Handbook of Research on Evidence-Based Perspectives on the Psychophysiology of Yoga and Its Applications 1799832546, 9781799832546

While yoga was originally intended to be practiced for spiritual growth, there is an increasing interest in applying yog

2,551 218 8MB

English Pages 612 [607] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Handbook of Research on Evidence-Based Perspectives on the Psychophysiology of Yoga and Its Applications
 1799832546, 9781799832546

Table of contents :
Editorial Advisory Board
List of Contributors
Table of Contents
Preface
Acknowledgment
Section 1: The Physiological Effects of Yoga Practice
1 Psychophysiological Effects and the Applications of Yoga Breathing Practices • Niranjan Kala, Dipak Chetry, Shirley Telles
2 Psychophysiology: Healing Effects of Voluntarily Regulated Breathing Practices • Heather Mason, Patricia L. Gerbarg, Richard P. Brown
3 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama • Ramakrishnan Angarai Ganesan
4 Neurobiology of Meditation • Danilo Forghieri Santaella
5 The Neurophilosophy of Meditation • Vinod D. Deshmukh,
6 The Yoga of the Bhagavad Gita: Spirituality, Meditation, and the Rise of a New Scientific Paradigm • Rubens Turci
7 Molecular Mechanisms Underlying the Effects of Yoga • Surabhi Gautam, Rima Dada
8 Neurocognitive Mechanisms of Yoga: Implications for Yoga Therapy • Irina Sheftel, Anneke Sips
Section 2: Yoga as a Therapeutic Intervention
9 Standards in Yoga Research and Reporting • Steffany Moonaz, Daryl Nault, Atiera Abatemarco
10 Yoga Therapy: An Overview of Key Research and the Underlying Mechanisms • Eileen M. Lafer, Margaret McCuiston, Ann Swanson
11 Yoga for Mental Health Disorders: Research and Practice • Bhargav Hemant, Rashmi Arasappa, Inbaraj G., Kaviraja Udupa, Shivarama Varambally
12 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders: From Suffering to Thriving • Elin-Kristin Hem Olsen
13 Clinicians Applying Yoga Principles and Practices in Pain Care: An Evidence-Informed Approach • Shelly Prosko, Matthew J. Taylor
14 Yoga and Lifestyle Factors in Heart Disease • M. Mala Cunningham
15 Advances in Understanding the Use of Yoga as Therapy in Lymphedema • S. R. Narahari, Madhur Guruprasad Aggithaya, Terence J. Ryan
16 Yoga and Weight Management: A Narrative Review • Sachin Kumar Sharma, Sushma Pal, Shirley Telles
Section 3: Applications and Mechanisms of Yoga
17 Overview of Yoga for Teenagers in the UK: The Rationale, Evidence Base, and the Application • Charlotta Martinus, Nicholas A. Kearney
18 Yoga and Aging: Neurobiological Benefits • Rui F. Afonso, Danilo F. Santaella, Elisa Harumi Kozasa
19 Special Yoga for Children and Young People With Special Needs • Jyoti Jo Manuel
20 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder • Pauline Jensen
21 Yoga: A Multi-Dimensional Therapeutic Approach to Autism Spectrum Disorder • Kankan Gulati, Praerna H. Bhargav, Shalu Elizabeth Abraham, Hemant Bhargav
22 Yoga and Polyvagal Theory: Effects and Applications for Well-Being • Marlysa Sullivan
23 Traditional Frameworks of Well-Being and Modern Science • Ganesh Mohan
24 The Science and Art of Contemporary Breathwork and Yoga • Jim Morningstar, Jessica Dibb
25 An Exploration of Influence of Duration on Physiological Effects of Asanas • Debadutta Subudhi, M. Manivannan
Compilation of References
About the Contributors
Index

Citation preview

Handbook of Research on Evidence-Based Perspectives on the Psychophysiology of Yoga and Its Applications Shirley Telles Patanjali Research Foundation, India Ram Kumar Gupta Patanjali Research Foundation, India

A volume in the Advances in Medical Diagnosis, Treatment, and Care (AMDTC) Book Series

Published in the United States of America by IGI Global Medical Information Science Reference (an imprint of IGI Global) 701 E. Chocolate Avenue Hershey PA, USA 17033 Tel: 717-533-8845 Fax: 717-533-8661 E-mail: [email protected] Web site: http://www.igi-global.com Copyright © 2021 by IGI Global. All rights reserved. No part of this publication may be reproduced, stored or distributed in any form or by any means, electronic or mechanical, including photocopying, without written permission from the publisher. Product or company names used in this set are for identification purposes only. Inclusion of the names of the products or companies does not indicate a claim of ownership by IGI Global of the trademark or registered trademark. Library of Congress Cataloging-in-Publication Data Names: Telles, Shirley, editor. | Gupta, Ram Kumar, 1986- editor. Title: Handbook of research on evidence-based perspectives on the psychophysiology of yoga and its applications / Shirley Telles and Ram Kumar Gupta, editors. Description: Hershey PA : Medical Information Science Reference, [2020] | Includes bibliographical references and index. | Summary: “This book explores the effects of yoga and its applications from physiological, psychological, therapeutic, and biomedical engineering perspectives”-Provided by publisher. Identifiers: LCCN 2019049661 (print) | LCCN 2019049662 (ebook) | ISBN 9781799832546 (hardcover) | ISBN 9781799832553 (ebook) Subjects: LCSH: Yoga--Therapeutic use--Case studies. | Yoga--Physiological aspects. | Yoga--Psychological aspects. | Psychophysiology. Classification: LCC RM727.Y64 E95 2020 (print) | LCC RM727.Y64 (ebook) | DDC 613.7/046--dc23 LC record available at https://lccn.loc.gov/2019049661 LC ebook record available at https://lccn.loc.gov/2019049662 This book is published in the IGI Global book series Advances in Medical Diagnosis, Treatment, and Care (AMDTC) (ISSN: 2475-6628; eISSN: 2475-6636) British Cataloguing in Publication Data A Cataloguing in Publication record for this book is available from the British Library. All work contributed to this book is new, previously-unpublished material. The views expressed in this book are those of the authors, but not necessarily of the publisher. For electronic access to this publication, please contact: [email protected].

Advances in Medical Diagnosis, Treatment, and Care (AMDTC) Book Series ISSN:2475-6628 EISSN:2475-6636 Mission

Advancements in medicine have prolonged the life expectancy of individuals all over the world. Once life-threatening conditions have become significantly easier to treat and even cure in many cases. Continued research in the medical field will further improve the quality of life, longevity, and wellbeing of individuals. The Advances in Medical Diagnosis, Treatment, and Care (AMDTC) book series seeks to highlight publications on innovative treatment methodologies, diagnosis tools and techniques, and best practices for patient care. Comprised of comprehensive resources aimed to assist professionals in the medical field apply the latest innovations in the identification and management of medical conditions as well as patient care and interaction, the books within the AMDTC series are relevant to the research and practical needs of medical practitioners, researchers, students, and hospital administrators.

Coverage • Chronic Conditions • Cancer Treatment • Medical Procedures • Experimental Medicine • Diagnostic Medicine • Internal Medicine • Patient Interaction • Alternative Medicine • Disease Management • Medical Testing

IGI Global is currently accepting manuscripts for publication within this series. To submit a proposal for a volume in this series, please contact our Acquisition Editors at [email protected] or visit: http://www.igi-global.com/publish/.

The Advances in Medical Diagnosis, Treatment, and Care (AMDTC) Book Series (ISSN 2475-6628) is published by IGI Global, 701 E. Chocolate Avenue, Hershey, PA 17033-1240, USA, www.igi-global.com. This series is composed of titles available for purchase individually; each title is edited to be contextually exclusive from any other title within the series. For pricing and ordering information please visit http:// www.igi-global.com/book-series/advances-medical-diagnosis-treatment-care/129618. Postmaster: Send all address changes to above address. © © 2021 IGI Global. All rights, including translation in other languages reserved by the publisher. No part of this series may be reproduced or used in any form or by any means – graphics, electronic, or mechanical, including photocopying, recording, taping, or information and retrieval systems – without written permission from the publisher, except for non commercial, educational use, including classroom teaching purposes. The views expressed in this series are those of the authors, but not necessarily of IGI Global.

Titles in this Series

For a list of additional titles in this series, please visit: http://www.igi-global.com/book-series/advances-medical-diagnosistreatment-care/129618

Diagnostic Techniques and Therapeutic Strategies for Parotid Gland Diorders Mahmoud Sakr (Alexandria University, Egypt) Medical Information Science Reference • © 2021 • 362pp • H/C (ISBN: 9781799856030) • US $295.00 Evaluation and Management of High-Risk Pregnancies Emerging Research and Opportunities Sapna Nanda (Panjab University, India) Medical Information Science Reference • © 2021 • 216pp • H/C (ISBN: 9781799843573) • US $195.00 Noninvasive Ventilation Technologies and Healthcare for Geriatric Patients César Fonseca (Universidade de Évora, Portugal) Manuel José Lopes (Universidade de Évora, Portugal) David Mendes (Universidade de Évora, Portugal) Jose Garcia-Alonso (University of Extremadura, Spain) and Felismina Mendes (Universidade de Évora, Portugal) Medical Information Science Reference • © 2020 • 252pp • H/C (ISBN: 9781799835318) • US $245.00 Diagnostic and Treatment Methods for Ulcerative Colitis and Colitis-Associated Cancer Ashok Kumar Pandurangan (B.S. Abdur Rahman Crescent Institute of Science and Technology, India) Medical Information Science Reference • © 2020 • 300pp • H/C (ISBN: 9781799835806) • US $295.00 Biopsychosocial Perspectives and Practices for Addressing Communicable and Non-Communicable Diseases Simon George Taukeni (University of Namibia, Namibia) Medical Information Science Reference • © 2020 • 368pp • H/C (ISBN: 9781799821397) • US $265.00 Role of Nutrition in Providing Pro-/Anti-Inflammatory Balance Emerging Research and Opportunities Uğur Günşen (Bandırma Onyedi Eylül University, Turkey) and Ramazan Mert Atan (Bandırma Onyedi Eylül University, Turkey) Medical Information Science Reference • © 2020 • 214pp • H/C (ISBN: 9781799835943) • US $235.00 Handbook of Research on Prenatal, Postnatal, and Early Childhood Development Neriman Aral (Ankara University, Turkey) Medical Information Science Reference • © 2020 • 470pp • H/C (ISBN: 9781799829522) • US $365.00 Ethnomedicinal Plant Use and Practice in Traditional Medicine Akash (Gurukul Kangri University, India) Navneet (Gurukul Kangri University, India) and B.S. Bhandari (Garhwal University, India) Medical Information Science Reference • © 2020 • 380pp • H/C (ISBN: 9781799813200) • US $295.00

701 East Chocolate Avenue, Hershey, PA 17033, USA Tel: 717-533-8845 x100 • Fax: 717-533-8661 E-Mail: [email protected] • www.igi-global.com

Shirley Telles and Ram Kumar Gupta’s book, Evidence-Based Perspectives on the Psychophysiology of Yoga and Its Applications is an inspiring guidebook that offers us a masterful collection of essays of practical wisdom for attaining health, healing and well-being that will motivate its reader for years to come. The rich variety of writing on the therapeutic benefits of yoga offers us a superb feast for our mind, body, and spirit. Bravo! I offer my deep bow of gratitude for the depth of expertise Telles and Gupta offer us through their care and loving attention in bringing forward this masterful treatise. Richard Miller, PhD. Founder iRest Institute, developer of iRest Meditation, and author of iRest Meditation: Restorative Practices for Health, Resiliency, and Well-Being. As medicine integrates yoga practices to support optimum mental and physical health, medical and yoga professionals need the scientific evidence that Evidence-Based Perspectives on the Psychophysiology of Yoga and Its Applications provides. The chapters, written by senior scientists and master yoga therapists from around the world, support the therapeutic use of yoga. If you are a healing professional, EvidenceBased Perspectives on the Psychophysiology of Yoga and Its Applications belongs on your bookshelf. Amy Weintraub, author of Yoga Skills for Therapists, Yoga for Depression and Temple Dancer



Editorial Advisory Board Gerry Bodeker, Columbia University, USA & Green Templeton College, University of Oxford, USA & Mental Wellness Initiative of the Global Wellness Institute, USA Ann C. DeBaldo, University of South Florida, USA Richard Miller, iRest Institute, USA Girishwar Misra, Mahatma Gandhi Antarrashtriya Hindi Vishwavidyalaya, India Terence Ryan, Oxford University, UK William Selvamurthy, Amity Science, Technology, and Innovation Foundation, Amity University, India Thaiyar M. Srinivasan, S-VYASA Yoga University, India

 

List of Contributors

Abatemarco, Atiera / Maryland University of Integrative Health, USA........................................... 144 Abraham, Shalu Elizabeth / National Institute of Mental Health and Neurosciences, Bengaluru, India.............................................................................................................................................. 361 Afonso, Rui F. / Hospital Israelita Albert Einstein, Brazil................................................................. 317 Aggithaya, Madhur Guruprasad / Institute of Applied Dermatology, Kasaragod, India................ 263 Arasappa, Rashmi / National Institute of Mental Health and Neurosciences, Bengaluru, India..... 179 Bhargav, Hemant / National Institute of Mental Health and Neurosciences, Bengaluru, India....... 361 Bhargav, Praerna H. / National Institute of Mental Health and Neurosciences, Bengaluru, India.. 361 Brown, Richard P. / College of Physicians and Surgeons, Columbia University, USA....................... 24 Chetry, Dipak / Patanjali Research Foundation, India......................................................................... 1 Cunningham, M. Mala / University of Virginia, USA....................................................................... 242 Dada, Rima / All India Institute of Medical Sciences, New Delhi, India.......................................... 103 Deshmukh, Vinod D. / University of Florida, USA.............................................................................. 72 Dibb, Jessica / Inspiration Consciousness School, USA.................................................................... 422 G., Inbaraj / National Institute of Mental Health and Neurosciences, Bengaluru, India................. 179 Ganesan, Ramakrishnan Angarai / Indian Institute of Science, Bangalore, India............................ 49 Gautam, Surabhi / All India Institute of Medical Sciences, New Delhi, India................................. 103 Gerbarg, Patricia L. / New York Medical College, USA..................................................................... 24 Gulati, Kankan / National Institute of Mental Health and Neurosciences, Bengaluru, India.......... 361 Hemant, Bhargav / National Institute of Mental Health and Neurosciences, Bengaluru, India...... 179 Jensen, Pauline / University of Sydney, Australia............................................................................. 347 Kala, Niranjan / Patanjali Research Foundation, India....................................................................... 1 Kearney, Nicholas A. / Teen Yoga Foundation, UK........................................................................... 305 Kozasa, Elisa Harumi / Hospital Israelita Albert Einstein, Brazil.................................................... 317 Lafer, Eileen M. / University of Texas Health Science Center at San Antonio, USA & Beyond Wellness Aljezur, Portugal............................................................................................................. 159 Manivannan, M. / Indian Institute of Technology, Madras, India.................................................... 451 Manuel, Jyoti Jo / Special Yoga, UK.................................................................................................. 330 Martinus, Charlotta / Teen Yoga Foundation, UK........................................................................... 305 Mason, Heather / The Minded Institute, UK....................................................................................... 24 McCuiston, Margaret / Southern California University of Health Sciences, USA & Seashore Healing, USA................................................................................................................................. 159 Mohan, Ganesh / Svastha Yoga and Ayurveda, Singapore................................................................ 410 Moonaz, Steffany / Maryland University of Integrative Health, USA............................................... 144 Morningstar, Jim / Transformations Incorporated, USA.................................................................. 422 



Narahari, S. R. / Institute of Applied Dermatology, Kasaragod, India.............................................. 263 Nault, Daryl / Maryland University of Integrative Health, USA....................................................... 144 Olsen, Elin-Kristin Hem / Drammen District Psychiatric Centre, Norway...................................... 199 Pal, Sushma / Patanjali Research Foundation, India........................................................................ 283 Prosko, Shelly / PhysioYoga, Canada................................................................................................ 221 Ryan, Terence J. / Department of Dermatology, Churchill Hospital, UK......................................... 263 Santaella, Danilo F. / Hospital Israelita Albert Einstein, Brazil........................................................ 317 Santaella, Danilo Forghieri / Sports Center, University of São Paulo, Brazil..................................... 61 Sharma, Sachin Kumar / Patanjali Research Foundation, India...................................................... 283 Sheftel, Irina / Network Yoga Therapy, The Netherlands.................................................................. 124 Sips, Anneke / Network Yoga Therapy, The Netherlands................................................................... 124 Subudhi, Debadutta / Indian Institute of Technology, Madras, India.............................................. 451 Sullivan, Marlysa / Maryland University of Integrative Health, USA............................................... 391 Swanson, Ann / Maryland University of Integrative Health, USA & Kripalu School of Integrative Yoga Therapy, USA........................................................................................................................ 159 Taylor, Matthew J. / MyRehab, LLC, USA........................................................................................ 221 Telles, Shirley / Patanjali Research Foundation, India................................................................. 1, 283 Turci, Rubens / Universidade Estadual do Rio de Janeiro, Brazil...................................................... 85 Udupa, Kaviraja / National Institute of Mental Health and Neurosciences, Bengaluru, India........ 179 Varambally, Shivarama / National Institute of Mental Health and Neurosciences, Bengaluru, India.............................................................................................................................................. 179

Table of Contents

Preface................................................................................................................................................. xxii Acknowledgment.............................................................................................................................xxviii Section 1 The Physiological Effects of Yoga Practice Chapter 1 Psychophysiological Effects and the Applications of Yoga Breathing Practices.................................... 1 Niranjan Kala, Patanjali Research Foundation, India Dipak Chetry, Patanjali Research Foundation, India Shirley Telles, Patanjali Research Foundation, India Chapter 2 Psychophysiology: Healing Effects of Voluntarily Regulated Breathing Practices............................... 24 Heather Mason, The Minded Institute, UK Patricia L. Gerbarg, New York Medical College, USA Richard P. Brown, College of Physicians and Surgeons, Columbia University, USA Chapter 3 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama...................... 49 Ramakrishnan Angarai Ganesan, Indian Institute of Science, Bangalore, India Chapter 4 Neurobiology of Meditation.................................................................................................................. 61 Danilo Forghieri Santaella, Sports Center, University of São Paulo, Brazil Chapter 5 The Neurophilosophy of Meditation...................................................................................................... 72 Vinod D. Deshmukh, University of Florida, USA Chapter 6 The Yoga of the Bhagavad Gita: Spirituality, Meditation, and the Rise of a New Scientific Paradigm................................................................................................................................................ 85 Rubens Turci, Universidade Estadual do Rio de Janeiro, Brazil 



Chapter 7 Molecular Mechanisms Underlying the Effects of Yoga..................................................................... 103 Surabhi Gautam, All India Institute of Medical Sciences, New Delhi, India Rima Dada, All India Institute of Medical Sciences, New Delhi, India Chapter 8 Neurocognitive Mechanisms of Yoga: Implications for Yoga Therapy............................................... 124 Irina Sheftel, Network Yoga Therapy, The Netherlands Anneke Sips, Network Yoga Therapy, The Netherlands Section 2 Yoga as a Therapeutic Intervention Chapter 9 Standards in Yoga Research and Reporting......................................................................................... 144 Steffany Moonaz, Maryland University of Integrative Health, USA Daryl Nault, Maryland University of Integrative Health, USA Atiera Abatemarco, Maryland University of Integrative Health, USA Chapter 10 Yoga Therapy: An Overview of Key Research and the Underlying Mechanisms............................... 159 Eileen M. Lafer, University of Texas Health Science Center at San Antonio, USA & Beyond Wellness Aljezur, Portugal Margaret McCuiston, Southern California University of Health Sciences, USA & Seashore Healing, USA Ann Swanson, Maryland University of Integrative Health, USA & Kripalu School of Integrative Yoga Therapy, USA Chapter 11 Yoga for Mental Health Disorders: Research and Practice.................................................................. 179 Bhargav Hemant, National Institute of Mental Health and Neurosciences, Bengaluru, India Rashmi Arasappa, National Institute of Mental Health and Neurosciences, Bengaluru, India Inbaraj G., National Institute of Mental Health and Neurosciences, Bengaluru, India Kaviraja Udupa, National Institute of Mental Health and Neurosciences, Bengaluru, India Shivarama Varambally, National Institute of Mental Health and Neurosciences, Bengaluru, India Chapter 12 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders: From Suffering to Thriving............................................................................................................................................... 199 Elin-Kristin Hem Olsen, Drammen District Psychiatric Centre, Norway



Chapter 13 Clinicians Applying Yoga Principles and Practices in Pain Care: An Evidence-Informed Approach.221 Shelly Prosko, PhysioYoga, Canada Matthew J. Taylor, MyRehab, LLC, USA Chapter 14 Yoga and Lifestyle Factors in Heart Disease....................................................................................... 242 M. Mala Cunningham, University of Virginia, USA Chapter 15 Advances in Understanding the Use of Yoga as Therapy in Lymphedema......................................... 263 S. R. Narahari, Institute of Applied Dermatology, Kasaragod, India Madhur Guruprasad Aggithaya, Institute of Applied Dermatology, Kasaragod, India Terence J. Ryan, Department of Dermatology, Churchill Hospital, UK Chapter 16 Yoga and Weight Management: A Narrative Review.......................................................................... 283 Sachin Kumar Sharma, Patanjali Research Foundation, India Sushma Pal, Patanjali Research Foundation, India Shirley Telles, Patanjali Research Foundation, India Section 3 Applications and Mechanisms of Yoga Chapter 17 Overview of Yoga for Teenagers in the UK: The Rationale, Evidence Base, and the Application..... 305 Charlotta Martinus, Teen Yoga Foundation, UK Nicholas A. Kearney, Teen Yoga Foundation, UK Chapter 18 Yoga and Aging: Neurobiological Benefits......................................................................................... 317 Rui F. Afonso, Hospital Israelita Albert Einstein, Brazil Danilo F. Santaella, Hospital Israelita Albert Einstein, Brazil Elisa Harumi Kozasa, Hospital Israelita Albert Einstein, Brazil Chapter 19 Special Yoga for Children and Young People With Special Needs..................................................... 330 Jyoti Jo Manuel, Special Yoga, UK Chapter 20 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder........................ 347 Pauline Jensen, University of Sydney, Australia



Chapter 21 Yoga: A Multi-Dimensional Therapeutic Approach to Autism Spectrum Disorder........................... 361 Kankan Gulati, National Institute of Mental Health and Neurosciences, Bengaluru, India Praerna H. Bhargav, National Institute of Mental Health and Neurosciences, Bengaluru, India Shalu Elizabeth Abraham, National Institute of Mental Health and Neurosciences, Bengaluru, India Hemant Bhargav, National Institute of Mental Health and Neurosciences, Bengaluru, India Chapter 22 Yoga and Polyvagal Theory: Effects and Applications for Well-Being............................................... 391 Marlysa Sullivan, Maryland University of Integrative Health, USA Chapter 23 Traditional Frameworks of Well-Being and Modern Science............................................................. 410 Ganesh Mohan, Svastha Yoga and Ayurveda, Singapore Chapter 24 The Science and Art of Contemporary Breathwork and Yoga............................................................ 422 Jim Morningstar, Transformations Incorporated, USA Jessica Dibb, Inspiration Consciousness School, USA Chapter 25 An Exploration of Influence of Duration on Physiological Effects of Asanas.................................... 451 Debadutta Subudhi, Indian Institute of Technology, Madras, India M. Manivannan, Indian Institute of Technology, Madras, India Compilation of References................................................................................................................ 475 About the Contributors..................................................................................................................... 562 Index.................................................................................................................................................... 574

Detailed Table of Contents

Preface................................................................................................................................................. xxii Acknowledgment.............................................................................................................................xxviii Section 1 The Physiological Effects of Yoga Practice Chapter 1 Psychophysiological Effects and the Applications of Yoga Breathing Practices.................................... 1 Niranjan Kala, Patanjali Research Foundation, India Dipak Chetry, Patanjali Research Foundation, India Shirley Telles, Patanjali Research Foundation, India Yoga is an ancient practice that originated in India and aims at purifying the mind for spiritual progress. In modern times yoga is widely practiced for general health and well-being as well as for therapeutic reasons. Voluntary breath regulation or pranayama has been given significant importance in traditional texts as well as by yoga masters. Research has shown beneficial effects of yoga breathing practices or pranayamas on neurocognitive, metabolic, respiratory, and autonomic functions, which are discussed in the chapter. The chapter also discusses the applications of these practices for the management of various clinical conditions as well as for alleviating psychological problems associated with particular illnesses. The beneficial effects of yoga breathing practices demonstrate the importance of these safe and cost effective non-pharmacological interventions for general health as well as for prevention and management of various diseases. Chapter 2 Psychophysiology: Healing Effects of Voluntarily Regulated Breathing Practices............................... 24 Heather Mason, The Minded Institute, UK Patricia L. Gerbarg, New York Medical College, USA Richard P. Brown, College of Physicians and Surgeons, Columbia University, USA This chapter describes the physiological mechanisms that underpin the varying effects of different types of breath practices inherent in the yoga tradition and ultimately the role that breathing techniques play in person and public health. Concurrently, the script elucidates how different practices alter psychophysiological states clarifying why and how they may be employed with specific health populations, how they may enhance and or maintain well-being, and clear guidance regarding precautions and contraindications. 



