and engages in
repetitive mental or physical action, while passively ignoring distracting thoughts (287).
This relaxed state has been associated with decreases in oxygen consumption, respiratory
rate, and blood pressure, and an increased sense of well-being (288) and has been used in
the treatment of immunological, cardiovascular, neurodegenerative diseases and
psychiatric disorders (289). A review by Khalsa (223) characterises yoga as a coordinated
psychophysiological relaxation response that leads to reductions in both cognitive and
somatic arousal and modifies the activity of the autonomic nervous system. In yoga
practice, a primary mechanism by which these benefits of relaxation occur are attention
to the breath and specific breathing practices. The regular practice of breathing exercises
increases parasympathetic tone, decreases sympathetic activity, improves cardiovascular
and respiratory functions, decreases the effects of stress on the body, and improves
physical and mental health (110,290). When combined with meditative practice, effects
of slow deep breathing are significantly greater as compared to other relaxation
techniques (286).
5.4.3 Yoga as Yoga
But is yoga more than just a set of relaxation and breathing exercises? Telles
suggests yoga elicits meditative states in which deep relaxation and increased focus of
attention co-exist (260). This state is putatively different than relaxation or sleep states, in
that one is still alert, mental activity is actively reduced as attention increases, and
187
positive affect increases (58). Traditional meditation texts in the yoga tradition suggest
meditative states (as opposed to meditative techniques) in yoga are characterized by a
progression from concentration or focused attention (dharana) to contemplation
(dhyana), a meditative state characterised by passive attention akin to mindfulness, and
finally absorption (samadhi), a state of mental containment in which the mind does not
waver (59,208). These meditative states, and the associated practices designed to elicit
them, can be better viewed as a family of complex affective and attentional regulatory
training regimes developed for various ends, including cultivation of well-being and
emotional balance (208). Yoga may outwardly share commonalities with both exercise
and relaxation, but may also offer unique meditative benefits that “still the fluctuations of
the mind.”
5.5 Concluding Thoughts
Current findings suggest yoga does indeed, “still the fluctuations of the mind.” In
a single session, yoga practice up-regulates both positive valence and associative
attention linearly while activation, exertion, and autonomic function vary based on the
intensity of any given yoga sequence. Over time, this leads to an increased energy,
decreased tension and tiredness and an improved ability to act with awareness and be
both non-judgemental of and non reactive-to inner experience. These affective and
attentional mechanisms are related to reduced mood disturbance and stress symptoms and
improved HRQL. Ultimately, previous yoga experience, higher energy and a greater
ability to let thoughts come and go leads to increased maintenance of yoga practice,
arguably further engendering the benefits of these practices (see Figure 5.1).
188
Within this model, yoga is a unique form of exercise that creates a confluence of
positive affect and associative attention that effectively stills the fluctuations of the mind
via heightened attention to the breath and body. Given the high rate of distress in cancer
patients and related lower HRQL, and despite the benefits conventional aerobic exercise
may offer in these areas, the addition of yoga to conventional cancer treatment is
warranted. Yoga may serve as a complement to common forms of aerobic and resistance
exercise that emphasises yoga’s affective and attentional benefits, soothing both mind
and body. Given the identified complexity of the cancer experience, the unique
challenges faced in each diagnosis, and the tools that yoga offers, highly prescriptive
interventions may be developed to aid cancer survivors in their personal quest to still the
fluctuations of the mind. While specific cancer treatment must remain a top priority, the
health of the overall psychophysiological system must not be overlooked in meeting these
aims (61). A reasoned approach that addresses both issues is sure to encourage adherence
to treatment and maintenance of health behaviours antagonistic to cancer recurrence but
supportive of cancer survivor health. This research is very timely given the large number
of practitioners of yoga world-wide, the many touted benefits of yoga and the
contemporary discipline’s connection to contemplative teachings that lend themselves to
this type of examination.
Figure 5.1 Yoga Mechanisms Integrated Model.
189
189
190
Bibliography
(1) Culos-Reed SN, Mackenzie MJ, Sohl SJ, Jesse MT, Ross-Zahavich AN, Danhauer
SC. Yoga & Cancer Interventions: A review of the clinical significance of patientreported outcomes for cancer survivors. Evidence-based complementary and alternative
medicine : eCAM 2012;2012:1-17.
(2) Lin KY, Hu YT, Chang KJ, Lin HF, Tsauo JY. Effects of yoga on psychological
health, quality of life, and physical health of patients with cancer: a meta-analysis. Evid
Based Complement Alternat Med 2011;2011:659876.
(3) Smith JP. Applied Multivariate Statistics for the Social Sciences. Fifth ed. New York:
Routledge; 2009.
(4) Canadian Cancer Society’s Steering Committee on Cancer Statistics. Canadian
Cancer Statistics 2012. 2012.
(5) Patrick DL, Ferketich SL, Frame PS, Harris JJ, Hendricks CB, Levin B, et al.
National Institutes of Health State-of-the-Science Conference Statement: Symptom
Management in Cancer: Pain, Depression, and Fatigue, July 15-17, 2002. J Natl Cancer
Inst 2003 Aug 6;95(15):1110-1117.
(6) Hacker E. Exercise and quality of life: strengthening the connections. Clin J Oncol
Nurs 2009 Feb;13(1):31-39.
191
(7) Carlson LE, Waller A, Groff SL, Giese-Davis J, Bultz BD. What goes up does not
always come down: patterns of distress, physical and psychosocial morbidity in people
with cancer over a one year period. Psychooncology 2011 Oct 4.
(8) Brown JC, Huedo-Medina TB, Pescatello LS, Pescatello SM, Ferrer RA, Johnson BT.
Efficacy of exercise interventions in modulating cancer-related fatigue among adult
cancer survivors: a meta-analysis. Cancer Epidemiol Biomarkers Prev 2011
Jan;20(1):123-133.
(9) Brown LF, Kroenke K. Cancer-related fatigue and its associations with depression
and anxiety: a systematic review. Psychosomatics 2009 Sep-Oct;50(5):440-447.
(10) Laird BJ, Boyd AC, Colvin LA, Fallon MT. Are cancer pain and depression
interdependent? A systematic review. Psychooncology 2009 May;18(5):459-464.
(11) Pinquart M, Duberstein PR. Depression and cancer mortality: a meta-analysis.
Psychol Med 2010 Nov;40(11):1797-1810.
(12) Satin JR, Linden W, Phillips MJ. Depression as a predictor of disease progression
and mortality in cancer patients: a meta-analysis. Cancer 2009 Nov 15;115(22):53495361.
(13) Courneya KS, Friedenreich CM, Reid RD, Gelmon K, Mackey JR, Ladha AB, et al.
Predictors of follow-up exercise behavior 6 months after a randomized trial of exercise
training during breast cancer chemotherapy. Breast Cancer Res Treat 2009
Mar;114(1):179-187.
192
(14) Spence RR, Heesch KC, Brown WJ. Exercise and cancer rehabilitation: a systematic
review. Cancer Treat Rev 2010 Apr;36(2):185-194.
(15) Young-McCaughan S, Arzola SM. Exercise Intervention Research for Patients With
Cancer on Treatment. Semin Oncol Nurs 2007 11;23(4):264-274.
(16) Blanchard CM, Courneya KS, Stein K. Cancer survivors' adherence to lifestyle
behavior recommendations and associations with health-related quality of life: results
from the American Cancer Society's SCS-II. J Clin Oncol 2008 May 1;26(13):2198-2204.
(17) Knobf MT, Musanti R, Dorward J. Exercise and quality of life outcomes in patients
with cancer. Semin Oncol Nurs 2007 Nov;23(4):285-296.
(18) Courneya KS, Tamburrini AL, Woolcott CG, McNeely ML, Karvinen KH,
Campbell KL, et al. The Alberta Physical Activity and Breast Cancer Prevention Trial:
quality of life outcomes. Prev Med 2011 Jan;52(1):26-32.
(19) McNeely ML, Courneya KS. Exercise programs for cancer-related fatigue: evidence
and clinical guidelines. J Natl Compr Canc Netw 2010 Aug;8(8):945-953.
(20) Barbaric M, Brooks E, Moore L, Cheifetz O. Effects of Physical Activity on Cancer
Survival: A Systematic Review. Physiotherapy Canada 2010;62(1):25-34.
(21) Hamer M, Stamatakis E, Saxton JM. The impact of physical activity on all-cause
mortality in men and women after a cancer diagnosis. Cancer Causes Control 2009
Mar;20(2):225-231.
193
(22) Speed-Andrews AE, Courneya KS. Effects of exercise on quality of life and
prognosis in cancer survivors. Curr Sports Med Rep 2009 Jul-Aug;8(4):176-181.
(23) Schmitz KH, Courneya KS, Matthews C, Demark-Wahnefried W, Galvao DA, Pinto
BM, et al. American College of Sports Medicine roundtable on exercise guidelines for
cancer survivors. Med Sci Sports Exerc 2010 Jul;42(7):1409-1426.
(24) Doyle C, Kushi LH, Byers T, Courneya KS, Demark-Wahnefried W, Grant B, et al.
Nutrition and physical activity during and after cancer treatment: an American Cancer
Society guide for informed choices. CA Cancer J Clin 2006 Nov-Dec;56(6):323-353.
(25) Courneya KS, Katzmarzyk PT, Bacon E. Physical activity and obesity in Canadian
cancer survivors: population-based estimates from the 2005 Canadian Community Health
Survey. Cancer 2008 Jun;112(11):2475-2482.
(26) Maddocks M, Mockett S, Wilcock A. Is exercise an acceptable and practical therapy
for people with or cured of cancer? A systematic review. Cancer Treat Rev 2009
6;35(4):383-390.
(27) Haskell WL, Lee IM, Pate RR, Powell KE, Blair SN, Franklin BA, et al. Physical
activity and public health: updated recommendation for adults from the American
College of Sports Medicine and the American Heart Association. Med Sci Sports Exerc
2007 Aug;39(8):1423-1434.
194
(28) Trost SG, Owen N, Bauman AE, Sallis JF, Brown W. Correlates of adults'
participation in physical activity: review and update. Med Sci Sports Exerc 2002
Dec;34(12):1996-2001.
(29) Courneya KS, Freidenreich CM. Physical Activity and Cancer Control. Seminars in
Oncology Nursing 2007;23(4):242-252.
(30) Wolin KY, Schwartz AL, Matthews CE, Courneya KS, Schmitz KH. Implementing
the Exercise Guidelines for Cancer Survivors. The Journal of Supportive Oncology 2012
0;10(5):171-177.
(31) Rock CL, Doyle C, Demark-Wahnefried W, Meyerhardt J, Courneya KS, Schwartz
AL, et al. Nutrition and physical activity guidelines for cancer survivors. CA: A Cancer
Journal for Clinicians 2012;62(4):242-274.
(32) McNeely ML, Peddle CJ, Parliament M, Courneya KS. Cancer Rehabilitation:
Recommendations for Integrating Exercise Programming in the Clinical Practice Setting.
Current Cancer Therapy Reviews 2006 11;2(4):351-360.
(33) Smith KB, Pukall CF. An evidence-based review of yoga as a complementary
intervention for patients with cancer. Psychooncology 2009 May;18(5):465-475.
(34) Cramer H, Lange S, Klose P, Paul A, Dobos G. Yoga for breast cancer patients and
survivors: a systematic review and meta-analysis. BMC Cancer 2012 Sep 18;12(1):412.
(35) Cramer H, Lange S, Klose P, Paul A, Dobos G. Can yoga improve fatigue in breast
cancer patients? A systematic review. Acta Oncol 2012 Apr;51(4):559-560.
195
(36) Boehm K, Ostermann T, Milazzo S, Bussing A. Effects of yoga interventions on
fatigue: a meta-analysis. Evid Based Complement Alternat Med 2012;2012:124703.
(37) Zhang J, Yang KH, Tian JH, Wang CM. Effects of Yoga on Psychologic Function
and Quality of Life in Women with Breast Cancer: A Meta-Analysis of Randomized
Controlled Trials. J Altern Complement Med 2012 Aug 21.
(38) Evans S, Tsao JCI, Sternlieb B, Zeltzer LK. Using the Biopsychosocial Model to
Understand the Health Benefits of Yoga. Journal of Complementary & Integrative
Medicine 2009 01;6(1):1-22.
(39) Field T. Yoga clinical research review. Complement Ther Clin Pract 2011
Feb;17(1):1-8.
(40) Bussing A, Michalsen A, Khalsa SB, Telles S, Sherman KJ. Effects of yoga on
mental and physical health: a short summary of reviews. Evid Based Complement
Alternat Med 2012;2012:165410.
(41) Sengupta P. Health Impacts of Yoga and Pranayama: A State-of-the-Art Review. Int
J Prev Med 2012 Jul;3(7):444-458.
(42) Roland KP, Jakobi JM, Jones GR. Does yoga engender fitness in older adults? A
critical review. J Aging Phys Act 2011 Jan;19(1):62-79.
(43) da Silva TL, Ravindran LN, Ravindran AV. Yoga in the treatment of mood and
anxiety disorders: A review. Asian Journal of Psychiatry 2009 3;2(1):6-16.
196
(44) Sharma M, Haider T. Yoga as an Alternative and Complementary Therapy for
Patients Suffering From Anxiety: A Systematic Review. Journal of Evidence-Based
Complementary & Alternative Medicine 2012 September 20.
(45) Uebelacker LA, Epstein-Lubow G, Gaudiano BA, Tremont G, Battle CL, Miller IW.
Hatha Yoga for Depression: Critical Review of the Evidence for Efficacy, Plausible
Mechanisms of Action, and Directions for Future Research. Journal of Psychiatric
Practice® 2010;16(1).
(46) Chong CS, Tsunaka M, Tsang HW, Chan EP, Cheung WM. Effects of yoga on stress
management in healthy adults: A systematic review. Altern Ther Health Med 2011 JanFeb;17(1):32-38.
(47) Patel NK, Newstead AH, Ferrer RL. The Effects of Yoga on Physical Functioning
and Health Related Quality of Life in Older Adults: A Systematic Review and MetaAnalysis. J Altern Complement Med 2012 Aug 21.
(48) Ross A, Thomas S. The health benefits of yoga and exercise: a review of comparison
studies. J Altern Complement Med 2010 Jan;16(1):3-12.
(49) Prasad J. The Date of the Yoga-Sūtras. Journal of the Royal Asiatic Society of Great
Britain and Ireland 1930 Apr.(2):pp. 365-375.
(50) Singleton M. Yoga Body: The Origins of Modern Posture Practice. New York, NY:
Oxford University Press; 2010.
197
(51) Bower JE, Woolery A, Sternleib B, Garet D. Yoga for cancer patients and survivors.
Cancer Control 2005;12(3):165-171.
(52) Alter JS. Modern Medical Yoga: Struggling with a History of Magic, Alchemy and
Sex. Asian Medicine 2005 01;1(1):119-146.
(53) Alter JS. Yoga and Physical Education: Swami Kuvalayananda's Nationalist Project.
Asian Medicine 2007 01;3(1):20-36.
(54) Singleton M. Salvation through Relaxation: Proprioceptive Therapy and its
Relationship to Yoga. Journal of Contemporary Religion 2005 10/01; 2012/10;20(3):289304.
(55) Rao K. Applied Yoga Psychology: Studies of Neurophysiology of Meditation.
Journal of Consciousness Studies 2011;18(11-12):161-198.
(56) Prabhavananda, Isherwood C. How to Know God: The Yoga Aphorisms of
Patanjali. : Vedanta Society of Southern California; 1953.
(57) Desikachar TKV, Cravens RH. Health, Healing & Beyond: Yoga and the Living
Tradition of Krishnamacharya New York, NY: Aperture Foundation; 1998.
(58) Rubia K. The neurobiology of Meditation and its clinical effectiveness in psychiatric
disorders. Biol Psychol 2009 Sep;82(1):1-11.
(59) Telles S, Raghavendra BR. Neurophysiological Changes in Meditation Correlated
with Descriptions from the Ancient Texts. Biofeedback 2011;39(2):56-59.
198
(60) Satyapriya M, Nagendra HR, Nagarathna R, Padmalatha V. Effect of integrated yoga
on stress and heart rate variability in pregnant women. Int J Gynaecol Obstet 2009
Mar;104(3):218-222.
(61) Taylor AG, Goehler LE, Galper DI, Innes KE, Bourguignon C. Top-down and
bottom-up mechanisms in mind-body medicine: development of an integrative
framework for psychophysiological research. Explore (NY) 2010 Jan;6(1):29-41.
(62) Osoba D. Health-related quality of life and cancer clinical trials. Ther Adv Med
Oncol 2011 Mar;3(2):57-71.
(63) Craig P, Dieppe P, Macintyre S, Michie S, Nazareth I, Petticrew M, et al.
Developing and evaluating complex interventions: the new Medical Research Council
guidance. BMJ 2008 Sep 29;337:a1655.
(64) Demark-Wahnefried W, Jones LW. Promoting a Healthy Lifestyle Among Cancer
Survivors. Hematol Oncol Clin North Am 2008 4;22(2):319-342.
(65) Humpel N, Iverson DC. Review and critique of the quality of exercise
recommendations for cancer patients and survivors. Support Care Cancer 2005
Jul;13(7):493-502.
(66) Perkins HY, Baum GP, Taylor CL, Basen-Engquist KM. Effects of treatment
factors, comorbidities and health-related quality of life on self-efficacy for physical
activity in cancer survivors. Psychooncology 2009 Apr;18(4):405-411.
199
(67) Pressman SD, Cohen S. Does Positive Affect Influence Health? Psychol Bull 2005
November;131(6):925-971.
(68) Hou WK, Law CC, Fu YT. Does change in positive affect mediate and/or moderate
the impact of symptom distress on psychological adjustment after cancer diagnosis? A
prospective analysis. Psychol Health 2010 Apr;25(4):417-431.
(69) Hirsch JK, Floyd AR, Duberstein PR. Perceived health in lung cancer patients: the
role of positive and negative affect. Qual Life Res 2011 May 25.
(70) Sepah SC, Bower JE. Positive affect and inflammation during radiation treatment for
breast and prostate cancer. Brain Behav Immun 2009 11;23(8):1068-1072.
(71) Reed J, Buck S. The effect of regular aerobic exercise on positive-activated affect: A
meta-analysis. Psychol Sport Exerc 2009 11;10(6):581-594.
(72) Reed J, Ones DS. The effect of acute aerobic exercise on positive activated affect: A
meta-analysis. Psychol Sport Exerc 2006 9;7(5):477-514.
(73) Ekkekakis P, Parfitt G, Petruzzello SJ. The Pleasure and Displeasure People Feel
When they Exercise at Different Intensities: Decennial Update and Progress towards a
Tripartite Rationale for Exercise Intensity Prescription. Sports Med 2011 Aug
1;41(8):641-671.
(74) Garber CE, Blissmer B, Deschenes MR, Franklin BA, Lamonte MJ, Lee IM, et al.
Quantity and quality of exercise for developing and maintaining cardiorespiratory,
200
musculoskeletal, and neuromotor fitness in apparently healthy adults: guidance for
prescribing exercise. Med Sci Sports Exerc 2011 Jul;43(7):1334-1359.
(75) Williams DM. Exercise, Affect, and Adherence: An Integrated Model and a Case for
Self-Paced Exercise. J Sport Exercise Psychol 2008 10;30(5):471-496.
(76) Courneya KS, Friedenreich CM, Sela RA, Quinney HA, Rhodes RE, Jones LW.
Exercise motivation and adherence in cancer survivors after participation in a randomized
controlled trial: an attribution theory perspective. Int J Behav Med 2004;11(1):8-17.
(77) West J, Otte C, Geher K, Johnson J, Mohr DC. Effects of Hatha yoga and African
dance on perceived stress, affect, and salivary cortisol. Ann Behav Med 2004
Oct;28(2):114-118.
(78) Netz Y, Lidor R. Mood alterations in mindful versus aerobic exercise modes. J
Psychol 2003 Sep;137(5):405-419.
(79) Danhauer SC, Mihalko SL, Russell GB, Campbell CR, Felder L, Daley K, et al.
Restorative yoga for women with breast cancer: findings from a randomized pilot study.
Psychooncology 2009 Apr;18(4):360-368.
(80) Vadiraja HS, Rao MR, Nagarathna R, Nagendra HR, Rekha M, Vanitha N, et al.
Effects of yoga program on quality of life and affect in early breast cancer patients
undergoing adjuvant radiotherapy: A randomized controlled trial. Complement Ther Med
2009 12;17(5-6):274-280.
201
(81) Russell JA. Core Affect and the Psychological Construction of Emotion. Psychol
Rev 2003 January;110(1):145-172.
(82) Posner J, Russell JA, Peterson BS. The circumplex model of affect: an integrative
approach to affective neuroscience, cognitive development, and psychopathology. Dev
Psychopathol 2005 Summer;17(3):715-734.
(83) Yik MSM, Russell JA, Barrett LF. Structure of Self-Reported Current Affect:
Integration and Beyond. Journal of Personality & Social Psychology 1999
September;77(3):600-619.
(84) Ekkekakis P. The Dual-Mode Theory of affective responses to exercise in
metatheoretical context: I. Initial impetus, basic postulates, and philosophical framework.
International Review of Sport and Exercise Psychology 2009;2(1):73-94.
(85) Ekkekakis P. The Dual-Mode Theory of affective responses to exercise in
metatheoretical context: II. Bodiless heads, ethereal cognitive schemata, and other
improbable dualistic creatures, exercising. International Review of Sport and Exercise
Psychology 2009;2(2):139-160.
(86) Ekkekakis P, Hall EE, Petruzzello SJ. The relationship between exercise intensity
and affective responses demystified: to crack the 40-year-old nut, replace the 40-year-old
nutcracker! Ann Behav Med 2008 Apr;35(2):136-149.
(87) McArdle WM, Katch FI, Katch V,L. Exercise Physiology: Nutrition, Energy and
Human Performance. 7th ed. Baltimore, MD: Lippincott Williams & Wilkins; 2009.
202
(88) Morgan WP, Pollock ML. Psychological Characterization of the Elite Distance
Runner. Ann N Y Acad Sci 1977;301(1):382-403.
(89) Lind E, Welch AS, Ekkekakis P. Do 'mind over muscle' strategies work? Examining
the effects of attentional association and dissociation on exertional, affective and
physiological responses to exercise. Sports Med 2009;39(9):743-764.
(90) LaCaille RA, Masters KS, Heath EM. Effects of cognitive strategy and exercise
setting on running performance, perceived exertion, affect, and satisfaction. Psychol
Sport Exerc 2004 10;5(4):461-476.
(91) Tenenbaum G. A social-cognitive perspective on perceived exertion and exertion
tolerance. In: Singer RN, Hausenblas HA, Janelle CM, editors. Handbook of Sport
Psychology. 2nd ed. Hoboken, NJ: John Wiley & Sons Inc.; 2001. p. 810-820.
(92) Tenenbaum G, Connolly CT. Attention allocation under varied workload and effort
perception in rowers. Psychol Sport Exerc 2008 9;9(5):704-717.
(93) Hutchinson JC, Tenenbaum G. Attention focus during physical effort: The mediating
role of task intensity. Psychol Sport Exerc 2007 3;8(2):233-245.
(94) Brown KW, Cordon S. Toward a Phenomenology of Mindfulness: Subjective
Experience and Emotional Correlates. In: Didonna F, editor. Clincal Handbook of
Mindfulness New York, NY: Springer; 2009. p. 59-81.
203
(95) Salmon P, Hanneman S, Harwood B. Associative / Dissociative Cognitive Strategies
in Sustained Physical Activity: Literature Review and Proposal for a Mindfulness-Based
Conceptual Model. Sport Psychologist 2010 06;24(2):127-156.
(96) Salmon P, Lush E, Jablonski M, Sephton SE. Yoga and Mindfulness: Clinical
Aspects of an Ancient Mind/Body Practice. Cognitive and Behavioral Practice 2009
2;16(1):59-72.
(97) Froeliger BE, Garland EL, Modlin LA, McClernon FJ. Neurocognitive correlates of
the effects of yoga meditation practice on emotion and cognition: a pilot study. Front
Integr Neurosci 2012;6:48.
(98) Lengacher CA, Johnson-Mallard V, Post-White J, Moscoso MS, Jacobsen PB, Klein
TW, et al. Randomized controlled trial of mindfulness-based stress reduction (MBSR) for
survivors of breast cancer. Psychooncology 2009 Dec;18(12):1261-1272.
(99) Carmody J, Baer RA. Relationships between mindfulness practice and levels of
mindfulness, medical and psychological symptoms and well-being in a mindfulnessbased stress reduction program. J Behav Med 2008 Feb;31(1):23-33.
(100) Sauer-Zavala ,Shannon, Walsh ,Erin, Eisenlohr-Moul ,Tory, Lykins ,Emily.
Comparing Mindfulness-Based Intervention Strategies: Differential Effects of Sitting
Meditation, Body Scan, and Mindful Yoga. Mindfulness 2012:1-6.
204
(101) Task Force of the European Society of Cardiology and the North American Society
of Pacing and Electrophysiology. Heart rate variability: Standards of measurement,
physiological interpretation, and clinical use. Eur Heart J 1996 Mar;17(3):354-381.
(102) Thayer JF, Åhs F, Fredrikson M, Sollers III JJ, Wager TD. A meta-analysis of heart
rate variability and neuroimaging studies: Implications for heart rate variability as a
marker of stress and health. Neuroscience & Biobehavioral Reviews 2012 2;36(2):747756.
(103) Denver JW, Reed SF, Porges SW. Methodological issues in the quantification of
respiratory sinus arrhythmia. Biol Psychol 2007 Feb;74(2):286-294.
(104) Appelhans BM, Luecken LJ. Heart Rate Variability as an Index of Regulated
Emotional Responding. Review of General Psychology September 2006;10(3):229-240.
(105) Mauss IB, Robinson MD. Measures of emotion: A review. Cogn Emot 2009 Feb
1;23(2):209-237.
(106) Thayer JF, Lane RD. A model of neurovisceral integration in emotion regulation
and dysregulation. J Affect Disord 2000 Dec;61(3):201-216.
(107) Rajendra Acharya U, Paul Joseph K, Kannathal N, Lim CM, Suri JS. Heart rate
variability: a review. Med Biol Eng Comput 2006 Dec;44(12):1031-1051.
(108) Khattab K, Khattab AA, Ortak J, Richardt G, Bonnemeier H. Iyengar yoga
increases cardiac parasympathetic nervous modulation among healthy yoga practitioners.
Evid Based Complement Alternat Med 2007 Dec;4(4):511-517.
205
(109) Streeter CC, Gerbarg PL, Saper RB, Ciraulo DA, Brown RP. Effects of yoga on the
autonomic nervous system, gamma-aminobutyric-acid, and allostasis in epilepsy,
depression, and post-traumatic stress disorder. Med Hypotheses 2012 5;78(5):571-579.
(110) Pal GK, Velkumary S, Madanmohan. Effect of short-term practice of breathing
exercises on autonomic functions in normal human volunteers. Indian J Med Res 2004
Aug;120(2):115-121.
(111) Mourya M, Mahajan AS, Singh NP, Jain AK. Effect of slow- and fast-breathing
exercises on autonomic functions in patients with essential hypertension. J Altern
Complement Med 2009 Jul;15(7):711-717.
(112) Telles S, Joshi M, Dash M, Raghuraj P, Naveen KV, Nagendra HR. An Evaluation
of the Ability to Voluntarily Reduce the Heart Rate after a Month of Yoga Practice.
Integrative Physiological & Behavioral Science 2004 Apr;39(2):119-125.
(113) Telles S, Raghavendra BR, Naveen KV, Manjunath NK, Kumar S, Subramanya P.
Changes in Autonomic Variables Following Two Meditative States Described in Yoga
Texts. J Altern Complement Med 2012 Sep 4.
(114) Cowen VS, Adams TB. Heart rate in yoga asana practice: A comparison of styles. J
Bodywork Movement Ther 2007 1;11(1):91-95.
(115) Sarang P, Telles S. Effects of Two Yoga Based Relaxation Techniques on Heart
Rate Variability (HRV). International Journal of Stress Management 2006
November;13(4):460-475.
206
(116) Thayer JF, Lane RD. Claude Bernard and the heart–brain connection: Further
elaboration of a model of neurovisceral integration. Neuroscience & Biobehavioral
Reviews 2009 2;33(2):81-88.
(117) Thayer JF, Ruiz-Padial E. Neurovisceral integration, emotions and health: An
update. Int Congr Ser 2006 4;1287:122-127.
(118) Weinstein AA, Deuster PA, Kop WJ. Heart Rate Variability as a Predictor of
Negative Mood Symptoms Induced by Exercise Withdrawal. Medicine & Science in
Sports & Exercise 2007 April;39(4):735-741.
(119) Sakuragi S, Sugiyama Y. Effects of daily walking on subjective symptoms, mood
and autonomic nervous function. J Physiol Anthropol 2006 Jul;25(4):281-289.
(120) Ray US, Pathak A, Tomer OS. Hatha yoga practices: energy expenditure,
respiratory changes and intensity of exercise. Evid Based Complement Alternat Med
2011;2011:241294.
(121) Lutz A, Thompson E. Neurophenomenology: Integrating Brain Dynamics in the
Neuroscience of Consciousness. Journal of Consciousness Studies 2003;10(9-10):31-52.
(122) Lutz A. Toward a neurophenomenology as an account of generative passages: a
first empirical case study. Phenomenology and the Cognitive Sciences 2002;1(2):133167.
(123) Thompson E. Neurophenomenology and Contemplative Experience. New York:
Oxford University Press; 2006. p. 226-235.
207
(124) Farb NAS, Segal ZV, Mayberg H, Bean J, McKeon D, Fatima Z, et al. Attending to
the present: mindfulness meditation reveals distinct neural modes of self-reference.
Social Cognitive and Affective Neuroscience 2007 December 01;2(4):313-322.
(125) Lutz A, Greischar LL, Rawlings NB, Ricard M, Davidson RJ. Long-term
meditators self-induce high-amplitude gamma synchrony during mental practice.
Proceedings of the National Academy of Sciences of the United States of America 2004
November 16;101(46):16369-16373.
(126) Lutz A, Brefczynski-Lewis J, Johnstone T, Davidson RJ. Regulation of the Neural
Circuitry of Emotion by Compassion Meditation: Effects of Meditative Expertise. PLoS
ONE 2008 03/26;3(3):1897.
(127) Lutz A, Slagter HA, Rawlings NB, Francis AD, Greischar LL, Davidson RJ.
Mental training enhances attentional stability: neural and behavioral evidence. J Neurosci
2009 Oct 21;29(42):13418-13427.
(128) Lutz A, Lachaux JP, Martinerie J, Varela FJ. Guiding the study of brain dynamics