Chapter 3 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama...................... 49 Ramakrishnan Angarai Ganesan, Indian Institute of Science, Bangalore, India The benefits of pranayama for positive health are well known. Even though there are many studies published on the effectiveness of pranayama, there are very few papers that actually have systematically studied the physiological mechanisms involved, causing the benefits of pranayama, especially with respect to the cardiac function. This chapter attempts to have a detailed look at the physiology behind deep breathing. The chapter also conjectures that voluntary, deep breathing with attention may have a role to play in faster recovery from surgeries and prevent or delay the onset of Alzheimer’s disease, Parkinson’s disease, and maybe even cancer. Extended, carefully controlled, and detailed studies are needed to prove or disprove these conjectures. Chapter 4 Neurobiology of Meditation.................................................................................................................. 61 Danilo Forghieri Santaella, Sports Center, University of São Paulo, Brazil Meditation should not be considered a simple activity that is performed with focused attention; this is concentration. When practicing concentration correctly, with a good “anchor” for attention, a specific state of mind takes place, in which logic relaxation happens, and there is a relative freedom from selfidentification. Such states of mind are to be experienced and cannot be practiced; thus, meditation techniques (concentrations) are the means to reach this goal. Those who achieve such a state experience positive neurophysiological effects, which have been studied for decades, such as increased functionality and connectivity of the brain, and also increased gray matter volume in specific cortical areas, whether in the young or in the elderly. Meditation has, thus, a proven potential role to help one maintain a healthy cognition and should be included in daily life routines of everybody who wishes for it. Chapter 5 The Neurophilosophy of Meditation...................................................................................................... 72 Vinod D. Deshmukh, University of Florida, USA Meditation is the art and science of optimal self-integration. We are naturally nested in Nature. Meditation is being at-peace with oneself and at-home in the world. It is the art of self-observation, learning, and developing presence of mind and selfless love. It is a unique mode of being present in the moment. It is a natural state of health and happiness. We should let go of the vicious cycle of negative feelings and discover the virtuous cycle of positive feelings like joy, cheerfulness, equanimity, friendliness, compassion, and love. This is possible through meditative practices like yoga, mindfulness, flow-state, and nature-experience. With advancement in neurobiology, we can understand how these meditative skills are learned, developed, evolved, and mastered. These meditative skills and lifestyle are the key to positive psychology and mental health. These secular meditative practices are being recommended in our education and health care systems. Chapter 6 The Yoga of the Bhagavad Gita: Spirituality, Meditation, and the Rise of a New Scientific Paradigm................................................................................................................................................ 85 Rubens Turci, Universidade Estadual do Rio de Janeiro, Brazil



This chapter explores the fact that śraddhā (truth-force or heart force; love-in-action, self-reliance, faithin-oneself, and also enthusiasm) may be said to represent the key noticeable outcome of spirituality, yoga, and meditation, and also that śraddhā can be seen as the main category to bridge the gap between science and spirituality. Being itself the truth-force that is common to scientists (atheists or not) and people of different faiths, śraddhā represents a conceptual category that helps us to explain how science evolves shifting from one paradigm to another. Chapter 7 Molecular Mechanisms Underlying the Effects of Yoga..................................................................... 103 Surabhi Gautam, All India Institute of Medical Sciences, New Delhi, India Rima Dada, All India Institute of Medical Sciences, New Delhi, India Complex chronic lifestyle disorders are the leading causes of death and disability worldwide. Stress and anxiety associated with today’s hectic life schedule and polluted environment have contributed a lot in triggering and causing many chronic diseases and decreased quality of life, even with pharmacologic treatment. Most of the chronic complex diseases, such as chronic obstructive pulmonary disease, depression, autoimmune diseases, cancer, cardiovascular diseases, obesity, and diabetes mellitus share underlying mechanisms like high levels of stress, anxiety, depression, oxidative stress, shorter telomeres, persistent activation of hypothalamo-pituitary adrenal axis, inflammation and dysregulated immune system, and thus need to be managed by an integrated approach that targets both mind and body. The individuals with these conditions have been reported to benefit from yoga, but the underlying mechanism of action of yoga remains unclear. The aim of this chapter is to summarize the mechanism of action underlying the cumulative effect of yoga on multiple pathways at a cellular level. Chapter 8 Neurocognitive Mechanisms of Yoga: Implications for Yoga Therapy............................................... 124 Irina Sheftel, Network Yoga Therapy, The Netherlands Anneke Sips, Network Yoga Therapy, The Netherlands This chapter combines insights from the neuroscientific research on yoga and meditation with the practical application of yoga therapy. The chapter opens with an overview of neuroscientific concepts related to the practice of yoga and meditation, such as attention, emotional regulation, body awareness, and autonomic regulation. The authors summarise the known effects of yoga and meditation, in relation to these concepts and common mental health disorders. The chapter continues with a case study, a personalised yoga therapy intervention in a client experiencing psychotic symptoms. The intervention is grounded in yoga philosophy, and in the neuroscientific concepts introduced previously. This section is written from the perspective of a yoga therapist and includes practical tips. In the final section, the authors suggest possible directions for future research and implementation of yoga therapy as a complementary treatment in mental healthcare.



Section 2 Yoga as a Therapeutic Intervention Chapter 9 Standards in Yoga Research and Reporting......................................................................................... 144 Steffany Moonaz, Maryland University of Integrative Health, USA Daryl Nault, Maryland University of Integrative Health, USA Atiera Abatemarco, Maryland University of Integrative Health, USA Yoga research is growing rapidly in volume and rigor but varies in sample size, study design, and reporting transparency. Yoga professionals may not be versed in the current research due to a lack of research literacy and may not be well positioned to discern research relevance and quality. Research literacy is necessary to apply research in yoga teaching and therapeutics. Research is part of evidenceinformed practice, along with clinical experience and client preferences, and is aligned with concepts from yoga philosophy. Several strategies are available to improve research literacy and evidence-informed practice for yoga professionals, which can help to expand inter-professional collaboration and inform the trajectory of yoga research toward better alignment and application to clinical practice. Yoga research reporting guidelines will improve transparency in research dissemination for application to practice, policy, replication, comparison, and summarization. Chapter 10 Yoga Therapy: An Overview of Key Research and the Underlying Mechanisms............................... 159 Eileen M. Lafer, University of Texas Health Science Center at San Antonio, USA & Beyond Wellness Aljezur, Portugal Margaret McCuiston, Southern California University of Health Sciences, USA & Seashore Healing, USA Ann Swanson, Maryland University of Integrative Health, USA & Kripalu School of Integrative Yoga Therapy, USA This chapter reviews the evidence supporting the efficacy of yoga therapy for wellness, and as an adjunct to standard care for a number of chronic conditions. The underlying mechanisms are explored, including the physiological and biochemical changes that have been observed in yoga practitioners. Yoga has been found to activate the relaxation response—a physiological state which reduces stress on bodily systems. Yoga leads to changes in gene expression, including decreases in the expression of genes involved in stress and inflammation. The positive effects of yoga therapy are interpreted through the lens of the biopsychosocial-spiritual model, which cultivates eudaimonic well-being and salutogenesis. Researchers attribute a wide range of yoga’s therapeutic benefits largely to its whole-person approach to well-being. Chapter 11 Yoga for Mental Health Disorders: Research and Practice.................................................................. 179 Bhargav Hemant, National Institute of Mental Health and Neurosciences, Bengaluru, India Rashmi Arasappa, National Institute of Mental Health and Neurosciences, Bengaluru, India Inbaraj G., National Institute of Mental Health and Neurosciences, Bengaluru, India Kaviraja Udupa, National Institute of Mental Health and Neurosciences, Bengaluru, India Shivarama Varambally, National Institute of Mental Health and Neurosciences, Bengaluru, India



This chapter is divided into four subsections: the first section of the chapter provides an update on current evidence for yoga therapy in common mental health disorders, the second section provides brief overview on neurophysiological abnormalities in psychiatric disorders and their relationship with psychological stress, the third section deals with summary of evidence for neurophysiological effects of yoga in mental health disorders, and the last section emphasizes on practical aspects of yoga therapy with details of clinically useful yoga practices for common mental health disorders. The mental health disorders covered in this chapter include depression, anxiety, schizophrenia, child and adolescent psychiatric disorders, substance use disorders, and geriatric psychiatric disorders. Chapter 12 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders: From Suffering to Thriving............................................................................................................................................... 199 Elin-Kristin Hem Olsen, Drammen District Psychiatric Centre, Norway Healing after trauma is a long-lasting process involving the body and mind. The neurobiological foundation of trauma calls for more bodily and sensational, bottom-up regulatory approaches. Yoga has been proposed as a possible adjunctive treatment for trauma, and for more complex forms where talk-therapy have been proven insufficient. This chapter will give a theoretical and evidence-based summary of all (to our knowledge) relevant empirical data on yoga as a treatment for trauma-related disorders and symptoms. Chapter 13 Clinicians Applying Yoga Principles and Practices in Pain Care: An Evidence-Informed Approach.221 Shelly Prosko, PhysioYoga, Canada Matthew J. Taylor, MyRehab, LLC, USA This chapter provides a practical introduction and guidelines for clinicians and researchers to appreciate the relationship between two enormous topics: pain and yoga. The emphasis is on practicality for both audiences with frequent citation to recent related publications and their deeper citations. The intention to fuel the imaginations of both clinicians and researchers occurs with a foundational understanding of how yoga, pain, and pain care are related. The authors’ experience in the clinic using yoga for people with pain and future areas for clinical and research inquiry are included. A list of techniques for the clinic, their clinical rationale, and two case reports ground the material and invite additional reflection for the psychophysiological applications of yoga in pain care. Chapter 14 Yoga and Lifestyle Factors in Heart Disease....................................................................................... 242 M. Mala Cunningham, University of Virginia, USA This chapter reviews cardiovascular risk factors and presents a comprehensive yoga-based intervention for impacting on heart health. Lifestyle interventions and yoga have been shown to positively impact on heart disease, stress levels, inflammation, vagal tone, and homeostasis. The Cardiac Medical Yoga Lifestyle Change program along with the five-point model for heart health (BREAD) provides heart patients with a comprehensive approach for impacting on their disease process and assists patients in developing the necessary strategies and skills for changing their lifestyle.



Chapter 15 Advances in Understanding the Use of Yoga as Therapy in Lymphedema......................................... 263 S. R. Narahari, Institute of Applied Dermatology, Kasaragod, India Madhur Guruprasad Aggithaya, Institute of Applied Dermatology, Kasaragod, India Terence J. Ryan, Department of Dermatology, Churchill Hospital, UK Lymphedema may be caused by dysfunction of the lymphatic system due to damage, block, hypoplasia due to genetic causes or by lymph overload. Lymphatic Filariasis is most prevalent and among the leading causes of disability. This chapter describes the improvements in three yoga protocols of integrative treatment, for lower limb, upper limb, and genital lymphedema. There are two sessions of yoga in our treatment protocol. Yoga helps to drain lymph through various mechanisms. Asanas focus on the dermal stretch, joint movement, muscle pumps, and muscle stretch and pranayamas on lung expansion. Joint movements and muscle contractions are designed to mimic nodal drainage. The yoga protocol also provides knee strengthening, gait correction, shoulder joint strengthening. Yoga asanas for comorbidities like hypertension, cardiovascular diseases, and arthritis and joint surgeries with movement restrictions are customized in the revised protocol. Yoga is an effective treatment in lymphedema and considerably improves the patient’s quality of life. Chapter 16 Yoga and Weight Management: A Narrative Review.......................................................................... 283 Sachin Kumar Sharma, Patanjali Research Foundation, India Sushma Pal, Patanjali Research Foundation, India Shirley Telles, Patanjali Research Foundation, India A comprehensive lifestyle intervention that is effective, yet safe, for weight loss is recommended for weight management. Yoga is one such intervention that includes (1) increased physical activity and (2) suggestions about healthy eating and other behavioral changes. With this, there have been attempts to assess the effectiveness of yoga for weight management. The present review aimed at (1) evaluating studies assessing the effects of yoga on obesity and (2) grade them according to standard grading methods. Twenty-three studies, of which nine were RCTs, were included. The rating of RCTs was between 70.6 and 94.2, which can be considered fairly good. However the present review shows that there are fewer studies with (1) adequate sample sizes, (2) study designs, (3) long term follow up, and (4) adequate reporting of adverse events to conclude that yoga is an effective and safe intervention for weight loss. Hence, further studies with the points described above are required to conclude the safety and efficacy of yoga for weight loss. Section 3 Applications and Mechanisms of Yoga Chapter 17 Overview of Yoga for Teenagers in the UK: The Rationale, Evidence Base, and the Application..... 305 Charlotta Martinus, Teen Yoga Foundation, UK Nicholas A. Kearney, Teen Yoga Foundation, UK This chapter outlines the current state of affairs of yoga provision for teenagers in the UK. It looks at the history of yoga in the UK among young people, the research and the application in school contexts, as well as prison and Pupil Referral Units. It outlines the results from the Hippocampus project, an EU-funded



2 year project evaluating the impact of yoga on disadvantaged youth in five countries. The chapter also takes a look at the current financial support for yoga in schools and the possibilities of taking it forward. Chapter 18 Yoga and Aging: Neurobiological Benefits......................................................................................... 317 Rui F. Afonso, Hospital Israelita Albert Einstein, Brazil Danilo F. Santaella, Hospital Israelita Albert Einstein, Brazil Elisa Harumi Kozasa, Hospital Israelita Albert Einstein, Brazil Governments and societies need to be prepared to confront population aging. Such preparation includes policies that can improve quality of life, functional capacity, and health of the general population, encouraging a more active and healthier lifestyle. Normal aging is associated with changes in brain structure and function, which may cause behavioral and cognitive impairments. It is important to understand which changes make some individuals healthier than others. Yoga has been associated with improved quality of life, cognition, and physical health as well as brain functional and structural changes. Chapter 19 Special Yoga for Children and Young People With Special Needs..................................................... 330 Jyoti Jo Manuel, Special Yoga, UK Special Yoga has been a provider of therapeutic yoga practices and training since its inception. The non-profit organization started with a yoga center in London, UK. The Special Yoga London center offered a therapeutic and nourishing, nurturing space for families of children with special needs who were welcomed with love, compassion, and totally nonjudgmentally with open arms. The work spread globally through the London-based trainings and the therapeutic yoga that was offered to children at the center and within education. This chapter will discuss my experience and understanding of the efficacy of yoga for children and young people, specifically those with cerebral palsy and autism and/or attention deficit hyperactivity disorder (ADHD). The author shares some of the practices, benefits, and case studies of two separate research studies that were undertaken by Special Yoga for each population. The author also shares other case histories of children that they have worked with. Chapter 20 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder........................ 347 Pauline Jensen, University of Sydney, Australia Research conducted in both the field of yoga and the field of behavioural disorders in children and adolescents leads to the speculation that the benefits of yoga practice demonstrated with respect to physiological, psychological, emotional, and psychosocial functioning may be applicable to the impairments evidenced in these areas in behavioural disorders. The intervention—20 weekly one-hour sessions of yoga—required a large commitment for the participants, their families, and the yoga instructor. The results suggested that yoga appears to exert its impact on stabilising the emotions (a secondary symptom) and reducing oppositional behaviour, frequently co-morbid (40%) with attention deficit hyperactivity disorder (ADHD). Given the limitations of this study, the results do indicate some significant changes in the behaviour of some of the boys with ADHD. In conclusion, yoga shows promise as a non-invasive, inexpensive, adjuvant treatment for boys with ADHD.



Chapter 21 Yoga: A Multi-Dimensional Therapeutic Approach to Autism Spectrum Disorder........................... 361 Kankan Gulati, National Institute of Mental Health and Neurosciences, Bengaluru, India Praerna H. Bhargav, National Institute of Mental Health and Neurosciences, Bengaluru, India Shalu Elizabeth Abraham, National Institute of Mental Health and Neurosciences, Bengaluru, India Hemant Bhargav, National Institute of Mental Health and Neurosciences, Bengaluru, India Autism spectrum disorder (ASD), a neurodevelopmental disorder, manifests as impairment in social communication an interaction with restrictive and repetitive patterns of behaviour. Yoga therapy, a mind-body intervention, employs a multi-dimensional approach to reduce psychological distress and bring balance and harmony at the levels of body, breath and mind through physical postures, breathing practices, chanting, and relaxation techniques, respectively, thus enhancing overall well-being. Various yoga studies have shown promise in improving symptoms of ASD by improvement in sensory processing, gross motor skills, balance and coordination, cognition, imitation skills, and the ability to connect in relationships. This chapter aims to provide an overview of the potential role of Yoga therapy in the management of ASDs with emphasis on future standardized yoga trials with robust methodology and long-term follow-ups to establish the clinical utility of Yoga therapy for the same. Also, a tentative yoga lifestyle module for ASD with necessary contra-indications and practical tips has been provided. Chapter 22 Yoga and Polyvagal Theory: Effects and Applications for Well-Being............................................... 391 Marlysa Sullivan, Maryland University of Integrative Health, USA This chapter explores yoga as a salutogenic intervention supportive of eudaimonic well-being with its wide-spread health effects for various patient populations. Autonomic nervous system regulation and resilience are considered as important meditators for the promotion of biopsychosocial health. Polyvagal theory offers a novel perspective on how underlying neural platforms support combined physiological, psychological, and behavioral states—inclusive of eudaimonic well-being. This chapter describes the convergence of neurophysiological ideas of neuroception, interoception, and neural platforms with yoga foundational concepts such as discriminative wisdom and the gunas. This translatory language of eudaimonic well-being and polyvagal theory offers a framework for yoga to be understood and integrated into current healthcare and research contexts while maintaining its unique perspective and foundational wisdom. Chapter 23 Traditional Frameworks of Well-Being and Modern Science............................................................. 410 Ganesh Mohan, Svastha Yoga and Ayurveda, Singapore This chapter describes how to connect traditional frameworks from yoga to modern science and systems theory in a way that is evidence-based, clinically testable, inclusive, and extensible. This way of systematization enables leveraging the evidence for yoga from across the spectrum of well-being modalities. The key in the presentation is the choice of an approach that is equally valid for both ancient and modern systems. The authors begin by examining the importance of a skill-based approach and layout how well-being systems can be built by mapping the skills and qualities of the organism. They take a selection of domains or categories under which skills from traditional and modern sources can



be analyzed: attention, transcendence, movement, breathing, devotion, cognition, and emotion. They conclude the chapter by presenting a larger map of well-being that includes more domains arranged from a clinical perspective. Chapter 24 The Science and Art of Contemporary Breathwork and Yoga............................................................ 422 Jim Morningstar, Transformations Incorporated, USA Jessica Dibb, Inspiration Consciousness School, USA This chapter provides evidence for the benefit of bringing yoga and its breathing practices into a working relationship with contemporary breathwork techniques for the betterment of both in certain circumstances. The history and developmental overlap of both are explored as well as their applications in the field of physical heath, psychotherapy, and spiritual growth. The benefits of both faster than normal and slower than normal breathing techniques are reviewed for the treatment of anxiety, depression, and addictions as well as the neurological correlates for breathwork and trauma recovery. A pilot study utilizing the combination of the two disciplines is reported. Chapter 25 An Exploration of Influence of Duration on Physiological Effects of Asanas.................................... 451 Debadutta Subudhi, Indian Institute of Technology, Madras, India M. Manivannan, Indian Institute of Technology, Madras, India Different body postures that can be maintained for a certain duration with awareness are called yogaasanas. These asanas matter flexibility, coordination, and strength, while the breathing practices and meditation sharpens the mind for better awareness and reduce anxiety and thus adds quality into life. Other beneficial effects might involve a reduction of stress, blood pressure, and improvements in resilience, mood, and metabolic regulation. The asanas performed regularly for a short duration in hours is well studied in the literature. However, when performed for a long duration continuously for several hours (40-hour yogathon), without food and sleep has a significant effect in regulating homeostasis. The homeostasis is accessed through cardio-respiratory and galvanic skin response changes. The study shows the physiological changes after the yogathon and compares it with effects on physiology due to short term yoga. It also emphasizes on the reduction of dependency on food, because of energy compensation through yoga-asanas. Compilation of References................................................................................................................ 475 About the Contributors..................................................................................................................... 562 Index.................................................................................................................................................... 574

xxii

Preface

Yoga is a mind and body intervention which has its origin in ancient India. Archaeological seals depicting a male deity seated in a posture comparable to the yoga half-lotus posture were excavated from the site of the Indus-Valley civilization (circa 3300 – 1500 B.C.) (www.harappa.com). Ancient sages of India who practiced yoga and followed its principles passed on their experiences to seekers and scholars. These scholars received instruction through word of mouth (Feuerstein, 2003) and direct teachings in traditional schools of the time (gurukul, where guru = teacher, kula = home, family). The profound personal experience of the ancient sages formed the basis for the classical texts of yoga. There are several yoga texts. The philosophy and principles of yoga are derived from complex teachings in the Vedas (circa 1500 B.C.) and the Upanishads (the written form, circa 1500 – 135 B.C.). The sage Patanjali (circa 900 B.C.) compiled the findings in a set of 196 aphorisms. The aphorisms describe the stages of yoga practice in a logical sequence from the start to the culmination of a seeker’s journey, in a state of ultimate, spiritual emancipation – kaivalya. The eight step path is called astanga yoga. In the beginning a yoga practitioner is recommended to observe basic principles of correct conduct in all dealings with the Self and with others [these are the yamas (= virtuous self-restraints) and niyamas (= virtuous observances)]. Having a clear idea and observance of the basic principles of Life, the practitioner is recommended to undergo physical and mental training through asanas (= yoga postures) to promote physical health and wellbeing. This training of the physical body also involves awareness of the breath and regulation of the mental state. Training the body through asanas aims to free the practitioner of physical concerns, while attempting to have control over the mental state. Regulation of the breath is an important part of yoga, since the breath is the only physical function which is under both voluntary and involuntary regulation. By voluntary breath regulation through pranayama (= voluntary regulated yoga breathing), a yoga practitioner controls the vital life principle called prana in Indian thought; chi in Chinese medicine. Breath regulation has a direct effect on controlling the mental state. Once a practitioner is able to regulate the breath at will, their next step is to regulate the senses, directing the attention inwards and away from the input from both external and internal sensations (pratyhara = withdrawal of the senses). These four steps, i.e., yama, niyama, asana and pranayama constitute transactional yoga (bahiranga yoga). The next three steps form contemplative yoga (antaranga yoga). After this, the practitioner is ready to attempt meditation. As described in yoga, meditation is practiced as two distinct steps, dharana (= meditative focused awareness) and dhyana (= effortless meditation). During the phase of dharana, a practitioner directs the attention to the object of meditation, which could be a word, an object, a symbol or anything else. This is a phase of focused attention which requires effort and commitment on the practitioner’s part. The next stage involves effortless absorption in the object of meditation and going beyond focusing, to expand the conscious state. This effortless, expansive and transcendent state is dhyana. This state is  