by using first-person data: synchrony patterns correlate with ongoing conscious states
during a simple visual task. Proc Natl Acad Sci U S A 2002 Feb 5;99(3):1586-1591.
(129) Thompson E. Contemplative Neuroscience as an Approach to Volitional
Consciousness. In: Murphy N, Ellis G, O’Connor T, editors. Downward Causation and
the Neurobiology of Free Will: Springer Berlin / Heidelberg; 2009. p. 187-197.
208
(130) Speck RM, Courneya KS, Masse LC, Duval S, Schmitz KH. An update of
controlled physical activity trials in cancer survivors: a systematic review and metaanalysis. J Cancer Surviv 2010 Jun;4(2):87-100.
(131) Brawley LR, Culos-Reed S, Angove J, Hoffman-Goetz L. Understanding the
Barriers to Physical Activity for Cancer Patients: Review and Recommendations. J
Psychosoc Oncol 2002 12;20(4):1-21.
(132) DiStasio SA. Integrating yoga into cancer care. Cancer Care 2008;12(1):125-130.
(133) Hayat MJ. Understanding statistical significance. Nurs Res 2010 MayJun;59(3):219-223.
(134) Fethney J. Statistical and clinical significance, and how to use confidence intervals
to help interpret both. Aust Crit Care 2010 May;23(2):93-97.
(135) Sloan JA, Frost MH, Berzon R, Dueck A, Guyatt G, Moinpour C, et al. The clinical
significance of quality of life assessments in oncology: a summary for clinicians. Support
Care Cancer 2006 Oct;14(10):988-998.
(136) Kamath C, Dueck A. Incorporating Clinical Significance into a Study. Curr Probl
Cancer 2005 12;29(6):306-316.
(137) Wyrwich KW, Bullinger M, Aaronson N, Hays RD, Patrick DL, Symonds T, et al.
Estimating clinically significant differences in quality of life outcomes. Qual Life Res
2005 Mar;14(2):285-295.
209
(138) Wyrwich KW. Minimal important difference thresholds and the standard error of
measurement: is there a connection? J Biopharm Stat 2004 Feb;14(1):97-110.
(139) Sloan JA. Assessing the minimally clinically significant difference: scientific
considerations, challenges and solutions. COPD 2005 Mar;2(1):57-62.
(140) Norman GR, Sloan JA, Wyrwich KW. Interpretation of Changes in Health-Related
Quality of Life: The Remarkable Universality of Half a Standard Deviation. Med Care
2003 May;41(5):582-592.
(141) Citrome L. The Tyranny of the P-value: Effect Size Matters. Bulletin of Clinical
Psychopharmacology 2011;21(2):91-92.
(142) Cohen J. A power primer. Psychol Bull 1992 Jul;112(1):155-159.
(143) Kalinowski P, Fidler F. Interpreting Significance: The Differences Between
Statistical Significance, Effect Size, and Practical Importance. Newborn and Infant
Nursing Reviews 2010 3;10(1):50-54.
(144) Kirk R. W. Beyond the P: IV: Gain confidence in confidence intervals. Journal of
Cataract & Refractive Surgery 2004 12;30(12):2618-2619.
(145) Thompson B. Significance, Effect Sizes, Stepwise Methods, and Other Issues:
Strong Arguments Move the Field. The Journal of Experimental Education 2001
Fall;70(1):pp. 80-93.
(146) Biostat. Comprehensive Meta-analysis Version 2. 2005.
210
(147) Vadiraja HS, Raghavendra RM, Nagarathna R, Nagendra HR, Rekha M, Vanitha
N, et al. Effects of a yoga program on cortisol rhythm and mood states in early breast
cancer patients undergoing adjuvant radiotherapy: a randomized controlled trial. Integr
Cancer Ther 2009 Mar;8(1):37-46.
(148) Moadel AB, Shah C, Wylie-Rosett J, Harris MS, Patel SR, Hall CB, et al.
Randomized controlled trial of yoga among a multiethnic sample of breast cancer
patients: effects on quality of life. Journal of Clinical Oncology 2007;25(28):4387-4395.
(149) Culos-Reed SN, Carlson LE, Daroux LM, Hately-Aldous S. A pilot study of yoga
for breast cancer survivors: physical and psychological benefits. Psychooncology
2006;15(891-897).
(150) Littman AJ, Bertram LC, Ceballos R, Ulrich CM, Ramaprasad J, McGregor B, et
al. Randomized controlled pilot trial of yoga in overweight and obese breast cancer
survivors: effects on quality of life and anthropometric measures. Support Care Cancer
2011 Jan 5.
(151) Bower JE, Garet D, Sternlieb B. Yoga for persistent fatigue in breast cancer
survivors: results of a pilot study. Evid Based Complement Alternat Med
2011;2011:623168.
(152) Danhauer SC, Tooze JA, Farmer DF, Campbell CR, McQuellon RP, Barrett R, et
al. Restorative yoga for women with ovarian or breast cancer: findings from a pilot study.
J Soc Integr Oncol 2008 Spring;6(2):47-58.
211
(153) Duncan MD, Leis A, Taylor-Brown JW. Impact and outcomes of an iyengar yoga
program in a cancer centre. Curr Oncol 2008 Aug;15 Suppl 2:s109.es72-8.
(154) Galantino ML, Desai K, Greene L, Demichele A, Stricker CT, Mao JJ. Impact of
Yoga on Functional Outcomes in Breast Cancer Survivors With Aromatase InhibitorAssociated Arthralgias. Integr Cancer Ther 2011 Jul 6.
(155) Speed-Andrews AE, Stevinson C, Belanger LJ, Mirus JJ, Courneya KS. Pilot
evaluation of an Iyengar yoga program for breast cancer survivors. Cancer Nurs 2010
Sep-Oct;33(5):369-381.
(156) Ülger Ö, Yağlı NV. Effects of yoga on the quality of life in cancer patients.
Complementary Therapies in Clinical Practice 2010 5;16(2):60-63.
(157) Chandwani KD, Thornton B, Perkins GH, Arun B, Raghuram NV, Nagendra HR,
et al. Yoga improves quality of life and benefit finding in women undergoing
radiotherapy for breast cancer. J Soc Integr Oncol 2010 Spring;8(2):43-55.
(158) Cohen L, Warneke C, Fouladi RT, Rodriguez MA, Chaoul-Reich A. Psychological
adjustment and sleep quality in a randomized trial of the effects of a Tibetan yoga
intervention in patients with lymphoma. Cancer 2004;100(10):2253-2259.
(159) Revicki DA, Erickson PA, Sloan JA, Dueck A, Guess H, Santanello NC, et al.
Interpreting and reporting results based on patient-reported outcomes. Value Health 2007
Nov-Dec;10 Suppl 2:S116-24.
212
(160) Linden W, Satin JR. Avoidable pitfalls in behavioral medicine outcome research.
Ann Behav Med 2007 Apr;33(2):143-147.
(161) Stang A, Poole C, Kuss O. The ongoing tyranny of statistical significance testing in
biomedical research. Eur J Epidemiol 2010 Apr;25(4):225-230.
(162) Eton DT, Bauer BA, Sood A, Yost KJ, Sloan JA. Patient-Reported Outcomes in
Studies of Complementary and Alternative Medicine: Problems, Solutions, and Future
Directions. EXPLORE: The Journal of Science and Healing 2011 0;7(5):314-319.
(163) Garcia SF, Cella D, Clauser SB, Flynn KE, Lad T, Lai JS, et al. Standardizing
patient-reported outcomes assessment in cancer clinical trials: a patient-reported
outcomes measurement information system initiative. J Clin Oncol 2007 Nov
10;25(32):5106-5112.
(164) Miller AH, Ancoli-Israel S, Bower JE, Capuron L, Irwin MR. Neuroendocrineimmune mechanisms of behavioral comorbidities in patients with cancer. J Clin Oncol
2008 Feb 20;26(6):971-982.
(165) Mackenzie MJ, Wurz AJ, Culos-Reed SN. Evaluation of pre-post class affective
measures during a seven-week yoga for cancer survivors program. International Journal
of Yoga Therapy 2011;Supplement:S40.
(166) Lyubomirsky S, King L, Diener E. The Benefits of Frequent Positive Affect: Does
Happiness Lead to Success? Psychol Bull 2005 November;131(6):803-855.
213
(167) Rhodes RE, Fiala B, Conner M. A review and meta-analysis of affective judgments
and physical activity in adult populations. Ann Behav Med 2009 Dec;38(3):180-204.
(168) Ekkekakis P. Affect circumplex redux: the discussion on its utility as a
measurement framework in exercise psychology continues. International Review of Sport
& Exercise Psychology 2008 06;1(2):139-159.
(169) Kabat-Zinn J, Lipworth L, Burney R. The clinical use of mindfulness meditation
for the self-regulation of chronic pain. J Behav Med 1985 Jun;8(2):163-190.
(170) Jislin-Goldberg T, Tanay G, Bernstein A. Mindfulness and positive affect: Crosssectional, prospective intervention, and real-time relations. The Journal of Positive
Psychology 2012 09/01; 2012/09;7(5):349-361.
(171) Mackenzie M, Carlson L, Speca M. Mindfulness-based stress reduction (MBSR) in
oncology: rationale and review. Evidence-Based Integrative Medicine 2005;2(3):139145.
(172) Ledesma D, Kumano H. Mindfulness-based stress reduction and cancer: a metaanalysis. Psychooncology 2009 Jun;18(6):571-579.
(173) Matchim Y, Armer JM, Stewart BR. Mindfulness-based stress reduction among
breast cancer survivors: a literature review and discussion. Oncol Nurs Forum 2011
Mar;38(2):E61-71.
(174) Shennan C, Payne S, Fenlon D. What is the evidence for the use of mindfulnessbased interventions in cancer care? A review. Psychooncology 2011 Jul;20(7):681-697.
214
(175) Branstrom R, Kvillemo P, Brandberg Y, Moskowitz JT. Self-report mindfulness as
a mediator of psychological well-being in a stress reduction intervention for cancer
patients--a randomized study. Ann Behav Med 2010 May;39(2):151-161.
(176) Carlson L, Ursuliak Z, Goodey E, Angen M, Speca M. The effects of a mindfulness
meditation-based stress reduction program on mood and symptoms of stress in cancer
outpatients: 6-month follow-up. Support Care Cancer 2001;9(2):112-123.
(177) Carlson LE, Speca M, Faris P, Patel KD. One year pre-post intervention follow-up
of psychological, immune, endocrine and blood pressure outcomes of mindfulness-based
stress reduction (MBSR) in breast and prostate cancer outpatients. Brain Behav Immun
2007 Nov;21(8):1038-1049.
(178) Foley E, Baillie A, Huxter M, Price M, Sinclair E. Mindfulness-based cognitive
therapy for individuals whose lives have been affected by cancer: a randomized
controlled trial. J Consult Clin Psychol 2010 Feb;78(1):72-79.
(179) van der Lee ML, Garssen B. Mindfulness-based cognitive therapy reduces chronic
cancer-related fatigue: a treatment study. Psychooncology 2012 Mar;21(3):264-272.
(180) Baer RA, Smith GT, Hopkins J, Krietemeyer J, Toney L. Using Self-Report
Assessment Methods to Explore Facets of Mindfulness. Assessment 2006;13(1):27-45.
(181) Baer RA. Self-focused attention and mechanisms of change in mindfulness-based
treatment. Cogn Behav Ther 2009;38 Suppl 1:15-20.
215
(182) Branstrom R, Kvillemo P, Moskowitz JT. A Randomized Study of the Effects of
Mindfulness Training on Psychological Well-being and Symptoms of Stress in Patients
Treated for Cancer at 6-month Follow-up. Int J Behav Med 2011 Sep 20.
(183) Garland SN, Tamagawa R, Todd SC, Speca M, Carlson LE. Increased Mindfulness
Is Related to Improved Stress and Mood Following Participation in a Mindfulness-Based
Stress Reduction Program in Individuals With Cancer. Integr Cancer Ther 2012 Apr 13.
(184) Labelle LE, Campbell TS, Carlson LE. Mindfulness-Based Stress Reduction in
Oncology: Evaluating Mindfulness and Rumination as Mediators of Change in
Depressive Symptoms . Mindfulness 2010;1(1):28-40.
(185) Coffey KA, Hartman M, Fredrickson BL. Deconstructing Mindfulness and
Constructing Mental Health: Understanding Mindfulness and its Mechanisms of Action.
Mindfulness 2010;1(4):235-253.
(186) Ross Zahavich AN, Robinson JA, Paskevich D, Culos-Reed SN. Examining a
Therapeutic Yoga Program for Prostate Cancer Survivors. Integrative Cancer Therapies
2012 June 27.
(187) Sohl SJ, Schnur JB, Daly L, Suslov K, Montgomery GH. Development of the
Beliefs about Yoga Scale. International Journal of Yoga Therapy 2011;21(1):85-91.
(188) Hardy CJ, Rejeski JW. Not what, but how one feels: The measurement of affect
during exercise. Journal of Sport & Exercise Psychology 1989;11(3):304-317.
216
(189) Ekkekakis P, Petruzzello SJ. Acute aerobic exercise and affect: current status,
problems and prospects regarding dose-response. Sports Med 1999 Nov;28(5):337-374.
(190) Svebak S, Murgatroyd S. Metamotivational Dominance: A Multimethod Validation
of Reversal Theory Constructs. Journal of Personality & Social Psychology 1985
January;48(1):107-116.
(191) Tammen VV. Elite middle and long distance runners associative/dissociative
coping. Journal of Applied Sport Psychology 1996;8(1):1-8.
(192) Baden DA, Warwick-Evans L, Lakomy J. Am I Nearly There? The Effect of
Anticipated Running Distance on Oerceived Exertion and Attentional Focus. Journal of
Sport and Exercise Psychology 2004;26(2):215-231.
(193) Godin G, Shephard RJ. A simple method to assess exercise behavior in the
community. Can J Appl Sport Sci 1985 Sep;10(3):141-146.
(194) Thayer RE. The biopsychology of mood and arousal. New York , NY: Oxford
University Press; 1989.
(195) Shacham S. A Shortened Version of the Profile of Mood States. J Pers Assess 1983
06;47(3):305.
(196) Carlson LE, Thomas BC. Development of the Calgary Symptoms of Stress
Inventory (C-SOSI). Int J Behav Med 2007;14(4):249-256.
217
(197) Cella DF, Tulsky DS, Gray G, Sarafian B, Linn E, Bonomi A, et al. The Functional
Assessment of Cancer Therapy scale: development and validation of the general measure.
J Clin Oncol 1993 Mar;11(3):570-579.
(198) Hayes AF. A Primer on Multilevel Modelling. Human Communication Research
2006 10;32(4):385-410.
(199) Kahn JH. Multilevel modelling: overview and applications to research in
counseling psychology. J Couns Psychol 2011 Apr;58(2):257-271.
(200) Shek DT, Ma CM. Longitudinal data analyses using linear mixed models in SPSS:
concepts, procedures and illustrations. ScientificWorldJournal 2011 Jan 5;11:42-76.
(201) Nezlek JB. An Introduction to Multilevel Modelling for Social and Personality
Psychology. Social and Personality Psychology Compass 2008;2(2):842-860.
(202) Hox JJ. Multilevel analysis: techniques and applications. 2nd ed. New York, NY:
Routledge; 2010.
(203) Huh D, Flaherty B, Simoni J. Optimizing the Analysis of Adherence Interventions
Using Logistic Generalized Estimating Equations. AIDS and Behavior 2012 Feb
01(2):422-431.
(204) Birrer ,Daniel, Röthlin ,Philipp, Morgan ,Gareth. Mindfulness to Enhance Athletic
Performance: Theoretical Considerations and Possible Impact Mechanisms. Mindfulness
2012 Sep 01(3):235-246.
218
(205) Hölzel BK, Lazar SW, Gard T, Schuman-Olivier Z, Vago DR, Ott U. How Does
Mindfulness Meditation Work? Proposing Mechanisms of Action From a Conceptual and
Neural Perspective. Perspectives on Psychological Science 2011 November 01;6(6):537559.
(206) Jimenez SS, Niles BL, Park CL. A mindfulness model of affect regulation and
depressive symptoms: Positive emotions, mood regulation expectancies, and selfacceptance as regulatory mechanisms. Personality and Individual Differences 2010