Preface

the meditation state according to yoga. A practitioner who can remain in a state of effortless absorption with expanded consciousness is ready for the ultimate spiritual emancipation (Samadhi = established in Supreme Consciousness), which is attained in successive stages. This demonstrates that yoga is the union of individual consciousness with Supreme consciousness, or of the individual soul (Jivatma) with the Supreme soul (Paratma). Yoga practice has gained in popularity the world over from the 1960s, when Transcendental Meditation (TM) was introduced to the world. TM is a twenty minute practice which helped novice practitioners reach a stage of transcendence (Rousseau, 1962). Before the 1960s, most yoga research was intended to examine extraordinary abilities of long-term yoga practitioners. This included voluntary regulation of physiological functions considered involuntary, such as slowing the heart rate at will or reducing the need for oxygen to minimal (Anand & Chhina, 1961; Karambelkar et al., 1968). Research on Transcendental Meditation showed that naïve practitioners could show benefits within days of committed practice. These benefits included health, therapy, education and even societal transformation (Dillbeck, 1990). In the last decade there has been an exponential increase in research on yoga the world over. Yoga techniques (especially, asanas, pranayama and meditation) are increasingly practiced for benefits such as better strength, increased flexibility, reduced stress, enhanced quality of life, enthusiasm for physical activity and healthy food choices (Nichols, 2019). While most yoga practitioners recognize that yoga practice does impact aspects of life other than physical function, there are some yoga studios which remain focused on the benefits related to physical fitness, weight regulation and muscle tone. This emphasis on physical fitness seems limited given the wide range of benefits a person can derive from a more comprehensive practice. To understand the types of yoga with better clarity, it is necessary to understand how systems of yoga are formed. There are several schools of yoga, some of them are named after the yoga master, whereas others are named after the sage or visionary whose vision is the basis of that branch of yoga. Styles of yoga which are named after a yoga master include Iyengar yoga, Kripalu yoga, and Sivananda yoga. Patanjali yoga is named after the great sage Patanjali (circa 900 B.C.). Vivekananda yoga is based on the ideas of the great Indian visionary, Swami Vivekananda (1863-1902), who described four paths of yoga, [viz. Karma yoga (the path through action), Jnana yoga (the path through intellect), Bhakti yoga (the path of unconditional devotion) and Raja yoga (eight steps of yoga)]. Sometimes the type of yoga is named after characteristics of the practice or based on the emphasis of the practice, such as Hatha yoga (Hatha = force), Vinyasa yoga (Vinyasa = transitions) or Kundalini yoga (named after the spiritual energy described at the base of the spine). There are also yoga practices which have been given names which suggest their benefits, such as Sudarshan Kriya yoga. Most of these types of yoga have definite positive effects on health demonstrated through experimental research. These are some of the popular forms of yoga which are practiced all over the world. In order to reach out to a wider population or to increase interest in yoga, there have been attempts to practice yoga in specific locations (e.g., on a beach or in a forest) or with particular farmyard animals (e.g., goats). These practices may seem far removed from the prescribed attitude of mental detachment, however any practice which is safe and helps poorly motivated people to increase their level of physical activity does have a part in the wide spectrum of health and wellness strategies. However, if yoga is to reach its full potential in health and as a therapy, there needs to be clarity about the efficacy and safety of yoga in accordance with guidelines for complementary and alternative medicine in health and therapy (Weiger et al., 2002). Research on yoga helps to understand the degree to which yoga is effective and safe to be used in the management of various conditions. The number of academicians involved in carrying out research on yoga has increased exponentially. A search for publications on yoga in PubMed using xxiii

Preface

‘yoga’ as the search word for different time periods from 1970 onward showed the following trend (the numbers are approximate as they depend on the date of the search): from 1970 to 1979, 166 publications; 1980 to 1989, 136 publications; 1990 to 1999, 225 publications; 2000 to 2009, 842 publications; 2010 to 2019, 3876 publications. Similar to yoga research, the number of yoga practitioners has increased over the years. In keeping with this trend, people who are trained to instruct others in yoga and to use yoga as a therapy have also increased. There are several levels of yoga courses. There are certificate courses of various durations which are regulated and certified by non-governmental organizations. Similarly such organizations also offer yoga therapy courses which enable a trained yoga therapist to teach yoga in a clinical setting, in consultation with a doctor. Universities in India and in other countries have started to offer graduate level and post graduate level courses in yoga. In India there is also a yoga and naturopathy degree course, which can be followed with three-year specialization in a particular area. There are also academic degrees in yoga, such as MPhil and Ph.D. In addition researchers in several disciplines (a few examples are psychology, education and different branches of medicine), take up topics related to yoga for their doctoral research. The present edited book would be useful to persons who have the different level of yoga instruction mentioned above. These courses all require understanding about how the best known and most widely practiced yoga techniques such as asanas, pranayamas and meditation influence physiology. The effects of yoga practice on physiology include the impact on cardio-respiratory systems, the autonomic nervous system and higher brain functions including cognition, among other functions. Most people undergoing training in yoga are interested in the health benefits of yoga, including promoting positive health, preventing and managing disease as a non-pharmacological, complementary treatment. It is also important to understand how practice may be modified for different age groups, especially children and older people. There are studies which have shown the benefits of yoga practice for children with behavioural disorders, children with attention deficit hyperactivity disorder and children with special needs due to other challenges. These topics have been covered in the present book, as detailed below. ‘Evidence-based perspectives on the psychophysiology of yoga and its applications’ is compiled as an edited volume to give readers an opportunity to access adequately supported information about yoga under three broad categories: (i) the physiology of yoga, (ii) applications of yoga in health and as a therapy and (iii) yoga practice to meet specific needs of different populations. These three categories are presented as three main sections with eight chapters in the first two sections and nine chapters in the third section. The chapters are authored by persons who can be considered as authorities in that particular area of knowledge about yoga based on their published research or based on their long involvement in imparting knowledge about yoga; in some cases the authors combine both types of experience. These authors are drawn chiefly from India, U.S.A., U.K. Europe, Brazil and Australia. The first section covers the physiological effects of yoga practice. Yoga has several features which make this mind and body practice different and distinct from physical exercise. One such feature is the emphasis on the breath in yoga. A verse from Hatha Yoga Pradipika (circa 300 A.D.) states in this verse, “when the prana (used interchangeably with breath) is irregular, the chitta (mind) is unstable, when the prana is without movement the chitta is stable” (Hatha Yoga Pradipika, Chapter II, Verse 2) (Saraswati & Saraswati, 1998). Hence in yoga practice, voluntary breath regulation is considered very essential to influence and quieten the mental state. For this reason, three of the eight chapters in the first section of the book cover different aspects of yoga voluntarily regulated breathing. The first chapter is a narrative review of the psychophysiological effects of yoga breathing, with details of some studies which report applications of yoga breathing. The second chapter covers detail about the neural regulation of breathxxiv

Preface

ing as well as other mechanisms of regulating respiration, describing the applications of these practices in health and wellbeing. The third chapter on yoga breathing is a summary of the cardiorespiratory and endocrine factors which regulate and are influenced by voluntary breath regulation in yoga. According to the eight steps of yoga (the ashtanga of the sage Patanjali, circa 900 B.C.), after quietening the mental state through breath regulation, a yoga practitioner is ready to practice meditation. There are two chapters on meditation. The first chapter on meditation covers the neurobiology of meditation with an emphasis on neuroimaging studies which have helped to understand the neurophysiology of meditation with spatial and temporal accuracy. The second chapter on meditation combines concepts from the traditional yoga texts with the neurobiology of meditation. This leads to the next chapter which covers concepts from the profound text, the Bhagavad Gita (circa 200 B.C.), with a presentation of concepts which can bridge science and spirituality. The first section ends with two chapters which cover mechanisms underlying the effects of yoga. The seventh chapter describes cellular changes which could explain the effects of yoga on different physiological pathways and systems. The eighth chapter describes the effects of yoga practice on cognition and other brain functions relevant to health. The second section of the book is about ‘Yoga as a therapeutic intervention’. Yoga influences the mental state and the physical functions, the latter both directly and indirectly by changing the mental state. This is the background for yoga being considered as a therapy for psychosomatic conditions. There has been research to support the therapeutic use of yoga for the last seventy years, at the least. Most often the evidence supports yoga as a safe add-on treatment for non-communicable diseases which are related to poor lifestyle choices. Research literacy is necessary to use yoga as an effective and safe therapeutic intervention. Yoga practiced in this way, would be better accepted and integrated in mainstream medicine (Stephens, 2017). For this reason the first chapter in this section is about standards in yoga research. The second chapter in this section explores the physiological and biochemical mechanisms underlying the use of yoga as a measure to enhance wellness and as a complementary treatment for several chronic diseases. The six remaining chapters of this section are related to using yoga as an add-on therapeutic intervention for different conditions. The third chapter of this section provides a comprehensive overview of the role of yoga in mental health disorders. The chapter includes the current evidence to use yoga as a therapy in common mental health disorders, the effects of yoga on neurophysiology in mental health disorders and practical aspects of yoga therapy with details of clinically useful yoga practices for common mental health conditions. The fourth chapter in this section is a theoretical and evidence-based summary of the relevant empirical research related to yoga as a treatment for trauma-related disorders. The fifth chapter in this section is about using yoga in the management of pain. The authors combine a practical approach interspersed with relevant citations. The conditions described in the third, fourth and fifth chapters are directly to mental health and neurology. The remaining conditions described in the sixth, seventh and eighth chapters of this section are more closely connected to lifestyle choices. The sixth chapter of this section reviews yoga for modifiable cardiac risk factors and presents the Cardiac Medical Yoga Lifestyle Change program, which provides patients with a comprehensive approach to adopt the necessary strategies to change their lifestyle. Lymphedema causes considerable physical and psychological discomfort. The seventh chapter of this section presents the authors’ novel approach which includes yoga to successfully manage lower limb lymphedema in cases of filariasis. The authors also mention that a comparable treatment program can be used to manage the lymphedema which follows surgery in many oncology cases. Finally, there is a chapter on yoga for weight management, which is relevant in this section, since obesity can predispose a person to many of the chronic conditions covered in the other chapters in this section. xxv

Preface

The third and last section of the book is ‘Applications and mechanisms of yoga’. This section has a unique blend of research and personal experience of the authors working with specific groups of people. The first chapter in this section examines the history of yoga in the U.K. among young people, the research and the application in school contexts, in prison and Pupil Referral Units, as well as the results of a two year project on yoga for disadvantaged youth. The second chapter in this section covers yoga for healthy aging especially in relation to brain functional and structural changes. The next two chapters deal with special needs of specific groups of children. This is an important area in which yoga has the potential to bring about beneficial change, since behavioural and physiological change is most feasible in this age group. In the third chapter in this section the author discusses the efficacy of yoga for children with cerebral palsy, autism and attention deficit hyperactivity disorder (ADHD), along with practises, two separate research studies and case histories. The fourth chapter discusses yoga as an intervention for children with ADHD and with conduct disorder enrolled at behavior school, presents research in this area. The fifth chapter discusses the potential of yoga to improve cognition, social deficits and overall quality of life of individuals with Autism Spectrum Disorder. While yoga has wide applications in education, across age groups and in rehabilitation, there are several theories which can further our understanding about additional possible benefits of yoga practice. The polyvagal theory (https://www.stephenporges. com) is increasingly quoted by experts in yoga as the mechanism for the biopsychosocial benefits associated with yoga practice. The sixth chapter in the section details the relevance of eudemonic wellbeing and the polyvagal theory to yoga in health. The seventh chapter discusses and yoga and wellbeing, connecting traditional descriptions from yoga to modern science in a way that is aimed to be clinically testable, inclusive, and extensible. Wellbeing is an important benefit seen with yoga practice. The last two chapters discuss the benefits of two allied and related practices. The eighth chapter of this section is related to Breathwork. Breathwork is a general term used to describe any type of therapy that utilizes breathing exercises to improve mental, physical, and spiritual health. This chapter provides evidence for the benefit of bringing yoga and its breathing practices into a working relationship with contemporary Breathwork techniques. The chapter includes several case histories which demonstrate the potential of Breathwork in managing mental health conditions. The final chapter in this section compares the physiological differences between a short and long duration of yogasana practice, with a brief review of previous research on yogasanas. Hence this evidence-based edited book is intended to provide clarity about the way yoga practice influences functioning in healthy persons and the benefits of yoga for different age groups, for people with special needs, as a therapy for lifestyle related diseases and in conjunction with allied mind and body interventions. Shirley Telles Ram Kumar Gupta

xxvi

Preface

REFERENCES Anand, B. K., & Chhina, G. S. (1961). Investigations on yogis claiming to stop their heart beats. The Indian Journal of Medical Research, 49, 90–94. Dillbeck, M. C. (1990). Test of a field theory of consciousness and social change: Time series analysis of participation in the TM-Sidhi program and reduction of violent death in the US. Social Indicators Research, 22(4), 399–418. doi:10.1007/BF00303834 Feuerstein, G. (2003). The deeper dimension of yoga: Theory and practice. Shambhala Publications. Harappa.com. (n.d.). Around the Indus in 90 slides. https://www.harappa.com/indus/33.html Karambelkar, P. V., Vinekar, S. L., & Bhole, M. V. (1968). Studies on human subjects staying on an air-tight pit. The Indian Journal of Medical Research, 56, 1282–1288. PMID:5711607 Nichols, H. (2019). What are the health benefits of yoga? https://www.medicalnewstoday.com/articles/326414 Porges, S. W., Doussard-Roosevelt, J. A., & Maiti, A. K. (1994). Vagal tone and the physiological regulation of emotion. Monographs of the Society for Research in Child Development, 59(2-3), 167–186. doi:10.1111/j.1540-5834.1994.tb01283.x PMID:7984159 Rousseau, D. M. (1962). Meditation: Easy system propounded by maharishi mahesh yogi. International Meditation Centre. Saraswati, S. M., & Saraswati, S. S. (1998). Hatha yoga pradipika. Yoga Publications Trust. Stephens, I. (2017). Medical yoga therapy. Children (Basel, Switzerland), 4(2), 12. doi:10.3390/children4020012 PMID:28208599 Weiger, W. A., Smith, M., Boon, H., Richardson, M. A., Kaptchuk, T. J., & Eisenberg, D. M. (2002). Advising patients who seek complementary and alternative medical therapies for cancer. Annals of Internal Medicine, 137(11), 889–903. doi:10.7326/0003-4819-137-11-200212030-00010 PMID:12458989

xxvii

xxviii

Acknowledgment

We express our deep gratitude to Swami Ramdev and Acharya Balkrishna, the founders of Patanjali Research Foundation, Haridwar, India, for their support and for giving us unstinted academic freedom. We are very grateful to the authors of individual chapters in the book for their valuable contributions. The many reviewers are gratefully appreciated for the time they spent to send their useful and constructive comments which were very essential to improve the quality of the book. We also thank the staff of Patanjali Research Foundation, especially Dipak Chetry who has been involved in the book from the very beginning. We are grateful to Kumar Gandharva and Deepak Pal for their assistance. We thank the publishers, IGI Global and especially thank Jan Travers and Maria Rohde of IGI global publishers for their support during the production of this book. This volume was put together during the extraordinary circumstances of the COVID-19 pandemic. Individually, we express our deep gratitude to specific people. Shirley Telles is grateful to Naveen Visweswaraiah for enduring academic and other support from 1994 onwards. Ram Kumar Gupta expresses gratitude to his beloved parents (Shri Bhagwan Sahay Gupta and Shrimati Nirmala Devi), family, teachers and friends for their consistent guidance and support. This edited book is based on the wisdom of all the great masters of yoga through the ages. We salute them. Shirley Telles Patanjali Research Foundation, India Ram Kumar Gupta Patanjali Research Foundation, India

 

Section 1

The Physiological Effects of Yoga Practice

1

Chapter 1

Psychophysiological Effects and the Applications of Yoga Breathing Practices Niranjan Kala Patanjali Research Foundation, India Dipak Chetry https://orcid.org/0000-0002-8117-6504 Patanjali Research Foundation, India Shirley Telles Patanjali Research Foundation, India

ABSTRACT Yoga is an ancient practice that originated in India and aims at purifying the mind for spiritual progress. In modern times yoga is widely practiced for general health and well-being as well as for therapeutic reasons. Voluntary breath regulation or pranayama has been given significant importance in traditional texts as well as by yoga masters. Research has shown beneficial effects of yoga breathing practices or pranayamas on neurocognitive, metabolic, respiratory, and autonomic functions, which are discussed in the chapter. The chapter also discusses the applications of these practices for the management of various clinical conditions as well as for alleviating psychological problems associated with particular illnesses. The beneficial effects of yoga breathing practices demonstrate the importance of these safe and cost effective non-pharmacological interventions for general health as well as for prevention and management of various diseases.

DOI: 10.4018/978-1-7998-3254-6.ch001

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

INTRODUCTION Yoga is an ancient science which originated thousands of years ago in India primarily for spiritual progress and self-enlightenment. Though the origin of yoga is difficult to trace, its presence can be found in the Indus valley civilization, one of the world’s oldest known civilizations which dates more than nine thousand years back (Sarkar et al., 2016; Dhyansky, 1987). Archaeological seals, such as that of a deity sitting in a yoga-like position suggests that yoga may have been practiced at that time (Dhyansky, 1987). The written descriptions of yoga can be found in the Vedas which are among the oldest written texts (Bhavanani, 2012). Though the Vedas were not in the written form originally, the compilations today date from Circa 10000 B.C. (Bhavanani, 2012). The Rigveda contains a number of verses describing yoga as a practice to control the mind (Bhavanani, 2012). The Atharvaveda talks about prana, the vital energy and the chakras, the energy centers in the pranic body (Bloomfield, 1897). After this the sage Patanjali wrote the yoga-sutras in which he systematically and exclusively compiled the wisdom of yoga in the form of aphorisms (Circa 900 B.C.) (Miller, 1996). This work remains as the most fundamental manual for yoga scholars as well as for yoga enthusiasts in the present time. There are 196 aphorisms in four chapters which define yoga as a means of self-enlightenment through the course of eight limbs (Miller, 1996). This is also known as the eight-fold path or astanga yoga and consists of the ethical precepts (yamas and niyamas), physical postures (asanas), breathing practices (pranayamas), control of the senses (pratyahara), deep concentration (dharana), meditation (dhyana) and the highest state of consciousness (samadhi). Out of the eight stages of yoga, voluntarily controlled breathing or pranayama, the fourth stage is the topic of this chapter. Etymologically, the word pranayama (in Sanskrit) means voluntarily slowing down and prolonging breathing. Ancient yoga masters realized the close connection between the breath and the mind hence these pranayamas or voluntarily regulated yoga breathing techniques are given special emphasis. This can be understood by a verse from the Hatha Yoga Pradipika (Circa 300 A.D.) which states “when the prana (used interchangeably with breath) is irregular, the chitta (mind) is unstable, when the prana is without movement the chitta is stable” (Hatha Yoga Pradipika, Chapter II, Verse 2) (Muktibodhananda, 2002). The relationship between the breath pattern and health is well established (Lieber & Mohsenin, 1992). Apart from this, breathing has been associated with higher brain functions. Zelano and colleagues (2016) found brain activity to be synchronized with the act of breathing during voluntary breath regulation measured through the recordings of intracranial electroencephalography (iEEG) in patients with epilepsy. In a separate study by the same authors and reported in the same article, nasal inhalation was found to facilitate recognition of facial expressions and recall of the objects shown (Zelano et al., 2016). When participants were instructed to breathe through the mouth these effects were not seen. Hence breathing can influence the brain and this is modulated by several factors such as nasal versus mouth breathing, inhalation in relation to exhalation and the nostril which is patent. In yoga the voluntary control of breathing involves one of the following aspects: change in the rate of breathing, change in the depth of breathing, manipulation of the nostril breathed through, including a period of breath holding and producing a sound during breathing. These alterations to the breathing pattern are the basis for various yoga breathing practices such as alternate nostril yoga breathing or anulom-vilom pranayama, bellows yoga breathing or bhastrika pranayama, high frequency yoga breathing or kapalabhati pranayama, as examples. The psychophysiological effects of these yoga breathing practices were discussed earlier (Telles & Singh, 2018) and are further elaborated here. 2

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

STUDIES ON PRANAYAMAS INVOLVING REGULATING THE NOSTRIL WHICH IS BREATHED THROUGH Nostril regulating breathing is performed by gentle pressure on the nose with the fingers or thumb of the dominant hand (Niranjanananda, 2004). Inhalation through the right nostril and exhalation through the left nostril exclusively is called suryabhedana pranayama (suryabhedana means sun-piercing breath in Sanskrit) with a variation called surya anuloma-viloma pranayama in which inhalation as well as exhalation is performed through the right nostril exclusively. In contrast, inhalation through the left nostril and exhalation through the right nostril exclusively is called chandrabhedana pranayama (chandrabhedana means moon-piercing breath in Sanskrit) with a similar variation called Chandra anuloma-viloma pranayama in which inhalation as well as exhalation is performed through the left nostril exclusively. Breathing through both nostrils alternately is called anuloma-viloma pranayama (alternate nostril breathing). which involves breathing through both nostrils alternately without retention of the breath. The practitioners use the thumb and ring finger of the right hand to manipulate the nostrils. The breathing practice begins by exhaling through the left nostril with the right nostril occluded; then inhaling through the left nostril; followed by exhaling through the right nostril with the left nostril occluded; then inhaling through the right nostril and exhaling through the left nostril. This is one complete cycle. Anuloma-viloma pranayama should not be confused with nadishodhana (subtle energy cleansing breathing) which is alternate nostril yoga breathing which involves breath holding which may follow either inhalation or exhalation. The effects of breathing through the right nostril, the left nostril, or through both nostrils alternately have been described in a specific yoga text called Swara yoga (Muktibodhananda, 1999). Swara yoga describes the effects of ida (left nostril patency), pingla (right nostril patency) and sushumna (both nostrils patent) on one’s body, mind and behavior. The text describes favorable acts to be performed when breath flows through a particular swara or nostril. For example breathing through left nostril is believed to have cooling effects and it is mentioned that one should perform acts which are not vigorous but are spiritually inclined when the left nostril is patent. These include stationary work, construction of a temple or well, consecration of a Deity, charity, entry into a newly constructed house and seed sowing (Shiva Swarodaya, Chapter V, Verses 102-113). It is also mentioned that a person should carry out activities requiring energy when the right nostril is patent because breathing through the right nostril is believed to be heat generating. The texts mention to carry out activities such as chanting of vira mantra (mantra for obtaining vigor and energy), journey, hunting, taming a horse, driving a chariot and holding a sword (Shiva Swarodaya, Chapter V, Verses 114-123). When prana (breath) flows through both nostrils equally, it is advised to remain silent, become introspective, concentrate the mind on Iswara (God) and perform yoga practices (Shiva Swarodaya, Chapter V, Verse 130). The effects of these breathing practices as described in Swara yoga texts have been partially studied in scientific studies. Right nostril yoga breathing has been shown to increase oxygen consumption immediately after 45 minutes of practice as well as after a period of one month (Telles et al., 1996; Telles et al., 1994). Along with this, right nostril breathing also caused an increase in peripheral vasoconstriction, increased systolic blood pressure and heart rate suggesting sympathetic activation (Telles et al., 1994; Telles et al.,1996). Similar findings were later reported when the practice of suryabhedana pranayama increased systolic, diastolic and mean blood pressure (Raghuraj & Telles, 2008). Chandra bhedana pranayama caused a reduction in systolic and mean blood pressure while anuloma-viloma pranayama decreased the systolic and diastolic blood pressure in twenty yoga experienced individuals 3

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

(Bhavanani et al., 2014). These findings support the energizing and heat-generating effects of suryabhedana pranayama and relaxing effects of chandrabhedana pranayama and anuloma-viloma pranayama described in yoga texts. It is described that the body can be purified and complete health can be achieved within three months through the practice of nadishodhana pranayama and anuloma-viloma pranayama (Hatha Yoga Pradipika, Chapter II, Verses 10, 20). The imbalance between the sympathetic and parasympathetic divisions of the autonomic nervous system indicates significant morbidity and mortality, and nadishodhana pranayama and anuloma-viloma pranayama appear to correct the imbalance. In an earlier study where twenty healthy males were trained to practice nadishodhana pranayama, they showed a reduction in systolic blood pressure and heart rate after four weeks (Bhargava et al., 1988). Interestingly, these changes remained when the volunteers were instructed to hold their breath until the breaking point. This suggests that nadishodhana pranayama practice reduces sympathetic activation in the presence of a physical stressor (i.e., breath holding). Various studies showed a trend of parasympathetic activation as well as sympathetic withdrawal after alternate nostril yoga breathing (Bhagat et al., 2017; Telles et al., 2014; Raghuraj & Telles, 2008). However, a study on non practitioners of yoga reported different results with increased low frequency (LF) and decreased high frequency (HF) and increased LF/HF ratio after 6 and 12 minutes of alternate nostril yoga breathing (Subramanian et al., 2016). Apart from the effects on vagal modulation, uninostril breathing practices have shown lateralized effects on the cerebral hemispheres. Earlier studies which were based on electroencephalogram recordings as well as on performance in hemisphere-specific tasks suggested that forced uninostril breathing activates the contralateral cerebral hemisphere (Shannahoff-Khalsa et al., 1991; Werntz et al., 1987). A recent study supported this result with changes in brain hemodynamics that showed brain oxy-hemoglobin levels increased in the left prefrontal cortex during the practice of right nostril breathing (Singh et al., 2016). The study also reported a trend of increased oxy-hemoglobin levels in right prefrontal cortex after practice of left nostril breathing. In another study right nostril breathing and alternate nostril breathing but not left nostril breathing improved scores in an attention related task (Telles et al., 2007). When yoga-based uninostril breathing practices were compared for performance in verbal and spatial memory tasks in 108 school children, the spatial memory task scores increased after left nostril breathing, right nostril breathing as well as alternate nostril breathing (Naveen et al., 1997). The lateralized effects of suryabhedana pranayama were shown by the recordings of middle latency auditory evoked potentials. There was an increase in the peak amplitudes of Na wave and Nb wave on the right cerebral hemisphere during suryabhedana pranayama suggesting better neural allocation at the right cerebral hemisphere (Raghuraj & Telles, 2004). The increase in peak amplitude suggested an increased recruitment of neural resources at the thalamic medial geniculate and Heschl’s gyrus on the right hemisphere during the practice of suryabhedana (Polich, 1999). Nostril manipulative breathing has also been shown to improve performance in several psychomotor tasks. In fifteen yoga practitioners, alternate nostril yoga breathing reduced time taken to complete a vigilance related task as an immediate effect along with a simultaneous reduction in systolic and mean arterial blood pressure (Telles et al., 2017). Another study on fifty male volunteers showed improved scores in shape and size discrimination task after alternate nostril yoga breathing and decreased state anxiety after breath awareness and quiet sitting (Telles et al., 2019). Similarly, studies reported improved performance in a letter cancellation task after anuloma-viloma pranayama as well as increased verbal and spatial memory scores after anuloma-viloma pranayama, suryabhedana pranayama and chandrabhedana pranayama (Telles et al., 2007; Garg et al., 2016). These findings can be explained by another 4

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

study in which the P300 task performance was improved after anuloma-viloma pranayama (Telles et al., 2013). The P300 is an event related potential which is generated when a person pays attention to a specific stimulus and differentiates that from the other (Polich, 1999). Hence the P300 is a objective measure of selective attention. The P300 peak amplitudes were increased at frontal, central and parietal scalp sites immediately after 18 minutes of anuloma-viloma pranayama (Telles et al., 2013). Other than this, anuloma-viloma pranayama practiced for ten days improved hand grip strength in school children (Raghuraj et al., 1997). Effect of alternate nostril breathing to reduce experimentally induced anxiety has also been reported (Kamath et al., 2017).

STUDIES ON PRANAYAMAS INVOLVING ALTERATION IN THE DEPTH OF BREATHING Bhastrika pranayama (bhastrika = bellows, in Sanskrit), is a yoga breathing practice that involves inhalation and exhalation with force and increased depth of breathing. In traditional texts bhastrika pranayama is described to balance the three constitutional types (doshas) described in Ayurveda (i.e., vata, pitta and kapha) as well as to be effective in increasing the digestive fire or metabolism (Hatha Yoga Pradipika, Chapter II, Verse 65). Earlier studies have reported increased oxygen consumption during this practice consistently by up to 35 percent (Miles & Behanan, 1934; Miles, 1964). The effect of bhastrika pranayama on cardiopulmonary functions has also been studied. In 76 adults randomized to bhastrika or stretching for four months, bhastrika caused an increase in maximum expiratory and inspiratory pressure as well as a reduction in the low frequency (LF) power and low frequency/high frequency (LF/HF) ratio of the heart rate variability (HRV) (Santaella et al., 2011). These components of HRV viz., LF and LF/HF ratio are suggestive of cardiac sympathetic modulation and sympathovagal balance respectively. A report is also there which shows that bhastrika pranayama can increase the heart rate, rate pressure product and double product (Madanmohan et al., 2005). Another study counters this in which slow paced bhastrika pranayama reduced systolic and diastolic blood pressure (Pramanik et al., 2009). These results indicate that the breath rate at which bhastrika pranayama is practiced has certain influences on the autonomic balance. A recent study on 26 yoga practitioners found that yogic slow breathing does not appear to augment cardiac vagal control (Bertisch et al., 2017). Bhastrika pranayama also has an effect on the reaction time. Typically reaction time is the time taken by an individual to respond to a specific stimulus or event (Jensen, 2006). In a study conducted on healthy males, 18 minutes of bhastrika pranayama decreased the anticipatory responses suggesting improved reaction time (Telles et al., 2013). A similar study was done on 25 females which reported no significant changes after bhastrika, however an improvement in reaction time occurred after breath awareness and quiet sitting (Telles et al., 2018). These findings suggest that bhastrika pranayama has gender-specific effects on reaction time. Apart from this, pulmonary functions were studied on thirty males in a randomized controlled trial (Budhi et al., 2019). There were significant improvements in all variables including forced vital capacity, forced expiratory volume in the first second, peak expiratory flow rate and maximum voluntary ventilation after bhastrika pranayama for one month as well as an increase in peak expiratory flow rate and maximum voluntary ventilation after physical activity with magnitude of change lesser than bhastrika. Similar findings were reported in another study which was conducted on swimmers in which bhastrika pranayama along with other pranayamas improved the pulmonary parameters (Hakked et al., 2017). 5

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

A RCT on 100 males examined the effects of slow breathing or bhastrika on perceived stress, body mass index, waist-hip ratio, heart rate and blood pressure (Naik et al., 2018). Heart rate, systolic and diastolic blood pressure and perceived stress decreased following 12 weeks of bhastrika with no significant changes in body mass index and waist-hip ratio. Bhastrika pranayama has a variation called mukhabhastrika pranayama which involves practicing bhastrika with mouth-breathing (mukha = mouth and bhastrika = bellows). Mukhabhastrika also appears to have similar trends as those of bhastrika pranayama. In twenty-two healthy school going boys, mukhabhastrika reduced visual and auditory reaction after nine rounds of practice (Bhavanani et al., 2003). Another study reported that mukhabhastrika can reduce sympathetic activity and increase parasympathetic activity by reducing basal heart rate, fall of systolic blood pressure on posture variation and increasing the Valsalva ratio (Veerabhadrappa et al., 2011). Recently, a RCT was conducted on 30 healthy adults aged between 18 to 40 years to assess the effect of bhastrika on fMRI, states of anxiety and positive and negative affect (Novaes et al., 2020). After 4 weeks of pranayama the states of anxiety and negative affect were significantly decreased. Also there were changes in the activity of brain regions involved in emotional processing, particularly the amygdala, anterior cingulate, anterior insula, and prefrontal cortex.

STUDIES ON PRANAYAMAS WITH BREATH RETENTION Certain yoga breathing practices include a period of breath holding which could follow either inhalation or exhalation or both. This is called kumbhak (kumbhak = a pot in Sanskrit). Generally, kumbhak pranayamas are practiced along with specified physiological locks or bandhas which are three in number i.e., Jalandhar bandha or chin lock, uddiyan bandha or abdominal lock and mula bandha or perennial lock (Nagendra, 1998). The duration of breath holding can also differ in different pranayamas. Generally breath holding is practiced by experienced yoga masters and requires steady and careful learning as described in this verse, “just as lions, elephants and tigers are controlled, so the prana should be controlled through gradual practice, otherwise the sadhaka (practitioner) is destroyed.” (Hatha Yoga Pradipika, Chapter II, Verse 15). An earlier study examined the oxygen consumption during pranayamas with breath holding and also compared the effects of the duration of breath holding (Telles & Desiraju, 1991). The results were interesting including a 19 percent decrease in oxygen consumption during the practice of long episodes of breath holding as well as a 56 percent increase in the oxygen consumption during the practice of short episodes of breath holding. In addition, kumbhak reduced the heart rate, stroke volume and cardiac output in 39 healthy adults following 20 minutes of practice (Saoji et al., 2018). In another study by the same authors, the practice for the same duration improved reaction time (Saoji et al., 2018). Pranayama in a ratio of 2:1:2:1 for inhalation:breath-hold:exhalation:breath-hold is called savitri pranayama. A study found significant changes after a combination of pranayamas including savitri pranayama on respiratory sinus arrhythmia suggesting cardiovascular effect of these pranayamas (Bhavanani et al., 2016).

6

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

STUDIES ON PRANAYAMAS INCORPORATING ALTERATIONS IN THE BREATHING FREQUENCY Most yoga breathing practices alter the rate of breathing by slowing down the breath frequency, kapalabhati pranayama however increases the breath rate by up to 60 to 120 breaths per minute. This technique has also been described as a cleansing process, or a kriya in Sanskrit. Kapalabhati is described as a pranayama when practiced at slower rates i.e., £ 60 breaths per minutes. The name kapalabhati consists of two Sanskrit words, kapala which means “forehead” or “cranium” and bhati which means shining, splendor or knowledge (Muktibodhananda, 2002) which indicates the possible benefits of kapalabhati on brain function. This is supported by a study in which the practice of kapalabhati was reported to improve the performance in a cancellation task that assesses selective attention, repetitive motor responses and visual scanning (Telles et al., 2008). Similar results were reported in a different study in which 18 minutes of practicing kapalabhati significantly reduced the degree of optical illusion perceived (Telles et al., 2011). Ten minutes of kapalabhati breathed at 1.0 Hz improved visual discrimination and finger dexterity in ninety four healthy volunteers (Telles et al., 2012). Though a recent study found no significant change after kapalabhati on performance in a Corsi-block tapping task, however an improvement in scores was reported after breath awareness (Gupta et al., 2019). Another study on functional near infrared spectroscopy found that kapalabhati practiced at the rate of 1.0 Hz does not cause changes in cerebral blood flow to the prefrontal region (Telles et al., 2016). The attention related effects were also seen in sixty one pre-teen children. Children showed improved performance in an attention based letter cancellation task as well as decreased levels of anxiety after kapalabhati (Telles et al., 2019). The improvements in performance in specific tasks following kapalabhati can be explained by the findings of a study on the P300 task (Joshi & Telles, 2009). The study found that P300 peak latency decreased after one minute of kapalabhati practiced at 2.0 Hz. A decrease in the P300 peak latency is suggestive of a decrease in time taken to complete the attention task hence faster neural processing. In addition, the P300 peak amplitude was increased after one minute of breath awareness suggesting increased neural allocation to complete the task. Apart from this, kapalabhati is reported to have autonomic and cardiovascular effects. In an earlier study, 24 volunteers practiced kapalabhati for 15 minutes at 2.0 Hz and their cardiovascular variables were observed including the electrocardiogram evaluated by spectral analysis of time series (Stancák et al., 1991a). During kapalabhati, the respiration rate was modulated by a 0.1 Hz rhythm in 82 percent of the experiments, which was also present in R-R intervals and blood pressure. Another study was conducted by the same authors in seventeen yoga experienced subjects (Stancák et al., 1991b). The effects of kapalabhati at a breath frequency of 2.0 Hz were examined before, three times during and after the practice for 15 minutes. During kapalabhati, the heart rate was increased by 9 beats per minute, systolic blood pressure by 15 mmHg and diastolic blood pressure by 6 mmHg. Also, the frequency bands of R-R interval variability as well as the baroreceptor-cardiac reflex sensitivity were reduced during the practice. Similar results were reported when the practice was for shorter duration. A study in which kapalabhati at 2.0 Hz was practiced for one minute the low frequency power and the low frequency/high frequency power increased along with a decrease in the high frequency power (Raghuraj et al., 1998). Findings of these studies suggest that kapalabhati can decrease the cardiac vagal modulation. Another study reported comparable results in which kapalabhati was practiced at slower rate i.e., 60 breaths per minutes (Telles et al., 2011). The study reported a significant decrease in the mean RR interval, NN50 and pNN50 components of heart rate variability during 15 minutes of kapalabhati. 7

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

STUDIES ON PRANAYAMAS INVOLVING PRODUCTION OF A SOUND WITH BREATHING Ancient yoga texts have given emphasis on sound and associated resonance and this construct in yoga is known as nadanushandhan (nad = sound, anusandhan = research in Sanskrit) (Hatha Yoga Pradipika). Several yoga breathing techniques involve production of a specific sound while breathing such as bhramari pranayama (bhramari = bumble bee in Sanskrit) and udgeeth pranayama or chanting of ‘OM’, a syllable with sacred connotations. Bhramari pranayama involves producing a vibrating sound like buzzing of a bumble bee while exhaling through the nostrils. A pre and post study reported that bhramari pranayama can generate highfrequency paroxyxmal gamma waves during humming (Vialatte et al., 2008). The effects of bhramari pranayama were also reported on the cardiovascular system. Sixty healthy adolescents were randomized to 45 minutes of bhramari pranayama or equal duration of normal breathing (Kuppusamy et al., 2016). A significant reduction was reported in heart rate, pulse pressure, mean arterial pressure, rate pressure product and double product after bhramari pranayama. In addition, short durations such as 5 minutes of the practice was also associated with a decrease in diastolic and mean blood pressure (Pramanik et al., 2010). In contrast, a recent study reported signs of parasympathetic withdrawal including increase in heart rate and low frequency spectrum of heart rate variability and decrease in high frequency spectrum of heart rate variability (Nivethitha et al., 2017). The changes seen were during the practice of bhramari pranayama which revert to normal after practice. Effects on pulmonary function were studied in ninety adolescents in a randomized controlled trial (Kuppusamy et. al, 2017). Bhramari pranayama practice improved all pulmonary function parameters including forced vital capacity (FVC), forced expiratory volume in first second (FEV1), FEV1/FVC and maximum voluntary ventilation. Apart from this, practicing bhramari pranayama for 10 minutes was also shown improve the reaction time and response inhibition in 31 male students (Rajesh et al., 2014).

STUDIES ON PRANAYAMAS INVOLVING BREATHING THROUGH THE MOUTH Though, breathing through the mouth is not preferred generally, certain yoga breathing practices involve inhaling through the mouth and exhaling through the nose. Examples of such breathing techniques are sheetali pranayama (sheetali comes from the Sanskrit root sheet, which means “cold” or “frigid) and sitkari pranayama (in Sanskrit, sitkari means ‘sipping’ or ‘hissing’). These breathing techniques are generally known as cooling pranayamas, possibly because these involve inspiration through the mouth which prevents the warming effect on air when it passes through the nostrils. The benefits of sheetali and sitkari are mentioned in Hatha Yoga Pradeepika such as control of hunger, thirst, sleep and laziness, relieving stomach related diseases, fever, excess bile and eliminating poison from the body (Hatha Yoga Pradipika, Chapter II, Verses 55, 58). Scientific studies have not studied the effects of these pranayamas widely, though in experience of the yoga practitioners these practices are considered beneficial for oral health and may help decreasing the peripheral temperature of a person with a fever (Ramdev, 2005). However an EEG based study reported an increase in theta and alpha waves in frontal and occipital regions and an increase in delta waves in the frontal region during the practice of sheetali and sitkari pranayamas (Thanalakshmi et al., 2014). The study also reported reduced beta frequencies during the

8

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

practice. There are some scientific reports which suggest the therapeutic benefits of these practices and they are discussed further in the chapter.

STUDIES ON PRANAYAMAS INVOLVING BREATHING WITH A PARTIALLY CLOSED GLOTTIS There is a yoga breathing practice known as ujjayi pranayama (ujjayi = victorious in Sanskrit), which involves voluntarily constricting the glottis along with production of a sound with breathing. It is mentioned that this practice removes phlegm from the throat and stimulates the digestive fire (Hatha Yoga Pradipika, Chapter II, Verse 52). This suggests the metabolic effects of ujjayi pranayama which are shown by earlier studies in which ujjayi pranayama increased oxygen consumption by 12 to 33 percent (Miles et al., 1964). Another study reported that ujjayi pranayama increased the oxygen consumption by 52 percent when practiced along with short episodes of breath holding and decreased the oxygen consumption by 19 percent when practiced along with long episodes of breath holding (Telles & Desiraju, 1991). One study assessed the effects of ujjayi along with kapalabhati and bhastrika pranayama on lower limb isokinetic and iso-inertial power and found no significant changes in power output (Wooten et al., 2018).

THERAPEUTIC APPLICATIONS OF YOGA BREATHING PRACTICES Therapeutic effects of yoga have been examined with greater interest in recent years. Though the practice of yoga breathing techniques is generally considered harmless, one should practice them with adequate knowledge and under proper guidance. A verse from Hatha Yoga Pradipika would be worth quoting which states “by the proper practice of pranayama, all diseases can be eradicated; and by the improper practice, all diseases are caused” (Hatha Yoga Pradipika, Chapter II, Verse 16). While the use of combined yoga interventions for the management of various diseases is studied widely, there are relatively less studies pertaining to yoga breathing practices exclusively. Two studies in the late 1980s by Virendra Singh and colleagues examined the effects of pranayama on patients with bronchial asthma (Singh, 1987; Singh et al., 1990). The authors used an apparatus called the ‘pink city lung exerciser (PCLE)’ which could control the inhalation-exhalation ratio hence at least in this aspect, resembling pranayama breathing (Dr. Virendra Singh is from Jaipur in India, where Jaipur is known as the Pink City, due to the dominant color scheme of its buildings. The PCLE comprised of a cylindrical device through which the person was asked to breathe. The device had holes that were designed in a way that a ratio of 1:2 for inhalation and exhalation was maintained while breathing through it (Singh, 1987). A device which looked exactly the same but did not alter the respiration ratio was used as a placebo. The first study reported significant increases in peak expiratory flow rate of the 12 patients after a two week intervention (Singh, 1987). Another study was more important which showed improved airway reactivity to histamine more with use of the PCLE than with the placebo device (Singh et al., 1990). The study reported an increase in the histamine provocation dose causing a 20 percent decline in FEV1 (PD20) by 1.53 (log2 units) in the PCLE group and 0.96 (log2 units) in the placebo group (Singh et al., 1990). Though the pink city lung exerciser had certain limitations, it was a unique method because it allowed a placebo controlled intervention to be compared with yoga breathing. Another study included 9

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

one more breath-based intervention viz., Buteyko breathing along with PCLE and PCLE based placebo breathing to assess pulmonary functions in 69 asthmatic patients (Cooper et al., 2003). The symptoms were reduced in the Buteyko breathing group but remained relatively stable in the PCLE and placebo groups. Use of bronchodilator was reduced in the Buteyko breathing group by two puffs per day at six months while there was no change in the other two groups. One important component of yoga breathing is awareness towards breathing along with the breath alterations (Ramdev, 2005). Breathing through the PCLE resembles yoga breathing by altering the inhalation:exhalation ratio but lacks the mental component of breath awareness essential in yoga breathing practice. This could restrict the potential effects of yoga breathing to be seen through the PCLE breathing. Findings of another study support this in which 120 asthmatic patients were divided into Buteyko breathing group, pranayama group and control group (Prem et al., 2013). The outcomes were Asthma Quality of Life Questionnaire, Asthma Control Questionnaire and pulmonary function test. The Buteyko breathing group showed 26.1 percent improvement in the quality of life than the other two groups, while pranayama showed 14.3 percent improvement in the quality of life compared to the control group. Effects of dirgha pranayama (the three-part breathing exercise, dirgha means long in Sanskrit) were observed in patients with chronic obstructive pulmonary disease (COPD). Dirgha pranayama is a slow rhythmic breathing in which participants are instructed to breath in slowly through their nose to fill first the bottom of their lungs, then the middle of their lungs, and finally the top of their lungs. The exhalation can be in the same sequence or opposite depending upon the comfort of the practitioner. A study on 43 patients with COPD were randomized to dirgha pranayama or control group and their 6-minute walk distance, pulmonary function test, markers of oxidative stress and systemic inflammation, and measures of dyspnea and quality of life were measured (Kaminsky et al., 2017). The 6-minute walk distance increased in the pranayama group by 9 percent along with improvements in the inspiratory capacity and air trapping by 6.2 and 12.5 percent respectively. Effects of bhramari pranayama were assessed in chronic sinusitis in 60 patients using the Sino-Nasal Outcome Test (Abishek et al., 2019). The scores of the Sino-Nasal Outcome Test improved by 36.7 percent after 4 weeks and continued until 12 weeks of assessment. Apart from diseases of the respiratory system, certain yoga breathing practices were assessed in hypertensive patients. In a study with 90 hypertensive patients, alternate nostril yoga breathing decreased the systolic blood pressure by 4.2 percent and diastolic blood pressure by 1.56 percent (Telles et al., 2013). The scores in Purdue pegboard task were also increased significantly after the practice. Effects of sheetali and sitkari pranayama on blood pressure and cardiovascular parameters were assessed in two separate studies containing twenty and sixty hypertensive patients respectively (Kumar et al., 2018; Shetty et al., 2017). Kumar et al. (2018) reported 8.6 percent decrease in systolic blood pressure, 11.4 percent decrease in diastolic blood pressure, 8.9 percent decrease in heart rate, 8.1 percent decrease in pulse pressure and 9.1 percent decrease in double product after 20 minutes of sheetali and sitkari pranayama. Shetty et al. (2017) also reported similar findings including 10.9 percent decrease in systolic blood pressure and 22.8 percent decrease in respiratory rate along with 9.9 percent increase and 86 percent increase in NN50 and HF-power component of heart rate variability after sheetali and sitkari pranayama. In another study, slow and fast breathing practices were compared in 60 patients with essential hypertension (Mourya et al., 2009). Alternate nostril breathing was chosen as the slow breathing practice and 4-5 epochs of deep high frequency breathing for 15 minutes was chosen as the fast breathing. The blood pressure decreased longitudinally over a period of three months with both slow and fast pranayama while there was 70.9 percent increase in standing-to-lying ratio, 13.9 percent increase in immediate heart rate response to 10

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

standing, 4.5 percent increase in heart rate variation with respiration ratio, 33 percent increase in the hand grip test and 62.5 percent increase in the cold pressor test only in the slow breathing practice. This suggests that slow breathing such as alternate nostril breathing have better effect than fast breathing. Another study supports this statement in which sukha pranayama (sukha means happiness in Sanskrit) was intervened to examine the heart rate and blood pressure (Bhavanani et al., 2011). Sukha pranayama is a slow breathing practice done at the rate of six breaths per min by using a timed count of 5 for inhalation and exhalation. Following sukha pranyama for 5 minutes there was a significant reduction in heart rate, systolic pressure, pulse pressure, mean arterial pressure, rate-pressure product, and double product. The study reported a 3.4 percent decrease in heart rate, 6.7 percent decrease in systolic blood pressure, 12.1 percent decrease in pulse pressure, 7.4 percent decrease in mean arterial pressure and 9.1 percent decrease in double product after 5 minutes of intervention (Bhavanani et al., 2011). In another study, 60 females with premenstrual syndrome were randomized to three groups viz., anuloma-viloma pranayama, yoga asanas and control (Sharma et al., 2013). Their blood pressure, heart rate, electromyogram, galvanic skin response, respiratory rate and peripheral temperature were measured seven days before and after starting of the menstruation for three consecutive menstrual cycles. The anuloma-viloma pranayama group showed 17.7 percent decrease in heart rate, 10.7 percent and 8.4 percent decrease in systolic and diastolic blood pressure, 44 percent decrease in electromyogram and 13.8 percent decrease in galvanic skin resistance after the third menstrual cycle. Similar findings were also reported in yoga asana group. Yoga based deep breathing as an add-on to Scaling and Root Planning (SRP) was administered in 30 patients with chronic periodontitis and was compared with SRP alone for a period of three months (Mahendra et al., 2017). The pranayama group showed 52.2 percent improvement in the symptoms of periodontitis after three months compared to SRP alone. The presence of three Red Complex Microorganisms (RCM) viz., Treponemadenticola, Porphyromonasgingivalis and Tannerella forsythia was also decreased in the pranayama group by nearly 77 percent. Apart from this, a study showed 52.27 percent increase in the net scores of the Braille version of sixletter cancellation test by 19 visually impaired individuals after bhramari pranayama suggesting better attention after yoga breathing which is in line with other reports on healthy persons (Pradhan et al., 2018). Yoga breathing appears to be effective for psychological challenges faced by patients such as anxiety and poor emotional aspects. Anxiety was measured in 80 patients undergoing coronary angiography before and after five minutes of sukha pranayama (Bidgoli et al., 2016). The anxiety score decreased significantly after sukha pranayama from 53.37 to 40.75. In another study 160 patients with breast cancer who were undergoing radiation therapy were randomized to pranayama or control group and their emotional aspects such as impatience, tiredness, worry, anxiety and frustration were assessed (Chakrabarty et al., 2016). Pranayama included bhramari, sheetali and nadishodhana for 20-30 minutes in morning and evening 6 days per week for 6 weeks. Pranayama group resulted in about 70 percent reduction in these negative emotions than the control group. Yoga breathing has also been found effective on cravings in abstaining smokers. 96 smokers were randomized to yoga breathing or video control groups (Shahab et al., 2013). Yoga breathing included deep breathing and alternate nostril breathing for 10 minutes while the video control group was shown a breathing exercise video for 10 minutes. The symptoms including strength of urges, cravings and desire to smoke were immediately alleviated after yoga breathing compared to video control group.