10;49(6):645-650.
(207) Grabovac AD, Lau MA, Willett BR. Mechanisms of Mindfulness: A Buddhist
Psychological Model . Mindfulness 2011;2(3):154-166.
(208) Lutz A, Slagter HA, Dunne JD, Davidson RJ. Attention regulation and monitoring
in meditation. Trends Cogn Sci 2008 Apr;12(4):163-169.
(209) Brown KW, Ryan RM. The benefits of being present: Mindfulness and it's role in
psychological well-being. J Pers Soc Psychol 2003;84(4):822-848.
(210) Duijts SF, Faber MM, Oldenburg HS, van Beurden M, Aaronson NK.
Effectiveness of behavioral techniques and physical exercise on psychosocial functioning
and health-related quality of life in breast cancer patients and survivors--a meta-analysis.
Psychooncology 2011 Feb;20(2):115-126.
219
(211) Carlson LE, Campbell TS, Garland SN, Grossman P. Associations among salivary
cortisol, melatonin, catecholamines, sleep quality and stress in women with breast cancer
and healthy controls. J Behav Med 2007 Feb;30(1):45-58.
(212) Baer RA, Smith GT, Lykens E, Button D, Krietemeyer J, Sauer S, et al. Construct
Validity of the Five Facet Mindfulness Questionnaire in Meditating and Nonmeditating
Samples. Assessment 2008;15(3):329-342.
(213) Bicego D, Brown K, Ruddick M, Storey D, Wong C, Harris SR. Effects of exercise
on quality of life in women living with breast cancer: a systematic review. Breast J 2009
Jan-Feb;15(1):45-51.
(214) Bryan AD, Nilsson R, Tompkins SA, Magnan RE, Marcus BH, Hutchison KE. The
Big Picture of Individual Differences in Physical Activity Behavior Change: A
Transdisciplinary Approach. Psychol Sport Exerc 2011 Jan;12(1):20-26.
(215) Hallgren MÅ, Moss ND, Gastin P. Regular exercise participation mediates the
affective response to acute bouts of avigorous exercise. Journal of Sports Science &
Medicine 2010 12;9(4):629-637.
(216) Josefsson ,Torbjörn, Larsman ,Pernilla, Broberg ,Anders, Lundh ,Lars-Gunnar.
Self-Reported Mindfulness Mediates the Relation Between Meditation Experience and
Psychological Well-Being. Mindfulness 2011 Mar 01(1):49-58.
220
(217) Schroevers MJ, Brandsma R. Is learning mindfulness associated with improved
affect after mindfulness-based cognitive therapy? Br J Psychol 2010 Feb;101(Pt 1):95107.
(218) Speed-Andrews AE, Stevinson C, Belanger LJ, Mirus JJ, Courneya KS. Predictors
of adherence to an Iyengar yoga program in breast cancer survivors. Int J Yoga 2012
Jan;5(1):3-9.
(219) Bryan S, Pinto Zipp G, Parasher R. The effects of yoga on psychosocial variables
and exercise adherence: a randomized, controlled pilot study. Altern Ther Health Med
2012 Sep-Oct;18(5):50-59.
(220) Ulmer CS, Stetson BA, Salmon PG. Mindfulness and acceptance are associated
with exercise maintenance in YMCA exercisers. Behav Res Ther 2010 Aug;48(8):805809.
(221) Ross A, Friedmann E, Bevans M, Thomas S. Frequency of yoga practice predicts
health: results of a national survey of yoga practitioners. Evid Based Complement
Alternat Med 2012;2012:983258.
(222) Stange KC, Breslau ES, Dietrich AJ, Glasgow RE. State-of-the-art and future
directions in multilevel interventions across the cancer control continuum. J Natl Cancer
Inst Monogr 2012 May;2012(44):20-31.
(223) Khalsa SB. Yoga as a therapeutic intervention: a bibliometric analysis of published
research studies. Indian J Physiol Pharmacol 2004 Jul;48(3):269-285.
221
(224) Takahashi T, Murata T, Hamada T, Omori M, Kosaka H, Kikuchi M, et al.
Changes in EEG and autonomic nervous activity during meditation and their association
with personality traits. Int J Psychophysiol 2005 Feb;55(2):199-207.
(225) Achten J, Jeukendrup AE. Heart rate monitoring: applications and limitations.
Sports Med 2003;33(7):517-538.
(226) Ditto B, Eclache M, Goldman N. Short-term autonomic and cardiovascular effects
of mindfulness body scan meditation. Ann Behav Med 2006 Dec;32(3):227-234.
(227) Houtveen JH, Rietveld S, de Geus EJ. Contribution of tonic vagal modulation of
heart rate, central respiratory drive, respiratory depth, and respiratory frequency to
respiratory sinus arrhythmia during mental stress and physical exercise.
Psychophysiology 2002 Jul;39(4):427-436.
(228) Giese-Davis J, Wilhelm FH, Conrad A, Abercrombie HC, Sephton S, Yutsis M, et
al. Depression and stress reactivity in metastatic breast cancer. Psychosom Med 2006
Sep-Oct;68(5):675-683.
(229) Mouton C, Ronson A, Razavi D, Delhaye F, Kupper N, Paesmans M, et al. The
relationship between heart rate variability and time-course of carcinoembryonic antigen
in colorectal cancer. Autonomic Neuroscience 2012 1/26;166(1–2):96-99.
(230) Fagundes CP, Murray DM, Hwang BS, Gouin J, Thayer JF, Sollers III JJ, et al.
Sympathetic and parasympathetic activity in cancer-related fatigue: More evidence for a
222
physiological substrate in cancer survivors. Psychoneuroendocrinology 2011;38(8):137147.
(231) Morrow GR, Hickok JT, DuBeshter B, Lipshultz SE. Changes in clinical measures
of autonomic nervous system function related to cancer chemotherapy-induced nausea. J
Auton Nerv Syst 1999 10/8;78(1):57-63.
(232) Chauhan A, Sequeria A, Manderson C, Maddocks M, Wasley D, Wilcock A.
Exploring autonomic nervous system dysfunction in patients with cancer cachexia: A
pilot study. Autonomic Neuroscience 2012 1/26;166(1–2):93-95.
(233) Fadul N, Strasser F, Palmer JL, Yusuf SW, Guo Y, Li Z, et al. The association
between autonomic dysfunction and survival in male patients with advanced cancer: a
preliminary report. J Pain Symptom Manage 2010 Feb;39(2):283-290.
(234) Stone CA, Kenny RA, Nolan B, Lawlor PG. Autonomic dysfunction in patients
with advanced cancer; prevalence, clinical correlates and challenges in assessment. BMC
Palliat Care 2012 Mar 1;11:3.
(235) Kim DH, Kim J,A., Choi YS, Kim SH, Lee JY, Kim YE. Heart Rate Variability
and Length of Survival in Hospice Cancer Patients. J Korean Med Sci 2010
/8/;25(8):1140-1145.
(236) Niederer D, Vogt L, Thiel C, Schmidt K, Bernhorster M, Lungwitz A, et al.
Exercise Effects on HRV in Cancer Patients. Int J Sports Med 2012 Aug 15.
223
(237) Sunkaria RK, Kumar V, Saxena SC. A comparative study on spectral parameters of
HRV in yogic and non-yogic practitioners. International Journal of Medical Engineering
and Informatics 2010 01/01;2(1):1-14.
(238) Routledge FS, Campbell TS, McFetridge-Durdle JA, Bacon SL. Improvements in
heart rate variability with exercise therapy. Can J Cardiol 2010 Jun-Jul;26(6):303-312.
(239) Varela F. Neurophenomenology: A Methodological Remedy for the Hard Problem.
Journal of Consciousness Studies 1996;3(4):330-349.
(240) Lutz A. Neurophenomenology and the study of self-consciousness. Conscious
Cogn 2007 Sep;16(3):765-767.
(241) Faul F, Erdfelder E, Lang AG, Buchner A. G*Power 3: a flexible statistical power
analysis program for the social, behavioral, and biomedical sciences. Behav Res Methods
2007 May;39(2):175-191.
(242) Ekkekakis P, Backhouse SH, Gray C, Lind E. Walking is popular among adults but
is it pleasant? A framework for clarifying the link between walking and affect as
illustrated in two studies. Psychol Sport Exerc 2008 5;9(3):246-264.
(243) Rose EA, Parfitt G. A quantitative analysis and qualitative explanation of the
individual differences in affective responses to prescribed and self-selected exercise
intensities. J Sport Exerc Psychol 2007 Jun;29(3):281-309.
224
(244) Welch AS, Hulley A, Ferguson C, Beauchamp MR. Affective responses of inactive
women to a maximal incremental exercise test: A test of the dual-mode model. Psychol
Sport Exerc 2007 7;8(4):401-423.
(245) Hansen AL, Johnsen BH, Sollers JJ,3rd, Stenvik K, Thayer JF. Heart rate
variability and its relation to prefrontal cognitive function: the effects of training and
detraining. Eur J Appl Physiol 2004 Dec;93(3):263-272.
(246) Wu SD, Lo PC. Inward-attention meditation increases parasympathetic activity: a
study based on heart rate variability. Biomed Res 2008 Oct;29(5):245-250.
(247) Pober DM, Braun B, Freedson PS. Effects of a single bout of exercise on resting
heart rate variability. Med Sci Sports Exerc 2004 Jul;36(7):1140-1148.
(248) Sandercock GR, Brodie DA. The use of heart rate variability measures to assess
autonomic control during exercise. Scand J Med Sci Sports 2006 Oct;16(5):302-313.
(249) Braun V, Clarke V. Using thematic analysis in psychology. Qualitative Research in
Psychology 2006 04;3(2):77-101.
(250) Brennan C, Stevens J. A grounded theory approach towards understanding the self
perceived effects of meditation on people being treated for cancer. Aust J Holist Nurs
1998 Oct;5(2):20-26.
(251) Borg G. Borg's Perceived Exertion and Pain Scales. Champaign, IL: Human
Kinetics Books; 1998.
225
(252) Kop WJ, Stein PK, Tracy RP, Barzilay JI, Schulz R, Gottdiener JS. Autonomic
Nervous System Dysfunction and Inflammation Contribute to the Increased
Cardiovascular Mortality Risk Associated With Depression. Psychosomatic Medicine
2010(7):626-635.
(253) Nezlek JB. Multilevel modelling for psychologists. In: Cooper H, Camic PM, Long
DL, Panter AT, Rindskopf D, Sher KJ, editors. APA handbook of research methods in
psychology, Vol 3: Data analysis and research publication Washington, DC: American
Psychological Association; 2012. p. 219-241.
(254) Rejas J, Pardo A, Ruiz MA. Standard error of measurement as a valid alternative to
minimally important difference for evaluating the magnitude of changes in patientreported outcomes measures. J Clin Epidemiol 2008 Apr;61(4):350-356.
(255) Rose EA, Parfitt G. Pleasant for some and unpleasant for others: a protocol analysis
of the cognitive factors that influence affective responses to exercise. Int J Behav Nutr
Phys Act 2010 Feb 7;7:15.
(256) Bryman A. Integrating quantitative and qualitative research: how is it done? 2006
2006;6(1):97-113.
(257) Moran-Ellis J, Alexander VD, Cronin A, Dickinson M, Fielding J, Sleney J, et al.
Triangulation and integration: processes, claims and implications. Qualitative Research
2006 February 01;6(1):45-59.
226
(258) Creswell JW, Klassen AC, Plano Clark VL, Smith KC. Best practices for mixed
methods research in the health sciences. Office of Behavioral and Social Sciences
Research: National Institutes of Health August 2011.
(259) Preacher KJ, Curran PJ, Bauer DJ. Computational Tools for Probing Interactions in
Multiple Linear Regression, Multilevel Modelling, and Latent Curve Analysis. Journal of
Educational and Behavioral Statistics 2006 December 21;31(4):437-448.
(260) Telles S, Singh N, Balkrishna A. Heart rate variability changes during high
frequency yoga breathing and breath awareness. Biopsychosoc Med 2011 Apr 13;5:4.
(261) Tanaka H, Monahan KD, Seals DR. Age-predicted maximal heart rate revisited. J
Am Coll Cardiol 2001 Jan;37(1):153-156.
(262) Koelsch S, Remppis A, Sammler D, Jentschke S, Mietchen D, Fritz T, et al. A
cardiac signature of emotionality. Eur J Neurosci 2007 12;26(11):3328-3338.
(263) Improvement of HRV methodology for positive/negative emotion assessment.
Collaborative Computing: Networking, Applications and Worksharing, 2009.
CollaborateCom 2009. 5th International Conference on; 2009.
(264) Ekkekakis P, Lind E, Vazou S. Affective responses to increasing levels of exercise
intensity in normal-weight, overweight, and obese middle-aged women. Obesity (Silver
Spring) 2010 Jan;18(1):79-85.
(265) Hagemann D, Waldstein SR, Thayer JF. Central and autonomic nervous system
integration in emotion. Brain Cogn 2003 6;52(1):79-87.
227
(266) Hansen AL, Johnsen BH, Thayer JF. Vagal influence on working memory and
attention. Int J Psychophysiol 2003 Jun;48(3):263-274.
(267) Thayer JF, Friedman BH. Stop that! Inhibition, sensitization, and their
neurovisceral concomitants. Scand J Psychol 2002 04;43(2):123.
(268) Buchheit M, Al Haddad H, Laursen PB, Ahmaidi S. Effect of body posture on
postexercise parasympathetic reactivation in men. Exp Physiol 2009;94(7):795-804.
(269) Telles S, Naveen KV. Voluntary Breath Regulation in Yoga: Its Relevance and
Physiological Effects. Biofeedback 2008 Summer2008;36(2):70-73.
(270) Porto LG, Junqueira LF,Jr. Comparison of time-domain short-term heart interval
variability analysis using a wrist-worn heart rate monitor and the conventional
electrocardiogram. Pacing Clin Electrophysiol 2009 Jan;32(1):43-51.
(271) Nunan D, Jakovljevic DG, Donovan G, Hodges LD, Sandercock GR, Brodie DA.
Levels of agreement for RR intervals and short-term heart rate variability obtained from
the Polar S810 and an alternative system. Eur J Appl Physiol 2008 Jul;103(5):529-537.
(272) Wallen MB, Hasson D, Theorell T, Canlon B, Osika W. Possibilities and
limitations of the Polar RS800 in measuring heart rate variability at rest. Eur J Appl
Physiol 2012 Mar;112(3):1153-1165.
(273) Quintana ,Daniel, Heathers ,James, Kemp ,Andrew. On the validity of using the
Polar RS800 heart rate monitor for heart rate variability research. Eur J Appl Physiol
2012;Published Online.
228
(274) Maas CJM, Hox JJ. Sufficient Sample Sizes for Multilevel Modelling.
Methodology: European Journal of Research Methods for the Behavioral and Social
Sciences 2005 01/01;1(3):86-92.
(275) Parfitt G, Hughes S. The Exercise Intensity–Affect Relationship: Evidence and
Implications for Exercise Behavior. Journal of Exercise Science & Fitness 2009;7(2,
Supplement):S34-S41.
(276) Masters KS, Ogles BM. Associative and dissociative cognitive strategies in
exercise and running: 20 years later, what do we know? / Strategies cognitives
associatives et dissociatives dans l'exercice et la course: 20 ans plus tard, qu'en est il ?
Sport Psychologist 1998 09;12(3):253-270.
(277) Katsanos CS, Moffatt RJ. Reliability of heart rate responses at given ratings of
perceived exertion in cycling and walking. Res Q Exerc Sport 2005 Dec;76(4):433-439.
(278) Peng CK, Henry IC, Mietus JE, Hausdorff JM, Khalsa G, Benson H, et al. Heart
rate dynamics during three forms of meditation. Int J Cardiol 2004 May;95(1):19-27.
(279) Clay CC, Lloyd LK, Walker JL, Sharp KR, Pankey RB. The metabolic cost of
hatha yoga. J Strength Cond Res 2005;19(3):604-610.
(280) Hagins M, Moore W, Rundle A. Does practicing hatha yoga satisfy
recommendations for intensity of physical activity which improves and maintains health
and cardiovascular fitness? BMC Complementary and Alternative Medicine
2007;7(1):40.
229
(281) Bharshankar JR, Bharshankar RN, Deshpande VN, Kaore SB, Gosavi GB. Effect
of yoga on cardiovascular system in subjects above 40 years. Indian J Physiol Pharmacol
2003 Apr;47(2):202-206.
(282) Devasena I, Narhare P. Effect of yoga on heart rate and blood pressure and its
clinical significance. International journal of Biological & Medical Research
2011;2(3):750-753.
(283) Madanmohan, Udupa K, Bhavanani AB, Shatapathy CC, Sahai A. Modulation of
cardiovascular response to exercise by yoga training. Indian J Physiol Pharmacol 2004
Oct;48(4):461-465.
(284) Abel AN, Lloyd LK, Williams JS. The Effects of Regular Yoga Practice on
Pulmonary Function in Healthy Individuals: A Literature Review. J Altern Complement
Med 2012 Sep 14.
(285) Jayasinghe SR. Yoga in cardiac health (a review). Eur J Cardiovasc Prev Rehabil
2004 Oct;11(5):369-375.
(286) Jerath R, Edry JW, Barnes VA, Jerath V. Physiology of long pranayamic breathing:
neural respiratory elements may provide a mechanism that explains how slow deep
breathing shifts the autonomic nervous system. Med Hypotheses 2006;67(3):566-571.
(287) Benson H. The Relaxation Response. New York, NY: HarperCollins; 1975.
(288) Dusek JA, Benson H. Mind-body medicine: a model of the comparative clinical
impact of the acute stress and relaxation responses. Minn Med 2009 May;92(5):47-50.
230
(289) Esch T, Fricchione GL, Stefano GB. The therapeutic use of the relaxation response
in stress-related diseases. Med Sci Monit 2003 Feb;9(2):RA23-34.
(290) Busch V, Magerl W, Kern U, Haas J, Hajak G, Eichhammer P. The effect of deep
and slow breathing on pain perception, autonomic activity, and mood processing--an
experimental study. Pain Med 2012 Feb;13(2):215-228.
231
APPENDICES
232
Title:
Exploring the relationship between yoga practice, affect and
attention regulation, health outcomes and program adherence in
cancer survivors.
Investigators:
Dr. S Nicole Culos-Reed (Principal Investigator)
Michael Mackenzie (Co-Investigator)
Sponsor:
None
This consent form, a copy of which has been given to you, is only part of the
process of informed consent. It should give you the basic idea of what the
research is about and what your participation will involve. If you would like more
detail about something mentioned here, or information not included here, you
should feel free to ask. Please take the time to read this carefully and to
understand any accompanying information.
Background
Yoga appears to be a promising complementary exercise choice for cancer survivors.
Within cancer settings, positive effects have been seen on indices of mood, stress-related
symptoms, overall quality of life and sleep, as well as functional and physiological
measures. Despite preliminary findings, little attention has been paid to the mechanisms
by which benefits are attained. Examining these mechanisms for intervention effects will
enable us to understand not only if yoga works, but also how. The proposed study
addresses this research gap and provides a foundation to more fully explore the
applications and benefits of yoga for cancer survivors and support persons.
What is the purpose of the study?
The purpose of this study is to investigate the psychological benefits of a therapeutic
yoga program for cancer survivors and support persons. Specifically we are interested in
psychological benefits to mood, stress, and quality of life.
What would I have to do?
Participants in the Yoga Thrive program will be asked to complete pre and post
questionnaires in the week prior to and immediately following the yoga program, and
complete the questionnaire portion at three and six-months following the program (i.e., to
be completed online or mailed to you with a pre-paid envelope for return). The
questionnaire assesses current physical activity beliefs and behaviour, mood, stress and
quality of life. Completion of the questionnaires pre and post intervention and at the three
and six month follow-ups should take an hour each time for a total of four hours.
In addition, participants will be asked to provide feedback on their yoga experience via
focus groups. These one hours focus group sessions will be scheduled following
completion of the last yoga class and at the three and six month follow-ups for a total of
three hours.
233
What are the risks?
There are no known physical or psychological risks associated with completing the
questionnaires, but you may discontinue at any time if you choose to do so.
Will I benefit if I take part?
If you agree to participate in this study, there may or may not be a direct benefit to you.
Your quality of life may be improved during the study, but there is no guarantee this
research will help you. The information we get from this study may help us to provide
better and more accessible activity options for cancer survivors and support persons in
the future.
Do I have to participate?
Participation in this study is completely voluntary and you may withdraw at any time
without jeopardizing your health care. Please inform the research coordinator or the
principal investigator if you choose to withdraw from the study.
What else does my participation involve?
This study involves no further participation beyond the six-month follow-up mailed
questionnaire.
Will I be paid for participating, or do I have to pay for anything?
There are no costs associated with participating in this research study (questionnaire and
focus groups). The Yoga Thrive program all participants will be recruited from is an
ongoing community-based program offering fee-based yoga classes for cancer survivors
and support persons.
Will my records be kept private?
The information obtained in this study will be stored on a secure (password protected)
computer, and the written assessments will be stored in a secure and locked file cabinet.
The information will be kept in Dr. Culos-Reed’s office in the Kinesiology Complex. All
records will be coded with non-descriptive ID numbers and kept in the strictest
confidence by the principal investigator and individual results will not be disclosed to
anyone other than the participant or the University of Calgary Conjoint Health Research
Ethics Board.
If I suffer a research-related injury, will I be compensated?
In the event that you suffer injury as a result of participating in this research, no
compensation will be provided to you by the University of Calgary, or the Researchers.
You still have your legal rights. Nothing said in this consent form alters your right to seek
damages.
Consent
Your signature on this form indicates that you have understood to your
satisfaction the information regarding participation in the research project and
agree to participate as a subject. In no way does this waive your legal rights nor
release the investigators, or involved institutions from their legal and professional
234
responsibilities. If you have further questions concerning matters related to this
research, please contact:
Dr. Nicole Culos-Reed, Principal Investigator (403) 220-7540
or
Michael Mackenzie, Co-investigator and Research Coordinator (403) 210-8482
If you have any questions concerning your rights as a possible participant in this
research, please contact the Director, The Office of Medical Bioethics, University
of Calgary, at (403) 220-7990.
(name of participant)
(date)
(signature of participant)
(name of investigator)
(date)
(signature of investigator)
(name of witness)
(date)
(signature of witness)
A copy of this consent form has been given to you to keep for your records and
reference.
The University of Calgary Conjoint Health Research Ethics Board has approved this
research study (Ethics ID 23313).
235
Title:
Exploring the relationships between a single-session yoga practice
and self-regulation in cancer survivors.
Investigators:
Dr. S Nicole Culos-Reed (Principal Investigator)
Michael Mackenzie (Co-Investigator)
Sponsor:
None
This consent form, a copy of which has been given to you, is only part of the
process of informed consent. It should give you the basic idea of what the
research is about and what your participation will involve. If you would like more
detail about something mentioned here, or information not included here, you
should feel free to ask. Please take the time to read this carefully and to
understand any accompanying information.
Background
Yoga appears to be a promising complementary exercise choice for cancer survivors.
Within cancer settings, positive effects have been seen on indices of mood, stress-related
symptoms, overall quality of life and sleep, as well as functional and physiological
measures. Despite preliminary findings, little attention has been paid to the mechanisms
by which benefits are attained. Examining these mechanisms for intervention effects will
enable us to understand not only if yoga works, but also how. The proposed study
addresses this research gap and provides a foundation to more fully explore the
applications and benefits of yoga for both cancer survivors and support persons.
What is the purpose of the study?
The purpose of the present study is to examine the psychological effects of a single
session of yoga both during and after a yoga practice.
What would I have to do?
Your participation in this study will involve attending one 75 minute yoga session. The
session will be offered at either the University of Calgary, the Holy Cross Hospital Gym
or at WellSpring Calgary. The session will include postures, deep relaxation, breathing
practices and meditation. The session will be led by a certified yoga instructor with
cancer-specific training. You will complete brief psychological measures at the
beginning, during, and end of the of the yoga session. Physiological measures (Heart Rate
and Heart Rate Variability) will be assessed throughout the class via a wireless
ambulatory chest-band heart rate monitor. The yoga session will be conducted during the
morning to ensure physiological similar states for all participants. In addition, you will be
asked not to exercise or consume caffeine or alcoholic beverages 12 hours prior to the
experiment and to refrain from eating and drinking 3 hours prior to the experiment. The
entire session, including the yoga practice will take 2.5 hours.
236
What are the risks?
The practice of yoga as outlined in this consent form is gentle, should not cause you pain
or injury and is therefore considered to be of minimal risk. Nonetheless, there are risks
associated with involvement in any physical activity program. Some people have also
noticed some initial discomfort associated with the chest-band heart rate monitor. There
are no other known physical or psychological risks associated with involvement in the
program, but you may discontinue any activity at any time if you choose to do so.
Will I benefit if I take part?
If you agree to participate in this study, there may or may not be a direct benefit to you.
Your quality of life may be improved during the study, but there is no guarantee this
research will help you. The information we get from this study may help us to provide
better and more accessible activity options for cancer survivors and support persons in
the future.
Do I have to participate?
Participation in this study is completely voluntary and you may withdraw at any time
without jeopardizing your health care. Please inform the research coordinator or the
principal investigator if you choose to withdraw from the study.
What else does my participation involve?
This study involves no further active participation beyond the single yoga session, usage
of a wireless heart rate monitor and associated questionnaires previously described.
Will I be paid for participating, or do I have to pay for anything?
There will be no costs to you for participating in this study.
Will my records be kept private?
The information obtained in this study will be stored on a secure (password protected)
computer, and the written assessments will be stored in a secure and locked file cabinet.
The information will be kept in Dr. Culos-Reed’s office in the Kinesiology Complex. All
records will be coded with non-descriptive ID numbers and kept in the strictest
confidence by the principal investigator and individual results will not be disclosed to
anyone other than the participant or the University of Calgary Conjoint Health Research
Ethics Board.
If I suffer a research-related injury, will I be compensated?
In the event that you suffer injury as a result of participating in this research, no
compensation will be provided to you by the University of Calgary, or the Researchers.
You still have your legal rights. Nothing said in this consent form alters your right to seek
damages.
Signatures
Your signature on this form indicates that you have understood to your
satisfaction the information regarding participation in the research project and
agree to participate as a subject. In no way does this waive your legal rights nor
237
release the investigators, or involved institutions from their legal and professional
responsibilities. If you have further questions concerning matters related to this
research, please contact:
Dr. Nicole Culos-Reed, Principal Investigator (403) 220-7540
or
Michael Mackenzie, Co-investigator and Research Coordinator (403) 210-8482
If you have any questions concerning your rights as a possible participant in this
research, please contact the Director, The Office of Medical Bioethics, University
of Calgary, at (403) 220-7990.
Participant’s Name
Signature and Date
Investigator/Delegate’s Name
Signature and Date
Witness’ Name
Signature and Date
The University of Calgary Conjoint Health Research Ethics Board has approved this
research study (Ethics ID 23313).
A signed copy of this consent form has been given to you to keep for your records and
reference.
238
DEMOGRAPHICS (Please answer in relation to your CURRENT status.):
Please check only one option. If you do not wish to answer a question, please skip and go to the
next question. All information provided will be kept anonymous and confidential.
1. Marital status:
Married/ common law:
Divorced/ separated:
Widowed:
Never Married:




2. Education Level (please check highest level attained):
Some high school:

Completed high school:

Some university/ College:

Completed University/ College:

Some OR completed Grad school:

3. Annual Family Income:
<20,000:

20,000-39,999:

40,000-59,999:

60,000-79,999:

>80,000:

4. Employment Status:
Full-time:
Homemaker:
Retired:
Part-time:
Unemployed:
Disability/ sick leave:
5. Gender:
Male:
Female:


6. Age: ___________






239
MEDICAL INFORMATION
Cancer Diagnosis (type and stage): _________________________________________________
Date of Diagnosis (approximate): __________________________________________________
Month / Year
Type of Treatment:
Radiation therapy
 
Chemotherapy

Surgery
 

Other (please specify):___________________________________________________________
If applicable, approximate date you finished active treatment (i.e., radiation, chemo):
_____________________________________________________________________________
Month / Year
240
PREVIOUS YOGA EXPERIENCE
Have you previously participated in yoga (i.e., class, DVD, tape, or book)?
Never: 

I tried it once or twice:  
I used to participate regularly:  If yes, how often? __________________________________
I currently participate regularly:  If yes, how often? ________________________________ 
If you have any previous yoga experience, what was the format?
Instructed class or private lesson(s):  
DVD or video:  
CD or cassette tape:  
Book:  
Other (please specify):___________________________________________________________
YOGA THRIVE SUPPORT PERSON
Do you have a support person who will be attending Yoga Thrive with you?
Yes:
 
No:
 
241
Beliefs about Yoga
Likely
Neither
Unlikely
Very
Unlikely
If I practiced yoga…
Extremely
Unlikely
INSTRUCTIONS: For each of these statements, please indicate the extent to which you believe
the statement is true by circling the appropriate number.
1. I would become more flexible.
1
2
3
4
5
6
7
2. There would only be “new age” people
in a class.
1
2
3
4
5
6
7
3. It would help me focus.
1
2
3
4
5
6
7
4. It would help me gain self-awareness.
1
2
3
4
5
6
7
5. I would be embarrassed in a class.
1
2
3
4
5
6
7
6. I would have to be more flexible to take
a class.
1
2
3
4
5
6
7
7. It would improve my overall health.
1
2
3
4
5
6
7
8. There would only be women in a class.
1
2
3
4
5
6
7
9. I wouldn’t be good at it.
1
2
3
4
5
6
7
10. The instructor would make me
uncomfortable
1
2
3
4
5
6
7
11. I would sleep better.
1
2
3
4
5
6
7
242
Feeling Scale
Estimate how good or bad you feel right now.
-5
-4
Very
Bad
-3
-2
Bad
-1
0
+1
+2
Fairly Neutral Fairly
Bad
Good
+3
+4
Good
+5
Very
Good
Felt Arousal Scale
Estimate how aroused you feel right now (low arousal = calm or fatigued, high arousal =
anxious or energized).
1
2
3
4
5
6
Low
Arousal
High
Arousal
Associative & Dissociative Thoughts
Estimate ratio of associative to dissociative thoughts right now (associative = focused in
present moment, dissociative = unfocused in past or future).
1
Very
Associative
2
3
4
5
6
7
8
9
10
Very
Dissociative
243
Godin Leisure Time Exercise Questionnaire
We would like you to recall your average weekly exercise over the past month. How many times
per week on average did you do the following kinds of exercise over the past month?
When answering these questions please remember to:

Consider your average weekly exercise over the past month

Only count exercise sessions that lasted 15 minutes or longer in duration

Only count exercise that was done during free time (i.e., do not included occupation or
housework)

Note the main difference between the three categories is the intensity of the exercise

Write the average frequency on the first line and the average duration on the second
line
A. STRENUOUS EXERCISE (Heart beats rapidly, sweating)
(e.g., running, jogging, hockey, soccer, squash, cross country skiing, judo, roller skating, vigorous
swimming, vigorous long distance bicycling, vigorous aerobic dance classes, heavy weight
training)
In an average week I was involved in strenuous exercise __________ times/week for an average
duration of __________ minutes/each session.
B. MODERATE EXERCISE (Not exhausting, light perspiration)
(e.g., fast walking, baseball, tennis, easy bicycling, volleyball, badminton, easy swimming, alpine
skiing, popular and folk dancing)
In an average week I was involved in moderate exercise __________ times/week for an average
duration of __________ minutes/each session.
C. MILD EXERCISE (Minimal effort, no perspiration)
(e.g., easy walking, yoga, archery, fishing, bowling, lawn bowling, shuffleboard, horseshoes, golf,
snowmobiling)
In an average week I was involved in mild exercise __________ times/week for an average
duration of __________ minutes/each session.
244
The Activation-Deactivation Adjective Check List (AD ACL) Short Form
Each of the words on the back describes feelings or mood. Please use the rating scale
next to each word to describe your feelings at this moment.
Work rapidly, but please mark all the words. Your first reaction is best. This should take
only a minute or two.
Active
Placid
Sleepy
Jittery
Energetic
Intense
Calm
Tired
Vigorous
At-rest
Drowsy
Fearful
Lively
Still
Wide-awake
Clutched-up
Quiet
Full-of-pep
Tense
Wakeful
Definitely feel Feel slightly Cannot decide Definitely do not feel
245
Five Facet Mindfulness Questionnaire
Please rate each of the following statements using the scale provided. Write the number
in the blank that best describes your own opinion of what is generally true for you.
1
never or very
rarely true
2
rarely
true
3
sometimes
true
often
true
4
5
very often or
always true
_____ 1. When I’m walking, I deliberately notice the sensations of my body moving.
_____ 2. I’m good at finding words to describe my feelings.
_____ 3. I criticize myself for having irrational or inappropriate emotions.
_____ 4. I perceive my feelings and emotions without having to react to them.
_____ 5. When I do things, my mind wanders off and I’m easily distracted.
_____ 6. When I take a shower or bath, I stay alert to the sensations of water on my body.
_____ 7. I can easily put my beliefs, opinions, and expectations into words.
_____ 8. I don’t pay attention to what I’m doing because I’m daydreaming, worrying, or
otherwise distracted.
_____ 9. I watch my feelings without getting lost in them.
_____ 10. I tell myself I shouldn’t be feeling the way I’m feeling.
_____ 11. I notice how foods and drinks affect my thoughts, bodily sensations, and emotions.
_____ 12. It’s hard for me to find the words to describe what I’m thinking.
_____ 13. I am easily distracted.
_____ 14. I believe some of my thoughts are abnormal or bad and I shouldn’t think that way.
_____ 15. I pay attention to sensations, such as the wind in my hair or sun on my face.
_____ 16. I have trouble thinking of the right words to express how I feel about things
_____ 17. I make judgments about whether my thoughts are good or bad.
_____ 18. I find it difficult to stay focused on what’s happening in the present.
_____ 19. When I have distressing thoughts or images, I “step back” and am aware of the thought
or image without getting taken over by it.
246
1
never or very
rarely true
2
rarely
true
3
sometimes
true
4
often
true
5
very often or
always true
_____ 20. I pay attention to sounds, such as clocks ticking, birds chirping, or cars passing.
_____ 21. In difficult situations, I can pause without immediately reacting.
_____ 22. When I have a sensation in my body, it’s difficult for me to describe it because I can’t
find the right words.
_____ 23. It seems I am “running on automatic” without much awareness of what I’m doing.
_____24. When I have distressing thoughts or images, I feel calm soon after.
_____ 25. I tell myself that I shouldn’t be thinking the way I’m thinking.
_____ 26. I notice the smells and aromas of things.
_____ 27. Even when I’m feeling terribly upset, I can find a way to put it into words.
_____ 28. I rush through activities without being really attentive to them.
_____ 29. When I have distressing thoughts or images I am able just to notice them without
reacting.
_____ 30. I think some of my emotions are bad or inappropriate and I shouldn’t feel them.
_____ 31. I notice visual elements in art or nature, such as colors, shapes, textures, or patterns of
light and shadow.
_____ 32. My natural tendency is to put my experiences into words.
_____ 33. When I have distressing thoughts or images, I just notice them and let them go.
_____ 34. I do jobs or tasks automatically without being aware of what I’m doing.
_____ 35. When I have distressing thoughts or images, I judge myself as good or bad, depending
what the thought/image is about.
_____ 36. I pay attention to how my emotions affect my thoughts and behavior.
_____ 37. I can usually describe how I feel at the moment in considerable detail.
_____ 38. I find myself doing things without paying attention.
_____ 39. I disapprove of myself when I have irrational ideas.
247
POMS-SF
Below is a list of words that describe feelings people have. Please read each one carefully. Then
circle ONE answer to the right, which describes HOW YOU HAVE BEEN FEELING OVER THE PAST
WEEK INCLUDING TODAY.
Not at all
A Little
Moderately
Quite a
bit
Extremely
1. Tense
0
1
2
3
4
2. Angry
0
1
2
3
4
3. Worn out
0
1
2
3
4
4. Unhappy
0
1
2
3
4
5. Lively
0
1
2
3
4
6. Confused
0
1
2
3
4
7. Peeved
0
1
2
3
4
8. Sad
0
1
2
3
4
9. Active
0
1
2
3
4
10. On edge
0
1
2
3
4
11. Grouchy
0
1
2
3
4
12. Blue
0
1
2
3
4
13. Energetic
0
1
2
3
4
14. Hopeless
0
1
2
3
4
15. Uneasy
0
1
2
3
4
16. Restless
0
1
2
3
4
17. Unable to concentrate
0
1
2
3
4
248
Not at all
A Little
Moderately
Quite a
bit
Extremely
18. Fatigued
0
1
2
3
4
19. Annoyed
0
1
2
3
4
20. Discouraged
0
1
2
3
4
21. Resentful
0
1
2
3
4
22. Nervous
0
1
2
3
4
23. Miserable
0
1
2
3
4
24. Cheerful
0
1
2
3
4
25. Bitter
0
1
2
3
4
26. Exhausted
0
1
2
3
4
27. Anxious
0
1
2
3
4
28. Helpless
0
1
2
3
4
29. Weary
0
1
2
3
4
30. Bewildered
0
1
2
3
4
31. Furious
0
1
2
3
4
32. Full of pep
0
1
2
3
4
33. Worthless
0
1
2
3
4
34. Forgetful
0
1
2
3
4
35. Vigorous
0
1
2
3
4
36. Uncertain about things
0
1
2
3
4
37. Bushed
0
1
2
3
4
249
FACT-G Quality of Life Questionnaire
Below is a list of statements that other cancer survivors have said are important. By circling one
(1) number per line, please indicate how true each statement has been for you DURING THE
PAST 7 DAYS.
PHYSICAL WELL-BEING
Not at
all
A
little
bit
I have a lack of energy………………………………….
0
1
2
3
4
I have nausea……………………………………………
0
1
2
3
4
Because of my physical condition, I have trouble
meeting the needs of my family………………………...
0
1
2
3
4
I have pain………………………………………………
0
1
2
3
4
I am bothered by side effects of treatment……………...
0
1
2
3
4
I feel ill………………………………………………….
0
1
2
3
4
I am forced to spend time in bed………………………..
0
1
2
3
4
Not at
all
A
little
bit
I feel close to my friends………………………………..
0
1
2
3
4
I get emotional support from my family………………..
0
1
2
3
4
I get support from my friends…………………………...
0
1
2
3
4
My family has accepted my illness……………………..
0
1
2
3
4
I am satisfied with family communication about my
illness……………………………………………………
0
1
2
3
4
I feel close to my partner (or the person who is my main
support)………………………………………………….
0
1
2
3
4
0
1
2
3
4
SOCIAL/FAMILY WELL-BEING
Some- Quite Very
what
a bit much
Some- Quite Very
what
a bit much
Regardless of your current level of sexual activity,
please answer the following question. If you prefer not
to answer it, please check this box (□) and go to the
next section.
I am satisfied with my sex life………………………….
250
By circling one (1) number per line, please indicate how true each statement has been for you
during the past 7 days.
EMOTIONAL WELL-BEING
Not at
all
A little
bit
Somewhat
Quite
a bit
Very
much
I feel
sad………………………………………………...
0
1
2
3
4
I am satisfied with how I am coping with my
illness…...
0
1
2
3
4
I am losing hope in the fight against my
illness………...
0
1
2
3
4
I feel
nervous……………………………………………
0
1
2
3
4
I worry about
dying……………………………………..
0
1
2
3
4
I worry that my condition will get
worse……………….
0
1
2
3
4
FUNCTIONAL WELL-BEING
Not at
all
A little
bit
Somewhat
Quite
a bit
Very
much
I am able to work (include work at
home)……………...
0
1
2
3
4
My work (include work at home) is
fulfilling…………..
0
1
2
3
4
I am able to enjoy
life…………………………………...
0
1
2
3
4
I have accepted my
illness………………………………
0
1
2
3
4
I am sleeping
well………………………………………
0
1
2
3
4
I am enjoying the things I usually do for
fun…………...
0
1
2
3
4
I am content with the quality of my life right
now……...
0
1
2
3
4
251
Calgary Symptoms of Stress Inventory (SOSI – C)
This questionnaire is designed to measure the different ways people respond to stressful
situations. The questionnaire contains sets of questions dealing with various physical,
psychological and behavioral responses. We are particularly interested in the frequency with
which you may have experienced these stress related symptoms during the past week.
Never
Infrequently
Sometimes
Often
Very Frequently
0
1
2
3
4
Stress is often accompanied by a variety of emotions.
During the last week, have you felt:
1. Like life is entirely hopeless
0
1
2
3
4
2. Unhappy and depressed
0
1
2
3
4
3. Alone and sad
0
1
2
3
4
4. That worrying gets you down
0
1
2
3
4
5. Like crying easily
0
1
2
3
4
6. That you wished you were dead
0
1
2
3
4
7. Frightening thoughts keep coming back
0
1
2
3
4
8. You suffer from severe nervous exhaustion
0
1
2
3
4
9. You become made or anger easily
0
1
2
3
4
10. When you feel angry, you act angrily toward most everything
0
1
2
3
4
11. You are easily annoyed and irritated
0
1
2
3
4
12. That little things get on your nerves
0
1
2
3
4
13. Angry thoughts about an irritating event keep bothering you
0
1
2
3
4
14. You let little annoyances build up until you just explode
0
1
2
3
4
Does it seem:
252
Never
Infrequently
Sometimes
Often
Very Frequently
0
1
2
3
4
15. Your anger is so great that you want to strike something
0
1
2
3
4
Muscle tension is a common way of experiencing stress.
Have you noticed excessive tension, stiffness, soreness or cramping in the muscles in your:
16. Shoulders
0
1
2
3
4
16. Neck
0
1
2
3
4
18. Back
0
1
2
3
4
19. Jaw
0
1
2
3
4
20. Forehead
0
1
2
3
4
21. Eyes
0
1
2
3
4
22. Hands or arms
0
1
2
3
4
23. Tension headaches
0
1
2
3
4
24. Thumping of your heart
0
1
2
3
4
25. Rapid or racing heart beats
0
1
2
3
4
26. Rapid breathing
0
1
2
3
4
27. Irregular heart beats
0
1
2
3
4
28. Difficult breathing
0
1
2
3
4
29. Pains in your heart of chest
0
1
2
3
4
0
1
2
3
4
Does it seem:
Do you experience:
30. Difficulty in staying asleep at night
253
Never
Infrequently
Sometimes
Often
Very Frequently
0
1
2
3
4
31. Hot or cold spells
0
1
2
3
4
32. Having to get up in the night to urinate
0
1
2
3
4
33. Sweating excessively even in cold weather
0
1
2
3
4
34. Having to urinate frequently
0
1
2
3
4
35. Early morning awakening
0
1
2
3
4
36. Flushing of your face
0
1
2
3
4
37. Difficulty in falling asleep
0
1
2
3
4
38. Breaking out in cold sweats
0
1
2
3
4
39. Feeling faint
0
1
2
3
4
40. Feeling weak and faint
0
1
2
3
4
41. Spells of severe dizziness
0
1
2
3
4
42. Nausea
0
1
2
3
4
43. Blurring of your vision
0
1
2
3
4
44. Severe pains in your stomach