11

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

CONCLUSION Yoga breathing practices have drawn scientific attention on discovering the various health related benefits. Numerous studies have shown significant health effects of these practices which include, but are not limited to, improvement in cognitive functions, improvement in lung functions, metabolism, better performance in fine motor tasks and modulation of the autonomic nervous system towards a physiologically relaxed state. In addition, yoga breathing practices are found to be effective for managing symptoms and associated psychological problems in several diseases. Though the yoga breathing practices are easy to learn and are generally (but not completely) harmless, they are recommended to be practiced in competent supervision only, especially when practiced for therapeutic purpose. The body of literature is continuously growing in the field of yoga research, nevertheless, robust trials are required to conclusively investigate and understand the effects associated with yoga breathing practices. So far, the available literature suggests that these yoga breathing practices can be practiced for general health and wellness, as well as an add-on therapy in many diseases.

REFERENCES Abishek, K., Bakshi, S. S., & Bhavanani, A. B. (2019). The efficacy of yogic breathing exercise bhramari pranayama in relieving symptoms of chronic rhinosinusitis. International Journal of Yoga, 12(2), 120–123. doi:10.4103/ijoy.IJOY_32_18 PMID:31143019 Bertisch, S. M., Hamner, J., & Taylor, J. A. (2017). Slow yogic breathing and long-term cardiac autonomic adaptations: A pilot study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 23(9), 722–729. doi:10.1089/acm.2016.0074 PMID:28418684 Bhagat, O. L., Kharya, C., Jaryal, A., & Deepak, K. K. (2017). Acute effects on cardiovascular oscillations during controlled slow yogic breathing. The Indian Journal of Medical Research, 145, 503–512. PMID:28862183 Bhargava, R., Gogate, M. G., & Mascarenhas, J. F. (1988). Autonomic responses to breath holding and its’ variations following pranayama. Indian Journal of Physiology and Pharmacology, 32, 257–264. PMID:3215678 Bhavanani, A. B., Madanmohan, & Udupa, K. (2003). Acute effect of mukh bhastrika (a yogic bellows type breathing) on reaction time. Indian Journal of Physiology and Pharmacology, 47, 297–300. PMID:14723315 Bhavanani, A. B., Raj, J. B., Ramanathan, M., & Trakroo, M. (2016). Effect of different pranayamas on respiratory sinus arrhythmia. Journal of Clinical and Diagnostic Research: JCDR, 10, CC04–CC06. doi:10.7860/JCDR/2016/16306.7408 PMID:27134863 Bhavanani, A. B., Ramanathan, M., Balaji, R., & Pushpa, D. (2014). Differential effects of uninostril and alternate nostril pranayamas on cardiovascular parameters and reaction time. International Journal of Yoga, 7(1), 60–65. doi:10.4103/0973-6131.123489 PMID:25035609

12

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

Bhavanani, A. B., Sanjay, Z., & Madanmohan. (2011). Immediate effect of sukha pranayama on cardiovascular variables in patients of hypertension. International Journal of Yoga Therapy, 21(1), 73–76. doi:10.17761/ijyt.21.1.y007g51341634172 PMID:22398346 Bhavanani, M. D. (2012). The history of yoga from ancient to modern times. Satya Press. Bidgoli, M. M., Taghadosi, M., Gilasi, H., & Farokhian, A. (2016). The effect of sukha pranayama on anxiety in patients undergoing coronary angiography: A single -blind randomized controlled trial. Journal of Cardiovascular and Thoracic Research, 8(4), 170–175. doi:10.15171/jcvtr.2016.34 PMID:28210473 Bloomfield, M. (1897). Hymns of the Atharva-veda: together with extracts from the ritual books and the commentaries (Vol. 1). Clarendon Press. Budhi, R. B., Payghan, S., & Deepeshwar, S. (2019). Changes in lung function measures following bhastrika pranayama (bellows breath) and running in healthy individuals. International Journal of Yoga, 12(3), 233–239. doi:10.4103/ijoy.IJOY_43_18 PMID:31543632 Chakrabarty, J., Vidyasagar, M. S., Fernandes, D., & Mayya, S. (2016). Emotional aspects and pranayama in breast cancer patients undergoing radiation therapy: A randomized controlled trial. Asia-Pacific Journal of Oncology Nursing, 3, 199–204. PMID:27981159 Cooper, S., Oborne, J., Newton, S., Harrison, V., Coon, J. T., Lewis, S., & Tattersfield, A. (2003). Effect of two breathing exercises (Buteyko and pranayama) in asthma: A randomised controlled trial. Thorax, 58(8), 674–679. doi:10.1136/thorax.58.8.674 PMID:12885982 Dhyansky, Y. Y. (1987). The Indus valley origin of a yoga practice. Artibus Asiae, 48(1/2), 89–108. doi:10.2307/3249853 Garg, R., Malhotra, V., Tripathi, Y., & Agarawal, R. (2016). Effect of left, right and alternate nostril breathing on verbal and spatial memory. Journal of Clinical and Diagnostic Research: JCDR, 10, CC01–CC03. doi:10.7860/JCDR/2016/12361.7197 PMID:27042450 Gupta, R. K., Agnihotri, S., Telles, S., & Balkrishna, A. (2019). Performance in a Corsi block-tapping task following high-frequency yoga breathing or breath awareness. International Journal of Yoga, 12(3), 247–251. doi:10.4103/ijoy.IJOY_55_18 PMID:31543634 Hakked, C. S., Balakrishnan, R., & Krishnamurthy, M. N. (2017). Yogic breathing practices improve lung functions of competitive young swimmers. Journal of Ayurveda and Integrative Medicine, 8(2), 99–104. doi:10.1016/j.jaim.2016.12.005 PMID:28601355 Jensen, A. R. (2006). Clocking the mind: mental chronometry and individual differences. Elsevier. Joshi, M., & Telles, S. (2009). A non randomized non-naïve comparative study of the effects of kapalabhati and breath awareness on event-related potentials in trained yoga practitioners. Journal of Alternative and Complementary Medicine (New York, N.Y.), 15(3), 281–285. doi:10.1089/acm.2008.0250 PMID:19243275 Kamath, A., Urval, R. P., & Shenoy, A. K. (2017). Effect of alternate nostril breathing exercise on experimentally induced anxiety in healthy volunteers using the simulated public speaking model: A randomized controlled pilot study. BioMed Research International, 2017, 162–166. doi:10.1155/2017/2450670 PMID:29159176

13

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

Kaminsky, D. A., Guntupalli, K. K., Lippmann, J., Burns, S. M., Brock, M. A., Skelly, J., DeSarno, M., Pecott-Grimm, H., Mohsin, A., LaRock-McMahon, C., Warren, P., Whitney, M. C., & Hanania, N. A. (2017). Effect of yoga breathing (pranayama) on exercise tolerance in patients with chronic obstructive pulmonary disease: A randomized, controlled trial. Journal of Alternative and Complementary Medicine (New York, N.Y.), 23(9), 696–704. doi:10.1089/acm.2017.0102 PMID:28714735 Kumar, N., Thanalakshmi, J., Kannan, R., Kumar, M., Allu, A. R., & Vijayalakshmi, B. (2018). The immediate effect of sheethali and sheethkari pranayama on blood pressure and cardiovascular changes among hypertensive patients. International Journal of Research in Pharmaceutical Sciences, 9, 1249–1252. Kuppusamy, M., Dilara, K., Ravishankar, P., & Julius, A. (2017). Effect of bhrāmarī prāṇāyāma practice on pulmonary function in healthy adolescents: A randomized control study. Ancient Science of Life, 36(4), 196–199. doi:10.4103/asl.ASL_220_16 PMID:29269971 Kuppusamy, M., Kamaldeen, D., Pitani, R., & Amaldas, J. (2016). Immediate effects of bhramari pranayama on resting cardiovascular parameters in healthy adolescents. Journal of Clinical and Diagnostic Research: JCDR, 10, CC17–CC19. doi:10.7860/JCDR/2016/19202.7894 PMID:27437210 Lieber, C., & Mohsenin, V. (1992). Cheyne-stokes respiration in congestive heart failure. The Yale Journal of Biology and Medicine, 65, 39–50. PMID:1509783 Madanmohan, U. K., Bhavanani, A. B., Vijayalakshmi, P., & Surendiran, A. (2005). Effect of slow and fast pranayams on reaction time and cardio- respiratory variables. Indian Journal of Physiology and Pharmacology, 49, 313–318. PMID:16440849 Mahendra, J., Mahendra, L., Ananthalakshmi, R., Parthiban, P. S., Cherukuri, S., & Junaid, M. (2017). Effect of pranayama on Ppar-γ, Nf-κBexpressions and red complex microorganisms in patients with chronic periodontitis – a clinical trial. Journal of Clinical and Diagnostic Research: JCDR, 11, ZC82– ZC86. PMID:28764300 Miles, W. R. (1964). Oxygen consumption during three yoga breathing patterns. Journal of Applied Physiology, 19(1), 75–82. doi:10.1152/jappl.1964.19.1.75 PMID:14104293 Miles, W. R., & Behanan, K. T. (1934). A metabolic study of three unusual learned breathing patterns practiced in the cult of yoga. The American Journal of Physiology, 109, 74–75. Miller, B. S. (1996). Yoga: Discipline of freedom: the yoga sutra attributed to patanjali. University of California Press. Mourya, M., Mahajan, A. S., Singh, N. P., & Jain, A. K. (2009). Effect of slow-and fast-breathing exercises on autonomic functions in patients with essential hypertension. Journal of Alternative and Complementary Medicine (New York, N.Y.), 15(7), 711–717. doi:10.1089/acm.2008.0609 PMID:19534616 Muktibodhananda, S. (1999). Swara yoga. Yoga Publications Trust. Muktibodhananda, S. (2002). Hatha Yoga pradipika. Yoga Publications Trust. Nagendra, H. R. (1988). Pranayama: The art and science. Vivekanand Kendra Yoga Prakashan.

14

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

Naik, G. S., Gaur, G. S., & Pal, G. K. (2018). Effect of modified slow breathing exercise on perceived stress and basal cardiovascular parameters. International Journal of Yoga, 11, 53–58. PMID:29343931 Naveen, K. V., Nagarathna, R., Nagendra, H. R., & Telles, S. (1997). Yoga breathing through a particular nostril increases spatial memory scores without lateralized effects. Psychological Reports, 81(2), 555–561. doi:10.2466/pr0.1997.81.2.555 PMID:9354111 Niranjanananda, S. (2004). Prana, pranayama, prana vidya. Yoga Publications Trust. Nivethitha, L., Manjunath, N. K., & Mooventhan, A. (2017). Heart rate variability changes during and after the practice of bhramari pranayama. International Journal of Yoga, 10(2), 99–102. doi:10.4103/09736131.205518 PMID:28546681 Novaes, M. M., Palhano-Fontes, F., Onias, H., Andrade, K. C., Lobão-Soares, B., Arruda-Sanchez, T., Kozasa, E. H., Santaella, D. F., & de Araujo, D. B. (2020). Effects of yoga respiratory practice (Bhastrika pranayama) on anxiety, affect, and brain functional connectivity and activity: A randomized controlled trial. Frontiers in Psychiatry, 11, 467. doi:10.3389/fpsyt.2020.00467 PMID:32528330 Polich, J. (1999). P300 in clinical applications. Electroencephalography: basic principles, clinical applications and related field. Baltimore: Urban and Schwarzenberg. Pradhan, B., Mohanty, S., & Hankey, A. (2018). Effect of yogic breathing on accommodate braille version of six-letter cancellation test in students with visual impairment. International Journal of Yoga, 11(2), 111–115. doi:10.4103/ijoy.IJOY_40_17 PMID:29755219 Pramanik, T., Pudasaini, B., & Prajapati, R. (2010). Immediate effect of a slow pace breathing exercise bhramari pranayama on blood pressure and heart rate. Nepal Medical College Journal, 12, 154–157. PMID:21446363 Pramanik, T., Sharma, H. O., Mishra, S., Mishra, A., Prajapati, R., & Singh, S. (2009). Immediate effect of slow pace bhastrika pranayama on blood pressure and heart rate. Journal of Alternative and Complementary Medicine (New York, N.Y.), 15(3), 293–295. doi:10.1089/acm.2008.0440 PMID:19249921 Prem, V., Sahoo, R. C., & Adhikari, P. (2013). Comparison of the effects of buteyko and pranayama breathing techniques on quality of life in patients with asthma–a randomized controlled trial. Clinical Rehabilitation, 27(2), 133–141. doi:10.1177/0269215512450521 PMID:22837543 Raghuraj, P., Nagarathna, R., Nagendra, H. R., & Telles, S. (1997). Pranayama increases grip strength without lateralized effects. Indian Journal of Physiology and Pharmacology, 41, 129–133. PMID:9142556 Raghuraj, P., Ramakrishnan, A. G., Nagendra, H. R., & Telles, S. (1998). Effect of two selected yogic breathing techniques of heart rate variability. Indian Journal of Physiology and Pharmacology, 42, 467–472. PMID:10874345 Raghuraj, P., & Telles, S. (2004). Right uninostril yoga breathing influences ipsilateral components of middle latency auditory evoked potentials. Neurological Sciences, 25(5), 274–280. doi:10.100710072004-0354-9 PMID:15624085

15

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

Raghuraj, P., & Telles, S. (2008). Immediate effect of specific nostril manipulating yoga breathing practices on autonomic and respiratory variables. Applied Psychophysiology and Biofeedback, 33(2), 65–75. doi:10.100710484-008-9055-0 PMID:18347974 Rajesh, S. K., Ilavarasu, J. V., & Srinivasan, T. M. (2014). Effect of bhramari pranayama on response inhibition: Evidence from the stop signal task. International Journal of Yoga, 7(2), 138–141. doi:10.4103/0973-6131.133896 PMID:25035624 Ramdev, S. (2005). Pranayama: its philosophy and practice. Divya Prakashan. Santaella, D. F., Devesa, C. R., Rojo, M. R., Amato, M. B., Drager, L. F., Casali, K. R., Montano, N., & Lorenzi-Filho, G. (2011). Yoga respiratory training improves respiratory function and cardiac sympathovagal balance in elderly subjects: A randomised controlled trial. BMJ Open, 1(1), e000085. doi:10.1136/bmjopen-2011-000085 PMID:22021757 Saoji, A. A., Raghavendra, B. R., & Manjunath, N. K. (2018). Immediate effects of yoga breathing with intermittent breath retention on the autonomic and cardiovascular variables amongst healthy volunteers. Indian Journal of Physiology and Pharmacology, 62, 41–50. Saoji, A. A., Raghavendra, B. R., Rajesh, S. K., & Manjunath, N. K. (2018). Immediate effects of yoga breathing with intermittent breath holding on response inhibition among healthy volunteers. International Journal of Yoga, 11(2), 99–104. doi:10.4103/ijoy.IJOY_65_16 PMID:29755217 Sarkar, A., Mukherjee, A. D., Bera, M. K., Das, B., Juyal, N., Morthekai, P., Deshpande, R. D., Shinde, V. S., & Rao, L. S. (2016). Oxygen isotope in archaeological bioapatites from India: Implications to climate change and decline of bronze ageharappan civilization. Scientific Reports, 6(1), 1–9. doi:10.1038rep26555 Shahab, L., Sarkar, B. K., & West, R. (2013). The acute effects of yogic breathing exercises on craving and withdrawal symptoms in abstaining smokers. Psychopharmacology, 225(4), 875–882. doi:10.100700213012-2876-9 PMID:22993051 Shannahoff-Khalsa, D. S., Boyle, M. R., & Buebel, M. E. (1991). The effects of unilateral forced nostril breathing on cognition. The International Journal of Neuroscience, 57(3-4), 239–249. doi:10.3109/00207459109150697 PMID:1938166 Sharma, B., Misra, R., Singh, K., & Archana, R. S. (2013). Comparative study of effect of anulomaviloma (pranayam) and yogic asanas in premenstrual syndrome. Indian Journal of Physiology and Pharmacology, 57, 384–389. PMID:24968577 Shetty, P., Reddy, K. K., Lakshmeesha, D. R., Shetty, S. P., Kumar, S., & Bradley, R. (2017). Effects of sheetali and sheetkari pranayamas on blood Pressure and autonomic function in hypertensive patients. Integrative Medicine, 16, 38–43. PMID:30936803 Singh, K., Bhargav, H., & Srinivasan, T. M. (2016). Effect of uninostril yoga breathing on brain hemodynamics: A functional near-infrared spectroscopy study. International Journal of Yoga, 9(1), 12–19. doi:10.4103/0973-6131.171711 PMID:26865766 Singh, V. (1987). Personal experiences: Effect of respiratory exercises on asthma the pink city lung exerciser. The Journal of Asthma, 24(6), 355–359. doi:10.3109/02770908709070967 PMID:3443597

16

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

Singh, V., Wisniewski, A., Britton, J., & Tattersfield, A. (1990). Effect of yoga breathing exercises (pranayama)on airway reactivity in subject with asthma. Lancet, 335(8702), 1381–1383. doi:10.1016/01406736(90)91254-8 PMID:1971670 Stancák, A., Jr, & Kuna, M., Srinivasan, Vishnudevananda, S., & Dostálek, C. (1991b). Kapalabhati yogic cleansing exercise. I. Cardiovascular and respiratory changes. Homeostasis in Health and Disease, 33, 126–134. PMID:1818666 Stancák, A. Jr, Kuna, M., Novák, P., Srinivasan, M. A., Dostálek, C., & Vishnudevananda, S. (1991a). Observations on respiratory and cardiovascular rhythmicities during yogic high-frequency respiration. Physiological Research, 40, 345–354. PMID:1751482 Subramanian, R. K., Devaki, P. R., & Saikumar, P. (2016). Alternate nostril breathing at different rates and its influence on heart rate variability in non practitioners of yoga. Journal of Clinical and Diagnostic Research: JCDR, 10, CM01–CM02. doi:10.7860/JCDR/2016/15287.7094 PMID:26894062 Telles, S., & Desiraju, T. (1991). Oxygen consumption during pranayamic type of very slow-rate breathing. The Indian Journal of Medical Research, 94, 357–363. PMID:1794892 Telles, S., Gupta, R. K., Gandharva, K., Vishwakarma, B., Kala, N., & Balkrishna, A. (2019). Immediate effect of a yoga breathing practice on attention and anxiety in pre-teen children. Children (Basel, Switzerland), 6(7), 84. doi:10.3390/children6070084 PMID:31336661 Telles, S., Gupta, R. K., Singh, N., & Balkrishna, A. (2016). A functional near-infrared spectroscopy study of high-frequency yoga breathing compared to breath awareness. Medical Science Monitor Basic Research, 22, 58–66. doi:10.12659/MSMBR.899516 PMID:27351626 Telles, S., Maharana, K., Balrana, B., & Balkrishna, A. (2011). Effects of high frequency yoga breathing called kapalabhati compared with breath awareness on the degree of optical illusion perceived. Perceptual and Motor Skills, 112(3), 981–990. doi:10.2466/02.20.22.PMS.112.3.981-990 PMID:21853784 Telles, S., Nagarathna, R., & Nagendra, H. R. (1994). Breathing through a particular nostril can alter metabolism and autonomic activities. Indian Journal of Physiology and Pharmacology, 38, 133–137. PMID:8063359 Telles, S., Nagarathna, R., & Nagendra, H. R. (1996). Physiological measures of right nostril breathing. Journal of Alternative and Complementary Medicine (New York, N.Y.), 2(4), 479–484. doi:10.1089/ acm.1996.2.479 PMID:9395677 Telles, S., Pal, S., Gupta, R. K., & Balkrishna, A. (2018). Changes in reaction time after yoga bellowstype breathing in healthy female volunteers. International Journal of Yoga, 11(3), 224–230. doi:10.4103/ ijoy.IJOY_70_17 PMID:30233116 Telles, S., Raghuraj, P., Dhananjay, A., & Naveen, K. V. (2008). Immediate effect of high frequency yoga breathing on attention. Indian Journal of Medical Sciences, 62(1), 20–22. doi:10.4103/0019-5359.38919 PMID:18239268

17

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

Telles, S., Raghuraj, P., Maharana, S., & Nagendra, H. R. (2007). Immediate effect of three yoga breathing techniques on performance on a letter-cancellation task. Perceptual and Motor Skills, 104(3_suppl), 1289–1296. doi:10.2466/pms.104.4.1289-1296 PMID:17879663 Telles, S., Sharma, S. K., & Balkrishna, A. (2014). Blood pressure and heart rate variability during yoga-based alternate nostril breathing practice and breath awareness. Medical Science Monitor Basic Research, 20, 184–193. doi:10.12659/MSMBR.892063 PMID:25408140 Telles, S., & Singh, N. (2018). Research-Based Perspectives on the Psychophysiology of Yoga. IGI Global. doi:10.4018/978-1-5225-2788-6 Telles, S., Singh, N., & Balkrishna, A. (2011). Heart rate variability changes during high frequency yoga breathing and breath awareness. BioPsychoSocial Medicine, 5(1), 1–7. doi:10.1186/1751-07595-4 PMID:21486495 Telles, S., Singh, N., & Balkrishna, A. (2012). Finger dexterity and visual discrimination following two yoga breathing practices. International Journal of Yoga, 5(1), 37–41. doi:10.4103/0973-6131.91710 PMID:22346064 Telles, S., Singh, N., & Puthige, R. (2013). Changes in P300 following alternate nostril yoga breathing and breath awareness. BioPsychoSocial Medicine, 7(1), 11. doi:10.1186/1751-0759-7-11 PMID:23721252 Telles, S., Verma, S., Sharma, S. K., Gupta, R. K., & Balkrishna, A. (2017). Alternate-nostril yoga breathing reduced blood pressure while increasing performance in a vigilance test. Medical Science Monitor Basic Research, 23, 392–398. doi:10.12659/MSMBR.906502 PMID:29284770 Telles, S., Vishwakarma, B., Gupta, R. K., & Balkrishna, A. (2019). Changes in shape and size discrimination and state anxiety after alternate-nostril yoga breathing and breath awareness in one session each. Medical Science Monitor Basic Research, 25, 121–127. doi:10.12659/MSMBR.914956 PMID:31006767 Telles, S., Yadav, A., Gupta, R. K., & Balkrishna, A. (2013). Reaction time following yoga bellows type breathing and breath awareness. Perceptual and Motor Skills, 117(1), 1131–1140. doi:10.2466/22.25. PMS.117x10z4 PMID:24422341 Telles, S., Yadav, A., Kumar, N., Sharma, S., Visweshwaraiah, N. K., & Balkrishna, A. (2013). Blood pressure and purdue pegboard scores in individuals with hypertension after alternate nostril breathing, breath awareness, and no intervention. Medical Science Monitor, 19, 61–66. doi:10.12659/MSM.883743 PMID:23334063 Thanalakshmi, J., Ravindran, R., Sembulingam, K., & Sembulingam, P. (2014). Impact of Sheetali and Sheetkari Pranayama on the topographic mapping of the brain waves. IOSR Journal of Pharmacy, 4(10), 51–57. doi:10.9790/3013-04010051057 Veerabhadrappa, S. G., Baljoshi, V. S., Khanapure, S., Herur, A., Patil, S., Ankad, R. B., & Chinagudi, S. (2011). Effect of yogic bellows on cardiovascular autonomic reactivity. Journal of Cardiovascular Disease Research, 2(4), 223–227. doi:10.4103/0975-3583.89806 PMID:22135480

18

 Psychophysiological Effects and the Applications of Yoga Breathing Practices

Vialatte, F. B., Bakardjian, H., Prasad, R., & Cichocki, A. (2008). EEG paroxysmal gamma waves during bhramari pranayama: A yoga breathing technique. Consciousness and Cognition, 18(4), 977–988. doi:10.1016/j.concog.2008.01.004 PMID:18299208 Werntz, D. A., Bickford, R. G., & Shannahoff-Khalsa, D. (1987). Selective hemisphereic stimulation by unilateral forced nostril breathing. Human Neurobiology, 6, 165–171. PMID:3449485 Wooten, S. V., Cherup, N., Mazzei, N., Patel, S., Mooney, K., Rafiq, A., & Signorile, J. F. (2018). Yoga breathing techniques have no impact on isokinetic and isoinertial power. Journal of Strength and Conditioning Research, 34(2), 430–439. doi:10.1519/JSC.0000000000002771 PMID:30142133 Zelano, C., Jiang, H., Zhou, G., Arora, N., Schuele, S., Rosenow, J., & Gottfried, J. A. (2016). Nasal respiration entrains human limbic oscillations and modulates cognitive function. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 36(49), 12448–12467. doi:10.1523/ JNEUROSCI.2586-16.2016 PMID:27927961

19

Psychophysiological Effects and the Applications of Yoga Breathing Practices

APPENDIX 1 Table 1. Studies reporting the psychophysiological effects of pranayamas Sl. No. 1.

Citation Bertisch et al., 2017

Name of pranayama

Sample, study design and variables

Results

Yogic slow breathing or bhastrika pranayama

Sample: 26 long-term yoga practitioners and 26 age- and sex-matched controls Design: cross-sectional Variables: cardiac vagal outflow as assessed by respiratory sinus arrhythmia (RSA).

1. Yoga practitioners exhibited augmented RSA compared to controls during unpaced breathing. 2. RSA did not differ between groups during paced breathing at 0.25 Hz. 1. Significant increase in the standard deviation of all NN intervals, low-frequency (LF) component, LF/HF (low frequency/high frequency) ratio and significant decrease in the HF component during anuloma-viloma pranayama. 2. A significant increase in total power in systolic BPV, diastolic BPV and mean BPV.

2.

Bhagat et al., 2017

Anuloma-viloma pranayama

Sample: 12 adults Design: before and after Variables: heart rate variability (HRV), blood pressure variability (BPV) and baroreflex sensitivity (BRS)

3.

Bhavanani et al., 2003

Mukhabhastrika

Sample: 22 healthy school boys Design: before and after Variables: reaction time

Decrease in visual reaction time and auditory reaction time after mukhabhastrika.

4.

Bhavanani et al., 2014

Suryabhedana, chandrabhedana, nadishuddhi

Sample: 20 yoga trained volunteers Design: self as control Variables: reaction time, heart rate and blood pressure

Heart rate and blood pressure reduced following chandrabhedana, and nadisuddhi, and increased after suryabhedana.

5.

Budhi et al., 2019

Bhastrika pranayama

Sample: 30 males Design: randomized controlled trial Variables: pulmonary function test

1. Significant increase in all variables, i.e.,forced vital capacity, maximum voluntary ventilation and peak expiratory flow rate after bhastrika. 2. Significant increase in peak expiratory flow rate and maximum voluntary ventilation after physical activity.

6.

Garg et al., 2016

Left, right and alternate nostril breathing

Sample: 51 females Design: randomized comparative trial Variables: memory testing

Increase in the recall of digit span-forward, digit-span backward, associate learning and spatial memory scores with left, right and alternate nostril breathing.

7.

Gupta et al., 2019

Kapalabhati

Sample: 15 adults Design: self as control Variables: working memory and spatial memory scores using the Corsi block-tapping task

Backward total scores and backward Corsi span improved after breath awareness.

8.

Hakked et al., 2017

Sample: 27 swimmers Bhastrika pranayama, anulom- Design: randomized controlled trial vilom pranayama Variables: pulmonary function test and state anxiety

9.

Joshi &Telles, 2009

Kapalabhati

Sample: 30 male participants Design: before and after Variables: P300 event related potential

1. The P300 peak latency decreased after kapalabhati. 2. The P300 peak amplitude increased after breath awareness.

10.

Kamath et al., 2017

Anuloma-viloma pranayama

Sample: 30 medical students Design: randomized controlled trial Variables: visual analog mood scale and selfstatements during public speaking scale

A significant effect of phase was present, but group and gender did not have statistically significant influence on the mean anxiety scores. The test group showed a trend towards lower mean scores for the anxiety factor.

11.

Kuppusamy et al., 2017

Bhramari pranayama

Sample: 90 healthy adolescents Design: randomized controlled trial Variables: pulmonary function test

Improvement in all pulmonary function parameters including forced vital capacity (FVC), forced expiratory volume in first second (FEV1), FEV1/FVC ratio, maximum voluntary ventilation and peak expiratory flow rate in the bhramari pranayama group.

12.

Kuppusamy et al., 2016

Bhramari pranayama

Sample: 90 healthy adolescents Design: randomized controlled trial Variables: heart rate and blood pressure

There was a decrease in heart rate, pulse pressure, mean arterial pressure, rate pressure product and double product following bhramari pranayama. 1. Increase in respiratory pressures and respiratoryendurance following Savitri pranayama. 2. Heart rate, rate pressure product and double product decreased in savitri pranayama group and increased in bhastrikagroup.

Improvement in the pranayama group as compared to control group in maximal voluntary ventilation, forced vital capacityand number of strokes per breath.

13.

Madanmohan et al., 2005

Savitripranayam and Bhastrika pranayama

Sample: 30 students Design: pre-post Variables: visual reaction time and auditory reaction time, maximum expiratory pressure and maximum inspiratory pressure, blood pressure, heart rate, double product and ratepressure product.

14.

Miles & Behanan, 1934

Ujjayi, kapalabhati and bhastrika pranayama

Sample: single subject Design: before, during and after Variables: metabolic rate

Increase in oxygen consumption during pranayama compared to sitting and as well as compared to reclined.

15.

Miles, 1964

Ujjayi, kapalabhati and bhastrika pranayama

Sample: single subject Design: before, during and after Variables: metabolic rate

A 12-35 percent increase in the oxygen consumption.

Naik et al., 2018

Slow breathing or bhastrika pranayama

Sample: 100 male adults Design: randomized controlled trial Variables: perceived stress scale, body mass index, waist-hip ratio, heart rate and blood pressure

Decrease in heart rate, systolic blood pressure, diastolic blood pressure and perceived stress in the experimental group.

16.

continued on following page

20

Psychophysiological Effects and the Applications of Yoga Breathing Practices

Table 1. Continued Sl. No.

Citation

Name of pranayama

Sample, study design and variables

Results

17.

Naveen et al., 1997

Right nostril breathing, left nostril breathing, alternate nostril breathing, breath awareness

18.

Nivethitha et al., 2017

Bhramari pranayama

Sample: 16 adults Design: before, during and after Variables: heart rate variability

Increase in HR and low frequency spectrum of HRV and a significant reduction in high frequency spectrum of HRV during the practice of bhramari which revert to normal after the practice.

19.

Novaes et al., 2020

Bhastrika pranayama

Sample: 30 adults Design: randomized controlled trial Variables: fMRI, state-trait anxiety, positive and negative affect

     1. Modulated the activity of brain regions involved in emotional processing, particularly the amygdala, anterior cingulate, anterior insula, and prefrontal cortex.      2. States of anxiety and negative affect decreased.

20.

Pramanik et al., 2009

Bhastrika Pranayama

Sample: 39 adults Design: before and after Variables: blood pressure and heart rate

Systolic and diastolic blood pressure decreased with a slight fall in heart rate after bhastrika pranayama.

21.

Pramanik et al., 2010

Bhramari pranayama

Sample: 50 adults Design: before and after Measure: heart rate and blood pressure

Diastolic blood pressure and mean blood pressure decreased after bhramari pranayama.

22.

Raghuraj & Telles, 2004

Right nostril yoga breathing (RNB)

Sample:14 healthy male volunteers Design: self as control design Variables: middle latency auditory evoked potentials (MLAEPs)

The peak amplitudes of Na wave and Nb wave increased on the right side during right nostril breathing.

23.

Raghuraj et al., 1997

Sample: 130 school children Right, left and alternate- nostril Design: before and after breathing Variables: grip strength (both hands)

24.

Raghuraj et al., 1998

Kapalabhati and Nadisuddhi

Sample: 12 male volunteers Design: self as control Variables: high frequency and low frequency component of heart rate variability

1. Increase in low frequency (LF) power and LF/HF ratio while high frequency (HF) power was significantly decreased following kapalabhati. 2. There were no significant changes following nadisuddhi.

25.

Rajesh et al., 2014

Bhramari pranayama

Sample: 31 male students Design: self as control Variables: stop signal task

Significant decrease in stop signal reaction time after bhramari pranayama.

Sample:108 school children Design: randomized controlled trial Variables: verbal and spatial memory tests

All four groups showed a significant increase in spatial test scores at after practice.

Significant increase in grip strength of both hands following the right, left and alternate-nostril breathing.

26.

Santaella et al., 2011

Bhastrika

Sample: 29 healthy elderly subjects Design: randomized controlled trial Decrease in the low frequency component and low frequency/high Variables: pulmonary function, maximum frequency ratio as well as increases in maximum expiratory and expiratory and inspiratory pressures, heart rate inspiratory pressures. variability and blood pressure variability

27.

Saoji et al., 2018

Kumbhak

Sample: 39 adults Design: self as control Variables: autonomic and cardiovascular variables

Decrease in the heart rate, stroke volume, mean arterial pressure and cardiac output along with increase in SDNN, RMSSD, pNN50, LF power and baroreflex sensitivity.

28.

Saoji et al., 2018

Kumbhak

Sample: 36 adults (19 females) Design: self as control Stop-signal reaction time reduced in both breath holding and breath Variables:stop-signal reaction time, mean awareness groups. reaction time to go stimuli, and the probability of responding on-stop signal trials

29.

Singh et al., 2016

Right and left nostril breathing

Sample: 32 male adults Design: self as control Variables: brain hemodynamic changes

1. Oxy Hb levels increased significantly in the left prefrontal cortex during right nostril breathing. 2. Left nostril breathing showed a trend towards significance for reduction in oxy-Hb in the right hemisphere. 3. Breath awareness reduced deoxy-Hb in the left hemisphere. 1. Heart rate, systolic blood pressure and diastolic blood pressure increased during kapalabhati. 2. All frequency bands of R-R interval variability were reduced in kapalabhati. A 0.1 Hz rhythm was present in the record of R-R intervals as well as in the blood pressure during kapalabhati.

30.

Stancák et al.,1991b

Kapalabhati

Sample: 17 yoga practitioners Design: before, during and after Variables: electrocardiogram, systolic blood pressure and diastolic blood pressure, R-R intervals and respiration

31.

Stancák et al., 1991a

Kapalabhati

Sample:24 Design: before, during and after Variables: respiratory movements, blood pressure and R-R intervals of ECG

32.

Subramanian et al., 2016

Anuloma-viloma pranayama

Sample: 25 adults Low frequency power increased, high frequency decreased and their Design: before and after ratio increased after 6 and 12 minutes. Variables: heart rate variability and respiration

33.

Telles & Desiraju, 1991

Ujjayi pranayama with breath holding

Sample:10 healthy male subjects Design: before and after Variables: oxygen consumption

1. Short kumbhakpranayaa breathing increased the oxygen consumption compared to the pre-pranayama base-line period. 2. Long kumbhak pranayama breathing decreased the oxygen consumption.

34.

Telles et al., 2019

Kapalabhati

Sample: 61 pre-teen children Design: self as control Variables: attention and anxiety

1. Increase in total attempts and net scores after high frequency yoga breathing. 2. Increased in wrong attempts after breath awareness. 3. Decrease in state anxiety after all three interventions.

35.

Telles et al., 2016

Kapalabhati pranayama

Sample: 40 male adults Design: self as control design Variables: hemodynamic changes using a functional near-infrared spectroscopy

1. Significant reduction in oxy-haemoglobin during and after kapalabhati pranayama on the left and right sides. 2. Significant reduction in deoxy- haemoglobin during and after the quiet sitting on left and right sides.

continued on following page 21

Psychophysiological Effects and the Applications of Yoga Breathing Practices

Table 1. Continued Sl. No.

Citation

Name of pranayama

Sample, study design and variables

Results

Telles et al., 2011

Kapalabhati

Sample: 30 male participants Design: randomized controlled trial Variables: optical illusion

37.

Telles et al., 1994

Right nostril breathing, left nostril breathing or alternate nostril breathing

Sample: 48 male subjects Design: randomized controlled trial Variables: oxygen consumption

1. Right nostril breathing, left nostril breathing and alternate nostril breathing showed a significant increase, in baseline oxygen consumption. 2. The left nostril pranayama group showed an increase in galvanic skin resistance.

38.

Telles et al., 1996

Surya anuloma-viloma pranayama (right nostril breathing)

Sample: 12 volunteers Design: before and after Variables: oxygen consumption and blood pressure

1. Increase in oxygen consumption after right nostril breathing. 2. Increase in systolic blood pressure and decrease in digit pulse volume after right nostril breathing.

39.

Telles et al., 2018

Bhastrika pranayama

Sample: 25 female adults Design: self as control Variables: reaction time

Reduction in the time taken to obtain a correct response breath awareness and quiet sitting.

40.

Telles et al., 2008

Kapalabhati

Sample: 110 (46 medical students, 48 middle aged adults, 16 older adults) Design: before and after Variables: six letter cancellation task

Scores of six letter cancellation task improved after kapalabhati.

41.

Telles et al., 2007

Sample: 20 male volunteers Right, left, and alternate nostril Design: self as control design breathing Variables: letter-cancellation task

42.

Telles et al., 2014

Alternate nostril yoga breathing

Sample: 26 male volunteers Design: randomized controlled trial Variables: heart rate variability, respiration rate and blood pressure

Systolic BP and respiration rate decreased while RMSSD and NN50 components of heart rate variability increased during anuloma-viloma pranayama.

43.

Telles et al., 2011

Kapalabhati

Sample: 38 male volunteers Design: self as control design Variables: heart rate variability

Decrease in NN50, pNN50 and the mean RR interval during and after kapalabhati and after breath awareness was found.

44.

Telles et al., 2012

High frequency yoga breathing (kapalabhati) and breath awareness

Sample: 140 subjects Design: randomized controlled trial Variables: finger dexterity and visual discrimination

Significant improvement in the finger dexterity task and shape and size discrimination task after both interventions, though the magnitude of change was higher in kapalabhati.

45.

Telles et al., 2013

Alternate nostril yoga breathing

Sample:20 males Design: self as control Variables: P300 event related potential

Following alternate nostril yoga breathing, there was a significant increase in the P300 peak amplitudes at Fz, Cz, and Pz and a significant decrease in the peak latency at Fz alone. 1. Decrease in systolic blood pressure, mean arterial blood pressure and the time taken to complete the digit vigilance test after anulomaviloma pranayama. 2. Decrease in the time taken to complete the digit vigilance test after sitting quietly.

36.

There was a significant decrease in the degree of optical illusion after both kapalabhati and breath awareness.

Fewer errors in the letter cancellation task following right and alternate nostril yoga breathing were found.

46.

Telles et al., 2017

Anuloma-viloma pranayama

Sample: 15 male adults Design: self as control Variables: digit vigilance test and blood pressure

47.

Telles et al., 2019

Anuloma-viloma pranayama

Sample: 50 male adults Design: self as control Variables: shape and size discrimination task and state anxiety

1. Reduction in the errors scores in the shape and size discrimination task after anuloma-viloma pranayama. 2. Reduction was found in the level of state anxiety after breath awareness and quiet sitting sessions.

48.

Telles et al., 2013

Bhastrika pranayama

Sample: 70 healthy males Design: self as control design Variables: reaction time

Reaction time improved after bhastrika pranayama.

49.

Veerabhadrappa et al., 2011

Mukhabhastrika

Sample: 50 healthy male subjects Design: before and after Measure: cardiovascular autonomic reactivity

Decrease in basal heart rate, increase in valsalva ratio and deep breathing difference in heart rate; and reduction in fall of systolic blood pressure on posture variation was seen.

50.

Vialatte et al., 2008

Bhramari pranayama

Sample: 8 subjects Design: before, during and after Variables: electroencephalogram

High frequency paroxysmal gamma waves were observed during bhramari pranayama.

51.

Werntz et al., 1987

Forced uninostril breathing

Sample: 5 yoga naïve persons Design: before and after Variables: electroencephalogram

Forced uninostril breathing was associated with an increase in the EEG amplitude in the contralateral cerebral hemisphere.

52.

Wooten et al., 2018

Ujjayi, bhastrika, and Kapalabhati

Sample: 32 adults Design: before and after Variables: lower-limb power

No significant differences in power output before and during the specific breathing practices.

Abbreviations: BPV = Blood Pressure Variability; BRS = Baroreflex sensitivity; DBP = Diastolic Blood Pressure; ECG = Electrocardiogram; FEV1 = Forced Expiratory Volume in first second; FVC = Forced Vital Capacity; Hb = Haemoglobin; HF = High Frequency; HR = Heart Rate; HRV = Heart Rate Variability; LF = Low Frequency; MLAEPs = Middle Latency Auditory Evoked Potentials; NN50 = number of successive NN intervals that differ by more than 50 milliseconds; pNN50 = proportion of NN50 divided by the total number of NN intervals; RMSSD = Root Mean Square of the Successive Differences; R-R interval = time between two successive heart beats, also called NN intervals; RSA = Respiratory sinus arrhythmia; RT = Reaction Time; SBP = Systolic Blood Pressure; SDNN = Standard deviation of the NN intervals; SSRT = Stop-Signal Reaction Time.

22

Psychophysiological Effects and the Applications of Yoga Breathing Practices

Table 2. Studies reporting the therapeutic effects of pranayamas Sl. No.

Citation

Name of pranayama

Results SNOT-22 score improved in the pranayama group by the end of 4 weeks and continued until the 12th week of assessment.

1.

Abishek et al., 2019

2

Sample: 23 patients with hypertension Bhavanani et al., 2011 Sukha pranayama Design: before and after Variables:heart rate and blood pressure

3

Bidgoli et al., 2016

Sample: 80 patients undergoing coronary angiography Sukha pranayama Design: randomized controlled trial Variables: anxiety

Decrease in anxiety score after sukha pranayama.

4

Chakrabarty et al., 2016

Bhramari, sheetali and nadishodhana

Sample: 160 patients with breast cancer Design: randomized controlled trial Variables: impatience, worry, anxiety and frustration

Pranayama group showed lesser mean scores for impatience, worry, anxiety and frustration.

5

Cooper et al., 2003

Breathing through the Pink city lung exerciser

Sample: 69 asthmatic patients Design: randomized controlled trial Variables: pulmonary functions

Symptoms remained relatively stable in the PCLE and placebo groups but were reduced in the Buteyko group.

6

Kaminsky et al., 2017

Dirgha pranayama

Sample: forty-three patients with COPD. Design: randomized controlled trial Variables: 6-min walk distance (6MWD), lung function, markers of oxidative stress and systemic inflammation, and Variables of dyspnoea and quality of life.

The 6MWD increased in the pranayama group. Small improvements in inspiratory capacity and air trapping were also seen in pranayama group.

7

Kumar et al., 2018

Sheetali and sitkari

Sample: 20 hypertensive patients and 20 healthy individuals Design: before and after Variables: blood pressure and cardiovascular parameters

Reduction in blood pressure, heart rate, pulse pressure, mean arterial pressure and double product after sheetali and sitkari.

Mahendra et al., 2017

Deep breathing according to Swami Shivananda

Sample: 60 patients with chronic periodontitis Design: randomized controlled trial Variables: probing pocket depth (PPD), clinical attachment level (CAL), bleeding index and plaque index were recorded and the presence of PPAR-γ, NF-κB and RCM were assessed

1. CAL improved in the pranayama group. 2. Reduction in the expression of NF-κB and increase in PPAR-γ expression levels in pranayama group. 3. Reduction of subgingival microorganisms in pranayama group.

9

Mourya et al., 2009

Yogic fast breathing and slow breathing

Sample: 60 patients with essential hypertension Design: randomized controlled trial Variables: blood pressure, standing-to-lying ratio (S=L ratio), immediate heart rate response to standing (30:15 ratio), Valsalva ratio, heart rate variation with respiration (E=I ratio), hand-grip test, and cold presser response

1. BP decreased longitudinally with both interventions over a 3-month period. 2. S=L ratio, 30:15 ratio, E=I ratio, and BP response in the hand grip and cold pressor test showed significant change only in slow-breathing exercise.

10

Pradhan et al., 2018

Bhramari

Sample: 19 visually impaired adults Design: self as control design Variables: six letter cancellation test

1. Net scores improved after bhramari pranayama and breath awareness. 2. Wrong scores increased following breath awareness.

11

Prem et al., 2013

Diaphragmatic breathing, thoracic breathing, upper lobe breathing and full yogic breathing progressing to right nostril breathing, left nostril breathing and alternate nostril breathing

Sample: 120 adults with asthma Design: randomized controlled trial Variables: asthma quality of life questionnaire, asthma control questionnaire and pulmonary function test.

1. Buteyko breathing group showed better trends of improvement in total asthma quality of life questionnaire score than the pranayama and control groups. 2. Pranayama group showed significant improvement in total asthma quality of life questionnaire score than the control group.

12

Shahab et al., 2013

Deep breathing and alternate nostril breathing

Sample: 96 smokers Design: randomized controlled trial Variables: cigarette cravings, strength of urges, desire to smoke

All craving measures were reduced immediately after yoga breathing.

13

Sharma et al., 2013

Anuloma-viloma pranayama

Sample: 60 females with premenstrual syndrome Design: randomized controlled trial Variables: blood pressure, heart rate, electromyogram, galvanic skin response, respiratory rate and peripheral temperature

Heart rate, systolic blood pressure, diastolic blood pressure, electromyogram, galvanic skin response and respiratory rate showed a significant reduction after the third menstrual cycle in anuloma-viloma pranayama and asana group,

14

Shetty et al., 2017

Sheetali and sitkari

Sample: 60 hypertensive adults Design: randomized controlled trial Variables:blood pressure and autonomic and respiratory functions

Decrease in systolic blood pressure, respiratory rate, heart rate and heart rate variability parameters.

15

Singh, 1987

Breathing through the Pink city lung exerciser, inspiration: expiration at a ratio of 1:2

Sample: 12 patients with asthma Design: placebo controlled (non-randomized) trial Variables: pulmonary functions

Increase in peak expiratory flow rate and decrease in frequency of nocturnal wheeze.

16

Singh et al., 1990

Breathing through the Pink city lung exerciser

Sample: 22 patients with asthma Design: placebo controlled (non-randomized) trial Variables: pulmonary functions, histamine challenge test

Airway reactivity to histamine improved more with use of the PCLE than with the placebo device.

Telles et al., 2013

Anuloma-viloma pranayama

Sample: 90 patients with hypertension Design: self as control Variables: blood pressure and Purdue pegboard task score

1. Decrease in systolic and diastolic blood pressure and an improvement in Purdue pegboard task scores for both hands and for the right hand after anuloma-viloma pranayama. 2. Breath awareness showed a decrease in systolic blood pressure. 3. The right hand scores improved after reading a magazine.

8

17

Bhramari

Sample, design and variables Sample: 60 patients with chronic sinusitis Design: randomized controlled trial Variables: sino-nasal outcome test (SNOT-22 score).

A significant decrease in heart rate, systolic pressure, pulse pressure, mean arterial pressure, rate-pressure product, and double product was seen.

Abbreviations: 30:15 Ratio = Immediate heart rate response to standing; 6MWD = 6-min Walk Distance; BI = Bleeding Index; BP = Blood Pressure; CAL = clinical Attachment Level; COPD = Chronic Obstructive Pulmonary Disease; E=I ratio = Heart rate variation with respiration; HR = Heart Rate; MFEV1 = Mean forced expiratory volume in first second; NF-κB = Nuclear Factor-Kappa B; PCLE = Pink City Lung Exerciser; PI = Plaque Index; PPAR-γ = Peroxisome Proliferator-Activated Receptor Gamma; PPD = Probing Pocket Depth; RCM = Red Complex Microorganisms; S=L ratio = standing-to-lying ratio; SNOT = Sino-Nasal Outcome Test.