0
1
2
3
4
45. You must do things very slowly to do them without mistakes
0
1
2
3
4
46. You get directions and orders wrong
0
1
2
3
4
47. Your thinking gets completely mixed-up when you have to do
things quickly
0
1
2
3
4
Have you experienced:
Does it seem:
254
Never
Infrequently
Sometimes
Often
Very Frequently
0
1
2
3
4
48. You have difficulty in concentrating
0
1
2
3
4
49. You become suddenly frightened for no good reason
0
1
2
3
4
50. You become so afraid you can't move
0
1
2
3
4
51. Colds
0
1
2
3
4
52. Hoarseness
0
1
2
3
4
53. Colds with complications (e.g. Bronchitis)
0
1
2
3
4
54. Nasal stuffiness
0
1
2
3
4
55. Having to clear your throat often
0
1
2
3
4
56. Sinus headaches
0
1
2
3
4
Have you experienced:
255
Borg Rate of Perceived Exertion
Rate your perception of exertion (i.e. how strenuous the exercise felt to you / how tired
you feel as a result).
6
No Exertion at All
7
8
Extremely Light
9
10
Very Light
11
12
Light
13
14
Somewhat Hard
15
16
Hard (Heavy)
17
18
Very Hard
19
Extremely Hard
20
Maximal Exertion
256
Dissertation Proposal
Qualitative Research Questions
YOGA INTERVENTION FOCUS GROUPS
o
o
o
o
o
o
o
o
o
o
o
o
o
o
Tell me about your experience in the yoga intervention.
Why did you decide to participate in the yoga intervention?
What did you expect when you started to practice yoga?
What was it like for you when you started out?
What has changed for you practicing yoga from the beginning of the intervention until
now?
What motivated you to continue the program?
What do you think you have learned from this program?
What did you like best / least about this intervention?
What do you think are some of the benefits / deficits to practicing yoga?
What effects, if any, have you noticed since joining the yoga programme?
What role does yoga play in your cancer recovery?
When people don't practice yoga as often as they'd like, why not?
What suggestions do you have for us to improve the program?
What additional information would you like to add today?
THREE & SIX MONTH FOLLOW-UP FOCUS GROUPS
o
o
o
o
o
o
o
o
Tell me about your yoga practice since completion of the intervention.
What has changed for you practicing yoga from completion of the intervention until
now?
What effects, if any, have you noticed since completing the yoga programme?
What has motivated you to continue practicing yoga?
What do you think are some of the benefits / deficits to practicing yoga?
What role does yoga play in your cancer recovery?
When people don't practice yoga as often as they'd like, why not?
What additional information would you like to add today?
LAB-BASED YOGA SESSION INDIVIDUAL INTERVIEWS
o
o
o
o
Tell me about your experience in the yoga session.
Can you describe how you felt while you were practicing yoga?
What helped you decide you felt that way?
What additional information would you like to add today?
257
ATTENDANCE LOG
Class Location
Class Day & Time
Session Dates (date to date, year)
Instructor: [Name]
NAME
Week 1
Week 2
Week 3
Week 4
Week 5
Week 6
Week 7
Attendance log – missing codes
W = work; V = vacation; S = sick; M = medical (i.e. cancer-related or injury); U (unknown); other
(please note reason why, if known)
*If a participant misses more than 2 classes for unknown reasons or informs you they are
dropping out, please follow-up to track reasons for ending the program or adverse events
as a result of the yoga.
258
YOGA for CANCER RESEARCH STUDY
(3 & 6 Month Follow-Up)
We would like to know what type of activities you may or may not be
participating in since completion of the yoga program. Please answer the
following questions:
Have you continued your yoga practice (at least once per week) over the past three months?
Yes  No 
If yes, on average how often are practicing yoga (sessions per week, including any classes
and/or home practice)? Please select one:
Once a week  Twice a week 
Five times a week 
Three times a week  Four times a week 
Six times a week 
Seven times a week 
Are you currently engaging in other activities in place of or in addition to yoga?
Yes  No 
If yes, please list the activity(s) you are currently doing on a weekly basis:
______________________________________________________________________________
______________________________________________________________________________
______________________________________________________________________________
______________________________________________________________________________
______________________________________________________________________________
______________________________________________________________________________
______________________________________________________________________________
259
Yoga for Cancer Survivors
7-Week Protocol
260
Week 1: Introduction to the basics
1. Legs up the Wall
2. Legs up the Wall: Hip Cross
3. Legs up the Wall: Pullovers(thumb tips touching) – up and down, and overhead
4. Sitting: Shoulders Rolls
5. Sitting:Breath Stretches
6. Tadasana: (feet 3 points) Rebound energy
7. Tree Pose (and prep)
8. Warrior 1/Crescent (cowboy surrender)
9. Pyramid
10. Floor: Knee to Belly
11. Floor: Supine Twist
12. Floor: Knee circles
13. Floor: Feet to floor
14. Savasana
261
Week 2: Building on the foundation, connecting to body and breath
1. Legs up the Wall
2. Legs up the Wall: Hip Cross (nostrils breathing into hip)
3. Legs up the Wall: Pullovers
4. Sitting: Shoulder Rolls
5. Sitting: Twist (nose in line with navel)
6. Hands and Knees
7. Hands and Knees:Cat and Dog
8. Tadasana
9. Tree Pose – 3 foot placements
10. Shiva twist
11. Warrior sequence: warrior 2, reverse warrior
12. Tadasana
13. Tadasana:Arm movements (flow)
14. Tadasana: Stand aware
15. Floor: Little bridge
16. Floor: Knees to belly
17. Floor:Twist on side – arm movements (like ASS, but not)
18. Savasana
262
Week 3: Building Strength, stability and flexibility
1. Legs up Wall
2. Legs up Wall: Hip Cross
3. Hands and Knees: Cat and Dog both directions
4. Floor: Child’s pose
5. Floor: Child’s pose plus 11 and 1
6. Lunge: arms on thigh/overhead
7. Tadasana: Standing side bend
8. Tree
9. Warrior 1 (cowboy surrender)
10. Pyramid
11. Warrior 3
12. Floor: TrA – silver platter, clock, no gripping
13. Floor: Marching one foot at at time; both feet
14. Floor: Little bridge
15. Floor: Supine twist
16. Floor: Hamstring release with strap – leg across body
17. Savasana – inhale pause inhale exhale pause exhale
263
Week 4: Opening Heart and Chest
1. Legs up Wall
2. Legs up Wall: Hip cross
3. On Floor, Knees Bent Feet on Floor: Connecting to PF, petals –
4. On Floor: Knee drops (soft jaw)
5. Sitting: Arms overhead with strap – circles overhead
6. Sitting: Arms behind back with strap
7. Sitting: Arms shoulder-ish height chest release – add fingers and wrists
8. Sitting: Eagle arms
9. Kneeling: Gateway
10. Tadasana: Arms shoulder-ish height – chest release – dural stretch
11. Tadasana: Eagle Arms
12. Tadasana: Shoulder Rolls
13. Triangle Pose
14. Tree Pose
15. Floor: Laying on Rolled Mat – Vertical
16. Floor: Laying on Rolled Mat – Vertical – Add Pullovers
17. Floor: Laying on Rolled Mat – Horizontal
18. Floor: Laying Twist - Lead with Torso, Arms Follow
19. Floor: Hamstring Traction
20. Savasana
264
Week 5 – Introducing meditation following savasana
1. Floor: Laying over yoga mat
- breathe easy
2. Floor: Knee Drops (no mat)
3. Lunge: hands on thigh
Reverse Lunge
Lunge Sequence forward and back
4. Tadasana: samastiti, feel feet equally, collar bones broad, easy breath
5. Tadasana: Side Bend
- body allows for it, right arm alongside ear
- you are stability, strength and balancee
6. Tadasana: Arms behind – chest release
- neck at ease
- finger tips move behind
- feeling energy of heart moving through arms going out and coming back.
7. Tadasana:Arm flow
- bringing sky toward your heart
- follow movement with breath
8. Tadasana:Arm flow 2
- arms to sky, palms touch – to heart.
- come back together
- palms to sky, and then all the way down to hips
9.Tree pose into Warrior1 /Crescent, Pyramid and Warrior 3
- arms to cowboy surrender
- arms to sky
10. Tadasana
11. Tadasana: Shoulder rolls
12. Tadasana: Eagle Arms
13. Floor: Plank Pose
- Fingers interlaced –forearms on floor – connect with pelvic floor.
- Press heels back. Lowering hips – watch how much
14. Floor: Cobra
- Initially, don’t use arms or hands strength. Peel off.
- Then, engage torso with legs connect with energy in legs. Feet moving backward.
- Tone in legs. Pinky toe touching floor. Peel up. Then add arms.
15. Floor: Little Bridge
16. Floor: Knee to belly (one leg extended along the floor) into Supine Twists.
- Pinching in hip flexor tone in right leg. Head of thigh bone in hip socket. Opposite heel
to shoulder. Sense all the other parts of you.
17.Savasana
18. Meditation
265
Week 6 – Bringing it all together (part 1)
1. Legs up the Wall – so hum – 5 minutes
2. Legs up the Wall - Press heels up to the ceiling-release and press
3. Sitting
4. Sitting: Massage along inner side of shin.
5. Sitting: Downward energy into and through pelvis. Feel rebound
6. Sitting: Tick Tocks Side to Side with breath
7. Sitting: Tick Tocks Forward and Backward with breath
8. Sitting: Sequential Twist with hands on shoulders
9. Sitting: Shoulder Rolls
10. Sitting: Breath Stretches
11. Tadasana: Bottom of feet - inner and outer lines,energy flow – up inner lines and down
outer lines
12. Tadasana: Chest and Arm Release using the wall
13.
continue to feel the lines
half a step forward
openness in chest
step back, then step out, and raise hand up the wall repeat
repeat 3 or 4 times
Tadasana: Dancer’s Pose Modified with wall
14. Tadasana: Half Dog, Table Top (keep in mind energy lines)
15. Tadasana Standing Hip Release
16. Tadasana: Chair Pose
17. Tadasana: Table Top with Feet Wide (aka Standing Wide Legged FB)
18. Floor: Supine Twist (outer hip to outer ankle)
19. Floor: Hamstring Release
- Be aware of the pelvis as bottom leg moves away
- Move leg across body, Move leg 6 inches away from midline – energy lines
20. Savasana: Breath – Softening front of torso and body softening.
266
Week 7 – Bringing it all together (part 2)
1. A) Legs up the Wall or
B) Laying on Rolled Yoga Mat Vertically along spine
2. In the position as above: Pullovers – 2 positions
3. Hands and Knees: Cat and Dog
- Leading with tail
- Leading with head
4. Lunge: arms can go overhead.
5. Lunge: Reverse Lunge
6. Kneeling: Gateway Pose
7. Tadasana: Feel energetic roots
8. Sequence: Tree Pose, Warrior 1/Crescent Pose, Warrior 3, Warrior 1, Tadasana.
9. Tadasana: Shiva Twist
10. Tadasana: Muscular and Dural Stretch of the arms and chest
11. Tadasana: Eagle Arms
12. Tadasana: Dancer’s Pose
13. Tadasana: Table Top
14. Floor: Twist – ASS
15. Floor: Suping Twist
16. Floor: Knees to Belly – Circles
17. Floor: Marching
18. Floor: Knee Drops
19. Floor: Little Bridge
20. Floor: Hamstring Traction
21. Savasana
22. Sitting Quietly
267
Michael Mackenzie
Study 3 Yoga Protocol
Legs up the Wall
15. Static Legs up the Wall
16. Legs up the Wall: Hip Cross
17. Legs up the Wall: Pullovers (thumb tips touching)
Seated
18. Sitting: Shoulders Rolls
19. Sitting: Breath Stretches
20. Sitting: Twist (nose in line with navel)
Kneeling
21. Hands and Knees: Cat and Dog
- Leading with tail
- Leading with head
22. Chakravakasana
23. Lunge
24. Child’s pose
Standing
25. Tadasana:
- (feet 3 points) Rebound energy
- samastiti, feel feet equally, collar bones broad, easy breath
26. Tadasana: Arm flow
- bringing sky toward your heart
- follow movement with breath
27. Tree Pose (and prep)
28. Warrior 1/Crescent (cowboy surrender)
29. Forward bend
Supine
30. Floor: Breath awareness (hands on chest/belly)
31. Floor: Little bridge
32. Floor: Knee to Belly
33. Floor: Supine Twist
34. Floor: Knee circles
Resting
35. Savasana
- Breath @ nostrils
- Body scan
- Whole body breathing
- Softening breath/body
Téléchargement