23

24

Chapter 2

Psychophysiology:

Healing Effects of Voluntarily Regulated Breathing Practices Heather Mason The Minded Institute, UK Patricia L. Gerbarg New York Medical College, USA Richard P. Brown College of Physicians and Surgeons, Columbia University, USA

ABSTRACT This chapter describes the physiological mechanisms that underpin the varying effects of different types of breath practices inherent in the yoga tradition and ultimately the role that breathing techniques play in person and public health. Concurrently, the script elucidates how different practices alter psychophysiological states clarifying why and how they may be employed with specific health populations, how they may enhance and or maintain well-being, and clear guidance regarding precautions and contraindications.

INTRODUCTION For thousands of years, breathing techniques, called pranayama in Sanskrit and translated as “restraining of life force or breath”, have been practiced for their energetic, physiological, psychological, and spiritual benefits. Here we use the current secular terminology, Voluntarily Regulated Breathing Practices (VRBPs), to indicate techniques in which a person volitionally changes their respiratory pattern (Telles & Singh, 2013). The term VRBP recognizes the multi-cultural origins of breath practices and it is more readily accepted by scientists, governmental agencies, and people of diverse religious backgrounds. The physiological effects of VRBPs overlap but differ from, those of the automatic breathing they override. DOI: 10.4018/978-1-7998-3254-6.ch002

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Psychophysiology

VRBPs entail volitional alterations in breath patterns that include changes in the respiratory rate, the relative length of four components of each breath cycle (inhalation, pause, exhalation, pause), forcefulness of the inhale or exhale, depth of respiration, use of the diaphragm versus accessory muscles, nostril dominance, resistance to airflow, and breath holding (Brown & Gerbarg, 2012; Gerbarg & Brown, 2016). Effects may be synergistic when techniques are done in concert with visualization, movement, vocalization, and meditation, or through the sequencing of practices. A growing body of research is documenting the physiological and psychological effects of breath techniques, inspiring yoga teachers and healthcare professionals to apply VRBPs therapeutically. Health educators and policy makers are recognizing the value of VRBPs as public health interventions for an array of mental health disorders and physical health problems, for example, cardiovascular, gastrointestinal, inflammation, pain, and stress-related medical conditions. Notably, breath practices are valued for their therapeutic role in emotion regulation, autonomic balance, stress reduction, attentional control, mood, cognitive function, social engagement, and substance abuse. Accordingly, VRBPs offer significant benefits in prevention and treatment of mental disorders, including the global epidemic of anxiety, depression, trauma, and post-traumatic stress disorders. VRBPs are advantageous for both individual and public health interventions. They are easy to learn in a short period of time, teachable to large groups by small numbers of providers, inexpensive, rapidly effective, low risk, non-stigmatizing, and easily adapted for acceptance by diverse cultures when taught in a secular manner. Furthermore, VRBPs can be transmitted and sustained by local trainees without professional degrees. Many techniques can be practiced anywhere and would be undetectable, even in the company of others. This chapter explores the mechanisms that underlie physiological changes associated with specific breathing practices and, where possible, links these with clinical applications. To understand the rapid, global effects of Voluntarily Regulated Breathing Practices (VRBPs) on the mind and body, we must recognize that every breath activates a constellation of psychophysiological processes. We review the anatomy and physiology of natural, automatic breathing, including cardiorespiratory reflexes. From this foundation, we explore the evidence supporting neurophysiological theories about the mechanisms of action brought into play by changing the patterns of breathing and the effects these may have on psychophysiological states and pathophysiology. Taking into account the afferent, bottom-up messages rising from the respiratory system to the brain, the efferent, top-down messaging descending from the brain to the body, the feedback loops at all levels—peripheral, brainstem, subcortical, and cortical—we may well stand in awe of the complex systems we are trying to delineate.

Physiology of Natural Respiration – Automatic Breathing Automatic breathing is controlled by inspiratory neurons, the dorsal respiratory group (DRG), and expiratory neurons, the ventral respiratory group (VRG) located in the brainstem (medulla oblongata). Additionally, the pontine respiratory group (PRG) houses the pneumotaxic and apneustic centers, which jointly inhibit inspiration and prevent over distension of the lungs. The DRG lies in the nucleus tractus solitarius (NTS), a relay station for chemosensory and somatosensory information. Inhalation begins when the DRG signals the phrenic nerve to contract the diaphragm downwards reducing pressure within the pulmonary system and expanding the lungs. Air is pulled into the respiratory tract when the pressure within the lungs is lower than the external air pressure. When the lungs expand numerous stretch receptors within the alveoli (millions of small air sacs throughout the lungs), lung tissue, and airways discharge; 25

 Psychophysiology

hence transmitting information about the rate and volume of inflation through vagal nerve fibers. This data is fed through the NTS to the DRG, inhibiting further inspiration and allowing passive exhalation. Any change in the pattern of breathing may affect afferent bottom-up messaging and neurological processing. Conversely, thoughts and emotions generated in subcortical and neocortical structures can alter breath patterns. For example, when experiencing fear, subcortical structures such as the amygdala may send top-down descending signals that increase respiratory rate and mobilization for fight or flight, a state of increased oxygen demand. These changes upregulate sympathetic drive, which increases heart rate for more expedient oxygen delivery to meet the increased oxygen demand. Concurrently, signals generated by the rapid breathing and accelerated heart rate may feed back to the brain, amplifying the experience of fear.

The Cardiorespiratory Reflex and the Vagal Brake The main neural pathways between the respiratory system and the brain are the vagus nerves, the tenth cranial nerves. Millions of sensory receptors, triggered by breathing, send signals through vagal nerve afferents to the nucleus tractus solitarius (NTS), and from there to brain networks involved in regulating neurophysiological states. Descending efferent vagal pathways originate in two brain stem nuclei. The first, the nucleus ambiguus (NA) sends action potentials down the vagus to synapses in the sinoatrial node of the heart, releasing acetylcholine, its primary neurotransmitter, which reduces the rate and intensity of cardiac contractions. Vagal input to the heart’s sino-atrial node can override the intrinsic pacemaker, inducing changes in the beat to beat intervals (the components of the heart rate). The cardio inhibitory vagal pathway is a reflex that either increases or decreases the effects of the vagus nerve on the heart. Breathing regulates the vagal brake that slows the heartbeat. During exhalation vagal influence increases, slowing heart rate; during inhalation vagal influence decreases, allowing heart rate to speed up. This influence of breathing on vagal output to the heart produces respiratory sinus arrhythmia (RSA), slight variations in heart rate between inhalation and exhalation. Through this reflex mechanism it is possible to rapidly reduce heart rate and calm the body and mind simply by elongating the exhalation in relation to the inhalation. The second nucleus, the dorsal motor nucleus (DMN) sends descending messaging from the brain stem via the vagus to the gastrointestinal tract. Patterns of breathing generate signals transmitted through vagal pathways which can affect the entire gastrointestinal tract, including motility, release of digestive enzymes, insulin release, and anti-inflammatory processes.

Chemoreceptor Reflexes Two types of chemoreceptors play major roles in respiratory adaptations. The first are the peripheral arterial chemoreceptors. Located mainly in the carotid bodies and the aortic bodies, they respond primarily to changes in the partial pressure of oxygen in the bloodstream. The second type, the central chemoreceptors, located in the brainstem, respond to changes in blood pH mediated mainly by the partial pressure of carbon dioxide in the bloodstream.

26

 Psychophysiology

Oxygen Homeostasis Given that our moment-to-moment survival depends on breathing, it is understandable that messages about respiratory function receive high priority in the brain and greatly influence psychophysiological states. Neurons need oxygen to synthesize adenosine triphosphate (ATP), an essential energy transport molecule within cells. The continuous flow of oxygen across alveolar walls into the blood stream is vital for neurological functioning. Elevated carbon dioxide (CO2), hypercapnia, can increase blood acidity. Decreased levels of CO2, hypocapnia, lead to vasoconstriction (narrowing of blood vessels) reducing O2 delivery. Changes in blood gas levels have immediate effects on the respiratory, nervous, and cardiovascular systems. This information is transmitted to the medulla oblongata via the glossopharyngeal and vagus nerves, which initiate reflex mechanism necessary for recalibration to homeostatic levels.

Oxygen Saturation and Hypoxemia In a healthy person, normal arterial blood oxygen saturation (SAO2) is between 95-100 percent. Anything below 90 percent is considered hypoxemia (abnormally low blood oxygen) or hypoxia (a state of oxygen deprivation). Oxygen homeostasis is crucial. Hypoxia can trigger physiological changes to increase the availability of O2 to cells, for example, increased red blood cell production, elevated heart rate, and accelerated respiratory rate. Chronic hypoxia is linked to cardiovascular disease, cancer, and neurodegenerative conditions by hypoxia inducible factor-1 (HIF-1), a transcription factor meditating changes in gene expression essential to meet the needs of reduced O2 (Michiels, 2004). In the short-term HIF-1 supports adaptive strategies for managing diminishing SAO2, but when persistently elevated, HIF-1 triggers pathways that generate oxidative stress and mitochondrial damage (Thomas & Ashcroft, 2019). Chronically elevated HIF-1 communicates with another transcription factor, nuclear factor of the kappa-light-chain-enhancer of activated B cells (NF-KB), notorious for upregulating genes associated with disease and inflammation (Acker, 2004). Additionally, when SAO2 is lower, vagal output to the heart is withdrawn, resulting in a higher resting heart rate (Siebenmann et al., 2019). Although temporarily useful to hasten oxygen delivery, sustained elevations of resting heart rate and autonomic dysregulation correlate with reduced physiological and psychological resiliency. Consequently, even marginally reduced SAO2 can be pernicious. Practicing VRBPs that increase SAO2 may help prevent or reverse these pathogenic changes.

Hypercapnia – High Blood Levels of Carbon Dioxide When CO2 combines with water in the bloodstream, it forms carbonic acid, which lowers the blood pH (making it more acidic). Hypercapnia, an excess blood level of carbon dioxide (CO2), activates the respiratory chemoreflex to accelerate exhalations, expel excess CO2, and restore a healthy acid-base balance (pH) in the blood. Mediated by the sympathetic nervous system excess carbon dioxide can also trigger the stress response. Temporary hypercapnia may occur during intense exercise. As a byproduct of ATP synthesis, CO2 levels can increase up to tenfold when the body’s energy demands are exceedingly high (Ogoh et al., 2008). In lung diseases such as COPD, wherein air trapping and CO2 retention are common, the chemoreflex can become chronically activated, leading to heightened sympathetic drive with anxiety. In people with chronic anxiety, whose sympathetic activity is elevated, the rate of cellular respiration is accelerated, 27

 Psychophysiology

producing hypercapnia. When both mechanisms co-active each other, an upward spiral of anxiety and rapid, shallow accessory breathing may ensue and become habitual. Bernardi’s research with mountain climbers and healthy populations demonstrated that reducing the breath rate, thereby increasing oxygen saturation (SAO2) could reduce chemoreflex sensitivity, increase tolerance of elevated CO2 levels and enhance psychological and physiological resiliency (Bernardi et al., 2001; Spicuzza et al., 2000). Slowing the breath rate can also increase baroreceptor sensitivity (BRS) (Bernardi et al., 2001). The baroreflex is a primary endogenous means of regulating blood pressure, based on the sensitivity of pressure receptors to changes in blood pressure.

Hypocapnia – Low Blood Levels of Carbon Dioxide Hyperventilation occurs when breathing is faster, deeper, or more forceful than normal, such that too much CO2 is expelled causing hypocapnia, reduced levels of CO2 in the blood. This can occur in conditions of low oxygen, for example, at high altitudes, or by the practice of rapid forceful breathing techniques, such as kapalabhati, breath of fire, high frequency cyclical breathing (Kriya), Holotropic Breathing, excessive bhastrika, or prolonged Ha breath. Stress and anxiety can also cause hyperventilation escalating to a panic attack. In people who suffer from panic attacks, high frequency or forceful VRBPs can trigger a panic attack. CO2 serves as a vasodilator. When levels drop quickly, for example, during hyperventilation, small blood vessels throughout the body may constrict, sometimes causing cramps in the hands and feet. Cerebral blood flow and brain activity may also decline (Hayashi et al., 2008). Hypocapnia can be reversed by slow controlled breathing with pursed lips or by breathing into a paper bag, normalizing CO2 levels and efficient oxygen delivery.

How the Lungs Talk to the Brain Respiration, the only autonomic function that can be altered at will, provides a portal into the intrinsic systems of communication between the body and the brain (Gerbarg & Brown, 2016). Breathing is intimately linked with autonomic function. Voluntary reduction of breathing rate enhances parasympathetic activity (as indicated by increased respiratory sinus arrhythmia (RSA) and the high-frequency component of heart rate variability, HRV) and reduces sympathetic activity. Slow breathing decreases chemoreflex sensitivity and baroreflex sensitivity (Spicuzza et al., 2000). VRBPs change the interoceptive (perception of the internal state of the body), chemosensory, barosensory, and other messages being sent from the respiratory system to the central nervous system, messages that command top priority, probably because any impediment to the flow of breath constitutes an immediate life threat. When the lungs talk, the brain listens. Changes in respiratory pattern also affect signals from the diaphragm, thoracic musculature, and the heart, whose beat-to-beat interval is linked to the breath rate. The profound effects of breathing on how we think, feel, and behave can be understood as the sum of many psycho-neurophysiological mechanisms. Convergent evidence supports the impact of VRBPs on critical neuroanatomical pathways, the autonomic nervous system, the exchange of oxygen and carbon dioxide, and brain electrical rhythms.

28

 Psychophysiology

Innervation of the Respiratory System To understand the global effects of breathing and VRBPs, we begin by tracing the pathways through which moment-to-moment respiratory activity is communicated to brain central processing in the interoceptive and anterior cingulate cortices. The lungs are innervated by vagal and spinal nerves with both sensory and motor components (Wang et al., 2018). Afferent respiratory signals originate in the activation of sensory receptors in lung tissues, airways, pleura, and musculature of the thorax and diaphragm. The Central Autonomic Network (CAN) includes the communicating and regulating feedback loops between the heart, the lungs, and the central nervous system (brainstem, subcortex, and neocortex). The brainstem structures include the nucleus acumbens (NA), ventrolateral medulla, nucleus tratus solitarius (NTS), and parabrachial nucleus (PBN). The subcortical structures are the periaqueductal gray, hypothalamus, amygdala, hippocampus, and thalamus. Neocortical structures include the medial prefrontal cortex, anterior cingulate cortex, and insular cortex (Bates et al., 2019). Most respiratory receptors arise from fibers in the vagal nerves. Subtypes of vagal sensory receptors include slowly and rapidly adapting (stretch) receptors and C-fiber receptors. Three sensory receptor end organs have been identified: smooth muscle associated airway receptors (SMARS), neuroepithelial bodies (NEBs) and visceral pleura receptors (VPRs). Large numbers of neuroepithelial bodies (NEBs) are diffusely spread throughout the epithelium (lining) of intrapulmonary airways. NEBS are neuroendocrine cells, innervated by myelinated vagal afferent fibers (Adriaensen et al., 2006). NEBs store serotonin and peptide transmitters. They include polymodal sensors responsive to diverse stimuli, including hypoxia (low oxygen), hypercapnia (increased CO2), and mechanical stretch (Cutz et al., 2013). In addition to afferent vagal (visceral) homeostatic sensory information from the lungs, somatic sensory feedback from respiratory muscles and humoral homeostatic information about changes in concentrations of carbon dioxide ascend to the interoceptive and anterior cingulate cortices for processing and integration.

How The Brain Listens to the Lungs Bidirectional communication between the brain and the respiratory system engages multiple pathways. However, our focus is on the vagal pathways through the nucleus tractus solitarius (NTS), the parabrachial nucleus (PBN). From these centers, ascending homeostatic sensory pathways deliver information to the hypothalamus (e.g. regulation of the hypothalamic-pituitary adrenal access and cortisol release), periaqueductal gray, amygdala (e.g., emotion processing), ventral medial nucleus (VMb) of the thalamus, and the insular cortex. The insular (interoceptive) cortex, through connectivity with higher cortical areas such as the anterior cingulate cortex and the Salience Network, is thought to provide primary processing, re-representation, and interpretation of interoceptive information (Strigo & Craig, 2016). Any change in the pattern of breathing will change the pattern of sensory messages being transmitted from the respiratory apparatus to the brain. As we learn more about the downstream effects of specific breath patterns, we can develop more therapeutic breathing regimens to correct specific psychological and physical disturbances. This is a foundational concept for our neurophysiological models of the effects of VRBPs.

29

 Psychophysiology

Polyvagal Theory – Autonomic Function, Stress Response and Social Engagement The Polyvagal Theory (Porges, 2001) identifies three evolutionary developments in the human autonomic nervous system: the ancient ‘reptilian’ unmyelinated vagal system, the sympathetic nervous system, and the most recent, the myelinated vagal system. Myelin is a lipid sheath that covers nerve axons, enabling more rapid transmission of impulses. The unmyelinated vagal pathway originates in the brainstem dorsal motor nucleus (DMN) of the vagus and regulates sub diaphragmatic (abdominal) organs. The myelinated vagal pathways, found primarily in mammals and mainly in primates, originates in the brain stem nucleus ambiguus (NA) and regulates supra diaphragmatic organs (heart and lungs). Recent research attributes the myelinated vagus with a major role in human emotional and social development (Porges, 2003; Porges & Furman, 2010; Porges & Carter 2017). When we feel safe, the myelinated vagal system is working well, and the parasympathetic and sympathetic nervous systems are more balanced. In this psychophysiological state we are less likely to react defensively. Our social engagement systems are more responsive, supporting positive feelings, compassion, empathy, cooperation, connectedness, and pro-social behaviours. In contrast, when we feel threatened, the sympathetic nervous system may come into play with a greater tendency to react defensively with fight or flight, approach or avoidance behaviors. If mobilization by the sympathetic system is unable to resolve the threat (we can neither fight back nor escape), and if we feel as though our life is in danger, our system may default to the older unmyelinated vagal circuitry, responsible for death feigning in animals. In this state, we may become immobilized, frozen with fear (freeze reaction) or we may dissociate, disconnect, or feel numb. Feelings from the body, particularly interoceptions, contribute heavily to assessments of safety or danger by the insular and cingulate cortical networks. Polyvagal theory explains the relationship between interoception, autonomic nervous system function, emotions and behavior. Any stimuli, such as interoceptions, that increase parasympathetic tone, tend to recruit social engagement networks and reduce defensive behaviors. Social engagement networks orchestrate communication through facial expression, vocalization, and perceptions of communicatory signals, such as tone and prosidy of voice, facial expressions, and physical postures (Porges & Furman, 2010; Porges & Carter, 2017). Porges and Carter (2017) describe the face-heart connection and the emergence of the Social Engagement System (2017), “When the newer [myelinated] mammalian vagus is optimally functioning in social interactions (inhibiting the sympathetic excitation that promotes fight-or-flight behaviors), emotions are well regulated, vocal prosody is rich, and the autonomic state supports calm, spontaneous social engagement behaviors. Thus, it is highly advantageous to become adept at managing social interactions by using the newer mammalian vagus rather than the more limited options of recruiting the sympathetic nervous system to support fight-or-flight behaviors or the “older” [unmyelinated] vagus to support immobilization and death feigning.” “The face-heart system is bidirectional such that the newer myelinated vagal circuit influences social interactions, and positive social interactions influence vagal function to optimize health, dampen stress-related physiological states, and support positive reciprocal social interactions.” Respiratory rhythms and patterns generate rhythmic changes in parasympathetic/sympathetic activity, heart rate, brain electrical activity, and interoceptive messaging from the respiratory systems as described above. These messages are rapidly transmitted to brain regulatory networks where they can quickly

30

 Psychophysiology

convince the mind that conditions are safe or unsafe. Such mechanisms make the voluntary regulation of breathing a powerful means for influencing mental states and treating psychological disorders.

Interoception and Homeostatic Emotions From the first Interoception Summit in 2016, organized by the Laureate Institute for Brain Research, a consensus definition of interoception emerged: “Interoception refers to the process by which the nervous system senses, interprets, and integrates signals originating from within the body, providing a moment-by-moment mapping of the body’s internal landscape across conscious and unconscious levels. Interoceptive signaling has been considered a component process of reflexes, urges, feelings, drives, adaptive responses, and cognitive and emotional experiences, highlighting its contributions to the maintenance of homeostatic functioning, body regulation, and survival” (Khalsa et al., 2018). Malfunctions of interoceptive systems and the autonomic nervous system contribute to most psychological and physical illnesses. In post-traumatic stress disorder (PTSD), several errors of interoceptive processing could contribute to symptomatology, for example: hyper vigilance entails both an attentional bias and erroneous predictions of threat; autonomic imbalance with sympathetic over activity and parasympathetic underactivity; misperceptions and over reactivity to bodily sensations; misinterpretations of the interoceptive ‘map’ of the body condition; and cognitive bias, such as catastrophizing. In effect, VRBPs provide easy access to the body’s interoceptive communication system and a potential therapeutic tool for correcting malfunctions and improving homeostatic regulation. Furthermore, we can design breath interventions as noninvasive physiological probes for research. Strigo and Craig (2016) expanded the concept of interoception to include, “a homeostatic sensory motor hierarchy”. This includes the central autonomic network, the emotional motor system, the limbic system, and the afferent vagal homeostatic sensory input from the body to the brain (particularly to the interoceptive cortex). Their studies investigate the role of the anterior insular and anterior cingulate cortices in the regulation of sympathovagal and emotional balance”. At the top of the architectural sensory motor hierarchy, the right forebrain (cerebral cortex, thalamic nuclei, and hypothalamus) predominantly processes sympathetic activity, negative affect (emotion), avoidance behavior, and energy consumption; the left forebrain mainly processes parasympathetic activity, positive affect, approach behavior, and energy replenishment. Emotional input from the limbic system, somatosensory information, and interoceptive input from the body are integrated in the insular cortex and cingulate cortex, which serve as the homeostatic-emotionallimbic-sensorimotor neocortex, overseeing adaptive homeostatic control of the mind and body. Evidence suggests that feelings and awareness are generated in the insular cortex; motivations and agency in the cingulate cortex (Strigo & Craig, 2016). A functional magnetic resonance imaging (fMRI) study in healthy humans compared the effects of slow-paced breathing at 5 bpm with fast-paced breathing at 20 bpm on the activation and responsivity of the insular cortex and the cingulate cortex while participants viewed emotional images (Strigo & Craig, 2016). Analysis of the fMRI data comparing blood-oxygen level dependent (BOLD) signals observed increased insular and cingulate activation in the left hemisphere while viewing images with positive emotional valence during slow breathing, and increased cingulate activation in the right hemisphere while viewing images with negative emotional valence during fast breathing. Slow breathing (5 bpm) significantly correlated with reduced heart rate, increased heart rate variability (HRV) (i.e., parasympathetic 31

 Psychophysiology

tone) and increased activation of the left insula, the interoceptive brain area activated by emotionally positive images. This evidence supports the importance of forebrain areas for the interoceptive respiratory modulation of emotional affect and sympathovagal balance, in accord with the homeostatic model of adaptive (emotional) behavior. Studies are providing increasing support for the proposed integration of interoceptive and emotional activity, and the regulation of sympathovagal balance in the insular and cingulate cortices. These observations further our understanding of breathing techniques that can be used to modulate how we experience our emotions and, potentially, to treat a wide range of psychiatric conditions (Brown & Gerbarg, 2012; Gerbarg, 2008; Gerbarg & Brown, 2011, 2015).

Breathing Effects on GABA-ergic Transmission The calming effects of cardiorespiratory synchronization through increases vagal tone could also exert calming effects through gamma-aminobutyric acid (GABA), the primary inhibitory neurotransmitter in the brain. Noble and Hochman (2019) noted that slow adaptive receptors (SARs) also synapse with GABAergic interneurons that postsynaptically inhibit nucleus tractus solitarius (NTS) neurons that connect with the amygdala and hypothalamus. They also hypothesized that cardiorespiratory vagal afferents could inhibit excitatory input to the NTS. Streeter and colleagues (2012, 2018, 2019) have been testing the Vagal-GABA Theory of Inhibition, which proposed that stress induces: imbalance of the autonomic nervous system (ANS) with decreased parasympathetic nervous system (PNS) activity and increased sympathetic nervous system (SNS) activity; underactivity of the inhibitory neurotransmitter, gamma amino-butyric acid (GABA); and increased allostatic load. They hypothesized that yoga and coherent breathing practices would correct underactivity of the PNS and GABA system, in part through stimulation of the vagal nerves, and thereby reduce allostatic load. In patients with major depressive disorder (MDD) they found that a 12-week program of Iyengar Yoga and coherent breathing significantly reduced depression and increased levels of GABA in the thalamus on mass resonance spectrometry. These preliminary results support their hypothesis that increased GABAergic transmission, for example, to the central extended nucleus of the amygdala could also help quiet the amygdalar over activity and symptoms seen in PTSD and anxiety disorders.

The Breath Entrains Brain Electrical Activity Until recently, neurophysiological models of the effects of respiration focused on automatic brain stem circuitry. The scientific study of higher-level voluntary breathing mechanisms and cognitive aspects of breathing is now coming into focus. Conscious control and awareness of breathing are ripe for exploration. Many brain regulatory centers exhibit respiration-locked or entrained oscillations. Entrainment or respiration-locking means that the phase and frequency of oscillations seen in recordings of the electrical activity in an area of the brain become organized in a regular, repeating pattern at a particular frequency, determined by the rate or frequency of respirations. As larger areas become entrained, neurons fire in synchrony, generating larger amplitude wages. It is thought that when brain waves become synchronized in larger clusters of neurons and in groups of neurons in separate but interconnected brain structures, the brain functions more effectively and efficiently, for example solving more complex problems or understanding more complex situations quickly. Notably, strong respiration locking has been found in the insula, amygdala, premotor, olfactory, caudal-medial frontal (executive function) and temporal cortices 32

 Psychophysiology

(Herrero et al., 2018). The locking of brain oscillations to the respiratory cycle is increased by volitional pacing of the breath and by attentional focus (awareness or mindfulness) of the breath. For example, in a small study of 15 young adults, 18 minutes of breath awareness improved primary working memory, spatial memory, and attention, but high frequency yoga breathing did not (Gupta et al., 2019). Using direct intracranial electroencephalogram (iEEG) recordings in patients with refractory epilepsy, Herrero and colleagues (2018) examined changes in iEEG-breath coherence (a measure of the degree of locking or entrainment of electrical rhythms by the breath rate) while subjects performed tasks aimed at voluntarily controlling the rate of respiration or increasing awareness of the respiratory cycle. They found that volitional paced breathing (VRBP) increased iEEG-breath coherence (locking) in the frontotemporal-insular network. When attention (awareness) was focused on breathing, iEEG-breath coherence (locking) increased in anterior cingulate, premotor, insular, and hippocampal cortices. Their findings indicate that breathing-related oscillations have a role in driving neuronal activity. This constitutes an important neuronal mechanism for the effects of volitional breath pacing and interoceptive attention (mindful awareness) on the electrical rhythms in areas of the brain known to be involved in executive functions, (e.g., behavioral planning and follow-through) and in the interoceptive monitoring of body states (Farb et al., 2013). Herrero and colleagues (2018) proposed respiratory rhythms as an organizing principle of cortical oscillations in the human brain. They demonstrated a network of areas involved in volitional (caudalmedial frontal, premotor, orbitofrontal, and motor cortex, insula, superior temporal gyrus, and amygdala) and attentive (anterior cingulum, premotor, insula, hippocampus) breathing, providing insights into potential brain mechanisms involved in therapeutic breathing exercises.

Voluntary Breath Pacing - Different Breath Patterns Have Different Effects In general, rapid or forceful breath practices are activating and stimulate the sympathetic nervous system. In contrast, slow, gentle breath practices, such as coherent breathing at 3 to 6 breaths per minute (bpm), are calming, reduce sympathetic over activity, and increase parasympathetic tone. Slow breath practices have few adverse effects, but rapid or forceful breath practices may cause symptoms of hyocapnia (lower levels of carbon dioxide). Moderately fast practices, such as bhastrika or “Ha” breath at 15 to 20 bpm for short periods of time (1 to 3 minutes with a 15-30 second rest between each set of 20 breath repetitions), initiate stimulation followed by calmness. Techniques that modify or enhance these effects, for example, breathing against airway resistance (e.g., ujjayi, ocean breath) can further increase parasympathetic tone. Resistance breathing entails creating resistance to the flow of air, for example using pursed lips, slight contraction of laryngeal muscles, partial closure of the glottis, chanting (tensing of vocal cords), partial obstruction by the tongue, breathing through clenched teeth, and other methods (Brown & Gerbarg, 2005, 2012, 2017).

Slow Voluntarily Regulated Breathing Practices Russo and others (2017) reviewed the effects of slow breathing on cardiorespiratory and autonomic nervous systems, focusing on diaphragm activity, ventilation efficiency, hemodynamics, heart rate variability, cardiorespiratory coupling, respiratory sinus arrhythmia and sympathovagal balance. Elliot (2010) coined the term coherent breathing to describe a method of breathing that would increase the range of movement of the diaphragm from 10 percent to 40-60 percent of the potential diaphragmatic 33

 Psychophysiology

range by breathing slowly, deeply, and rhythmically, at a frequency of about 5 bpm and consciously relaxing during exhalation. The entrainment of the heart rate to respiratory phase during slow breathing, via the cardiorespiratory reflex, increases vagal power (Jalife et al., 1983). Vagal feedback from pulmonary stretch receptors was shown to mediate respiratory sinus arrhythmia (RSA) in awake human subjects (Taha et al., 1985). During controlled, slow, deep breathing, the respiratory phase modulation of sympathetic activity is also stronger. A long-term shift towards more sympathetic/parasympathetic balance may require regular, possibly daily, practice of slow breathing over time. Reducing the breath rate can increase oxygen saturation (SAO2) and reduce chemoreflex sensitivity and response to changes in serum CO2. Bernardi discovered that slow breathing at around 6 bpm increased tolerance of elevated CO2 levels, reducing chemoreflex hyperactivity and enhancing psychological resiliency (Spicuzza et al., 2000). Slowing the breath rate can also increase baroreceptor sensitivity (BRS) (Bernardi et al., 2001). The baroreflex is a primary endogenous means of regulating blood pressure, based on the sensitivity of pressure receptors to changes in blood pressure. Slow, gentle Coherent Breathing or resonant breathing can be used to lower blood pressure.

Adverse Effects of Slow Voluntarily Regulated Breathing Practices Slow VRBPs are generally quite safe with few side effects. Although most people with asthma benefit from the long-term practice of slow VRBPs, those with unstable or severe asthma may experience some difficulty when initiating slow practices such as coherent breathing. The first effect of increased parasympathetic action on smooth airway muscles can cause a slight contraction of fine airways, making it more difficult to breathe. This can be prevented by teaching unstable asthmatics to use a technique called Moving the Breath, whereby one imagines moving the breath in circuits inside the body, particularly moving the breath up to the top of the head during the inhale and down to the base of the spine during the exhale. This method of keeping the airways open for fragile asthmatic patients was developed by author Richard Brown (Brown & Gerbarg, 2012).

The Holy Grail - Resonant Breathing and Coherent Breathing Resonant breathing and coherent breathing, the two most studied slow breath practices, are closely related. In studies of heart rate variability (HRV) biofeedback, Lehrer and colleagues (2000, 2003) identified a range of paced breath rates (4.5 to 7.0 bpm) that optimized HRV and cardiovascular resonance (CVR). HRV, a measure derived mathematically from the effects of respiration on heart rate-slight increases in heart rate during inhalation and slight decreases in heart rate during exhalation - is an indicator of sympathovagal balance, physiological flexibility, health, and longevity. The cardiovascular system (CVS) reveals resonance properties at frequencies of about 0.1 and 0.03 Hz due to the arterial baroreflex (Vaschillo et al., 2002; van de Vooren et al., 2007; Hammer & Saul, 2005). The 0.1 Hz resonance property of the CVS has demonstrated clinical utility in the treatment of disorders that involve autonomic nervous system dysregulation because patients may be trained to voluntarily produce high-amplitude oscillations in cardiovascular functions. Paced breathing at a rate of 0.1 Hz (6 bpm) has been used effectively in HRV biofeedback (Lehrer et al., 2000) to improve symptoms of physical and mental disorders. 34

 Psychophysiology

Resonant breathing or resonance-paced breathing refers to the breath rate that optimizes HRV and cardiovascular resonance (CVR) (Steffan et al., 2017). Heart rate and breathing synchronize or become resonant around 6 bpm (0.1 Hz) on average. Every person has a specific measurable resonance breath rate that optimizes HRV and CVR, typically between 4.5 and 7.0 bpm. For example, the resonant breath rate of physically healthy individuals who are over six feet tall (1.83 metres) is between 3.0 and 4.5 breaths per minute. In studies of HRV biofeedback in adults, the most common resonant breathing rate is 5.5 bpm (Steffan et al., 2017). Children have higher resonant breath rates, between 5 and 10 breaths per minute, depending on age, height, and health. It is neither necessary, nor always desirable, for an individual to breathe at their precise resonant breath rate. Breathing within a close range is usually effective. For example, if a person’s resonant rate is 5.5 bpm, positive effects will occur (with subtle differences) when their breath rate is between 4.0 and 6.0 bpm. Thus, it is possible when working with adult groups, to achieve good results simply by pacing everyone at an average of 5.0 bpm. Brown and Gerbarg refined the technique and definition of coherent breathing based on clinical studies and their work with disaster survivors, veterans, and patients with post-traumatic stress disorder, anxiety and mood disorders, and stress-related medical conditions (Descilo et al., 2010; Gerbarg et al., 2011; Carter et al., 2013; Gerbarg et al., 2015; Streeter et al., 2017; Nyer et al., 2018; www.breath-body-mind. com). Their techniques are designed to reduce sympathetic over drive and optimize parasympathetic activation, emotional calming, symptom resolution, attentional focus, cognitive function, pain relief, and social relationships (Brown & Gerbarg, 2012; Gerbarg & Brown, 2019). Brown and Gerbarg (2012, 2017) define coherent breathing as “gentle breathing in and out, preferably through the nose, with equal duration of inspiration and expiration at 3.0 to 6.0 bpm for most adults (depending on height and physical fitness), a rate that rapidly improves sympathovagal balance. Ultra slow paced breathing at 1 to 2 breaths per minute is used to engender deep meditative states with reduced metabolic activity. Research is sparse, since it requires many years of training to acquire the ability to comfortably sustain such slow breath rates. One report on a Japanese Yogi breathing at 1.9 bpm noted an increase in end-tidal CO2, but normal oxygenation was maintained by an increased tidal volume. Occipital parietal alpha activity increased, consistent with a calm, internally focused state (Peper et al., 2005). What is the optimal breath rate and how can therapeutic effects be enhanced? The answer depends on the characteristics of the breather (height, weight, physical fitness, medical conditions that affect cardiorespiratory capacity, and psychiatric disorders that affect emotional aspects of breathing), the target symptoms of the treatment, and/or the parameters of interest to a research team. It also depends on the individual’s activity and environmental factors. For example, during physical exertion, such as jogging, most people need to increase their respiratory rate to maintain an adequate supply of oxygen. The environment in which VRBPs are used can significantly influence the effects. The therapeutic benefits of practicing in an environment that feels safe and supportive cannot be overstated. The feeling of safety and being cared for can arise from the relationship with a healthcare provider, yoga teacher, or group. Adverse childhood events (ACEs), neglect, abuse, loss, or trauma disrupt the development of emotion regulatory systems, such that defensive reactions (fear or anger) and maladaptive behaviours (avoidance or aggression) emerge. Neuroplastic change in the process of growth and recovery requires the disruption or dismantling of old, habitual patterns of negative thoughts and emotions, followed by learning and retaining new perceptions and experiences, through the creation of new neural connections. The repeating, consistent experience of the caring presence of others enables the individual to feel safe 35

 Psychophysiology

enough to let down old defenses and allow their mind to accept and eventually internalize new positive experiences, perceptions, emotions, attitudes, thoughts, and behaviors. Slow VRBPs, like Coherent Breathing, enable the body to make the mind feel safe, thereby facilitating psychotherapy, sympathovagal balance, and healing (Gerbarg & Brown, 2019).

Slow Deep Diaphragmatic Breathing Versus Shallow Accessory Breathing Diaphragmatic breathing emphasizes volitional control of diaphragmatic excursions. Increasing the diaphragmatic movement between inhalation and exhalation by actively expanding and contracting the diaphragm increases the changes in lung volume (tidal volume). However, slowing the respiratory rate also increases tidal volume. As more time is allowed for inspiration during slow paced breathing, more air is drawn into the lungs naturally. In effect, slow breathing increases the tidal volume without focusing attention or conscious effort on moving the diaphragm. Slow breathing at about 6 bpm, compared to 15 or 3 bpm, decreased alveolar dead space, increased tidal volume, and improved oxygen saturation and exercise performance (Bernardi et al., 1998). Slow deep diaphragmatic breathing improves ventilation-perfusion matching, especially at a rate of 5-6 bpm (the coherent or resonant breathing rate). Ventilation refers to the period when the lungs are full of air. Perfusion, in this case, refers to the period when blood in the capillaries maximally arrives at the alveolar wall (membrane). When the timing of alveolar expansion and capillary perfusion are optimally matched, oxygen most effectively diffuses across the alveolar and capillary membranes to bind with blood hemoglobin (Russo et al., 2017). During inhalation the negative pressure caused by the descent of the diaphragm pulls venous blood into the chest cavity. At breath rates between 5-6 bpm, tidal volume peaks at the same time as maximal pulmonary perfusion (Bernardi et al., 1998). Bernardi’s research revealed that in environments where the partial pressure of oxygen is lower, such as at high altitudes, or in people with health conditions that compromise arterial oxygen saturation (e.g., chronic heart failure and chronic obstructive pulmonary disease, COPD), slow breathing can transcend these limitations and maintain SAO2 at healthier levels (Bernardi et al., 2002; Pomidori et al., 2009). Slow breathing at 5-6 bpm positively influences two additional cardiovascular measurements: baroreflex sensitivity (BRS) and heart rate variability (HRV). Baroreflex Sensitivity (BRS) is blunted in hypertension and other cardiovascular conditions, wherein elevated blood pressure is sensed as the norm and cardio-autonomic processes cease to regulate it. Slow breathing has the potential to restore BRS when a long inhalation provokes a momentary increase in blood pressure, intensifying action potentials sent from stretch receptors and baroreceptors in the carotid bodies and aortic arch to the NTS. When accompanied by a long exhalation, wherein momentary blood pressure drops, the increased fluctuations in moment-to-moment blood pressure between the inhale and exhale can restore or resuscitate the mechanisms of the baroreflex, improving BRS, as seen in hypertension, heart failure, and type-1 diabetes (Rosengård-Bärlund et al., 2011; Bernardi et al., 2002; Joseph et al., 2005). Lehrer hypothesized that at a respiratory rate between 5 and 6 bpm, RSA reaches its maximum amplitude and RSA frequency aligns with the frequency of spontaneous oscillations in blood pressure that are reflexively compensated for by changes in the baroreflex. When baroreflex and RSA frequencies are in concordance, they mutually stimulate each other, causing very high oscillations in heart rate, invoking maximum levels of heart rate variability (Lehrer et al., 2000; Lehrer et al., 2003). Accordingly, slow controlled breathing with an equal inhale and exhale support greater oxygen saturation and higher HRV. The value of this extends beyond improvements in cardiovascular functioning. 36

 Psychophysiology

HRV is an indicator of autonomic flexibility, derived from alternating basal sympathetic and parasympathetic output via vagal efferent to the sino-atrial node, the pacemaker of the heart, during an RR-interval (The Task Force of the European Society of Cardiology, 1996). High heart rate variability is associated with general psychological and physical wellness, cognitive flexibility, and longevity (Thayer et al., 2009). Low levels of HRV are found in a myriad of mental health conditions reflecting problems with shifting from states of mobilization to calmness, as correlated with emotion regulation. As a fundamental marker of psychophysiological health, low HRV has been linked to cardiovascular conditions, diabetes, social isolation, sedentariness, the transition from middle to older age, and earlier mortality (Porges, 2003; Kemp & Quintana, 2013; Jandackova et al., 2016). Maintaining or increasing HRV improves physical and psychological health. Coherent breathing optimizes HRV, is simple to learn and relatively inexpensive, and has proven clinical benefits with a high safety profile. Therefore, coherent breathing should be considered the foundational VRBP for public health and it warrants further clinical trials.

Rapid and Forceful Breath Practices High frequency breath practices (HFBPs), such as kapalabhati, bhastrika, breath of fire in Kundalini Yoga, cyclic breathing in Sudarshan Kriya Yoga, Holotropic Breathing, and mind-body methods, are widely used in yoga. Rapid breathing can be used to energize, activate, or focus attention. Zelano and colleagues (2016) found that when subjects were asked to breathe faster, respiration-locked oscillations increased in power and followed the breath frequency at the higher rate, particularly in premotor, caudal-medial frontal, orbitofrontal, and motor cortex, insula, superior temporal gyrus, and amygdala. As subjects returned to natural breathing, respiration-locked oscillations gradually decreased in power and continued to track at a lower frequency. Stronger respiration-locked oscillations were noted particularly in the amygdala during voluntary fast breathing as compared with natural breathing, suggesting that such practices could affect the processing of emotions. One study conducted on Sudarshan Kriya, a practice that includes rapid cyclical breathing, variable rate breathing and slow breathing, found elevated brain-derived-neurotrophic-factor (BDNF), colloquially dubbed the “fertilizer of neuroplasticity” (Zope & Zope, 2013). BDNF initiates neurogenesis, promotes N-methyl-D-aspartate receptor (NMDAR) plasticity (NMDARs are glutamate receptors, which play an important role in the rapid encoding of memory), and upregulates genes necessary for the rapid formation of dendritic connections. Thus, rapid breathing may facilitate neuroplasticity. This hypothesis is supported by the observation of changes in neurotransmission that are associated with hypocapnia. Hypocapnia increases transmission of glutamate, the primary excitatory neurotransmitter in the central nervous system (Godoy et al., 2017). Heightened levels of glutamate reduce the threshold for neural receptor firing, while enabling long-term-potentiation (LTP), a persistent strengthening of synapses based on recent patterns of activity with a long-lasting increase in signal transmission between two neurons. Long-term memory depends on LTP. Theoretically, rapid breathing techniques might facilitate personal growth by supporting the capacity to shift out of old maladaptive patterns and lay down new neurological connections, a testable hypothesis.

37

 Psychophysiology

Rapid and Forceful Breath Practices – Precautions and Contraindications Rapid and forceful breath practices accelerate heart rate, temporarily reduce oxygen delivery, and trigger physiological mechanisms associated with increased sympathetic drive, for example, increasing blood pressure. Such rapid VRBPs may not be safe for people with hypertension, cardiac conditions or diabetes. High frequency breathing is particularly contraindicated for individuals with Bipolar Disorder because it can trigger hypomanic, manic, or psychotic states. This decompensation may be immediate or delayed. It can occur hours or days after the individual has left their yoga sessions, when no one is available to notice or help get emergency care. It has been estimated that as many as 7 percent of adults in the US are on the bipolar spectrum and some may not know they have this condition until it is triggered by the overstimulation of rapid or forceful breathing. Individuals known to have bipolar disorder should be advised to avoid forceful or high frequency breathing practices (HFBPs). Also, rapid breathing accelerates the excretion of lithium by the kidneys. Consequently, patients being treated with lithium should avoid HFBPs, which could lower their serum lithium levels, exposing them to destabilization. Warning signs include agitation, over excitement, irritability, rapid or loud speech, disorganized behavior, grandiose statements, and hallucinations. Individuals with mild Bipolar II Disorder, whose symptoms are well controlled with medication, could engage in short practices of rapid breathing, with physician supervision and monitoring. Obviously, it is safer to engage in short periods of rapid breathing (e.g., 1-2 minutes) followed by periods of restful normal breathing (12-20 bpm), rather than prolonged periods of rapid breathing. Also, moderate rates, such as 20 to 40 bpm are safer than faster rates (40 to 80 bpm) or extremely fast rates (above 80 bpm). Prolonged rapid breathing can induce altered mental states which can destabilize vulnerable individuals, set off panic attacks in those with anxiety disorders, trigger flashbacks or dissociative episodes in people with post-traumatic stress disorder, or induce seizures in those with epilepsy (Gerbarg & Brown, 2011; Brown & Gerbarg, 2012).

Breathing Through the Nose Versus Through the Mouth In most yoga traditions, nasal breathing is favored over oral breathing. The nasal passages are lined with mucous and hairs, which prevent unwanted particles and pathogens from entering the respiratory tract. Additionally, nasal breathing moisturizes the air reducing uncomfortable dryness in the throat. Differences in physiological effects of nasal versus oral VRBPs are being studied with current technologies. Zelano and colleagues (2016) observed that during quiet natural inhalation (0.24-0.36 Hz or 14.4-21.6 bpm) slow oscillations are generated in various brain regions, enhancing theta waves (4-8 Hz) within and across the prefrontal cortex, olfactory cortex, amygdala, and hippocampus, creating collective entrainment of these region. Increased theta firing in this network engages signaling pathways connecting these subcortical structures, and enhances learning, particularly associative memory, i.e., associations between the primary experience, in this case breathing, and other contemporaneous experiences (Wang et al., 2018). In a few subjects, inspiration-locked increases also occurred in beta (13-30 Hz) and delta (0.5-2 Hz) frequencies. Consequently, nasal breathing could facilitate learning, memory encoding and possibly reshaping of emotional responses, which are influenced by affect-laden memories (Zhong et al., 2017). In Zelano’s model (2016), the slow oscillatory entrainment and coupling of signals across the olfactory cortex, prefrontal cortex, amygdala, and hippocampus, seem to diminish during oral breathing in comparison to nasal breathing. Accordingly, the development of new perceptions, positive emotional 38

 Psychophysiology

experiences, and healthier prosocial behaviours may be even more strongly supported through nasal breathing in contrast with oral breathing. The prefrontal cortex is tasked with modulating emotional reactions in the limbic system. Synchronization of oscillations may enhance communication and efficacy in the emotion regulatory circuit (prefrontal cortex, thalamus and amygdala), supporting more appropriate, nuanced emotional responses. Therefore, we suggest that attending to the environmental conditions, wherein nasal breathing (or any other VRBP) is practiced, may further improve clinical outcomes. Memories encoded during nasal breathing in a safe, supportive environment are more likely to be encoded with a positive emotional valence that may carry over into subsequent nasal breathing sessions, thereby augmenting feelings of safety that tend to arise as sympathovagal balance is restored. VRBPs performed in the company of compassionate professional or a supportive group impart a greater sense of safety and peacefulness that may be evoked, even when practicing alone. Similarly, if a person practices breathing with pleasant visualizations, soothing sounds, music, or aromas, these positive experiences may become associated with the practice. Furthermore, improved emotion regulation is necessary to reduce over reactivity and inappropriate defensive or aggressive behaviors.

Nostril Dominance The physiological effects of right or left unilateral nostril breathing and alternate nostril breathing (ANB) (alternating occlusion of the right and left nostrils) are briefly review. Here we briefly review physiological effects. It has been proposed that right nostril breathing activates the sympathetic system while left nostril breathing activates parasympathetic response. Preliminary evidence suggests that alternate nostril breathing improves autonomic balance. Research by Telles also indicates an increase in heart rate with right nostril breathing, a decrease in heart rate with ANB, and a further reduction in heart rate with left nostril breathing, supporting the theory of lateralization of autonomic effects (Telles et al., 1994; Raghuraj & Telles, 2008). These findings support the therapeutic potential of ANB to improve autonomic balance. A study by Ghiya and Lee (2012) compared the effects of paced breathing at 5 bpm to ANB at 5 bpm on autonomic measures in 20 healthy adults. Analysis revealed that both practices increased parasympathetic dominance and autonomic regulation of cardiac functioning as indicated by HRV, but there were no statistically significant differences between them. This data suggests that autonomic changes in response to ANB were primarily mediated by the breathing rate in subjects without prior experience in yoga breathing. It is likely that longer practice of ANB or the use of different durations for the inhale, hold and exhale would produce different results. For example, prolonging the outbreath would be more parasympathetically activating, thereby creating a greater calming effect. More research is needed as studies of ANB report mixed results and often do not adequately document the duration of the phases of the breath cycle being utilized. Based on the premise that sensory fibers in the right and left nostril lead to the cerebral cortex, it has been hypothesized that ANB creates cerebral brain wave asymmetry between the left and right hemisphere. Telles and colleague (2017) found no supporting evidence for this hypothesis, contradicting a previous study. At this time, the evidence for the concept that ANB leads to cortical lateralization is inconclusive. Sensory fibers in the nose (other than olfactory fibers that project to the olfactory bulb) articulate on the trigeminal nerve, which projects to synapses in the NTS. However, in a rodent study, anatomical and

39

 Psychophysiology

functional connections between the orbital bulb and the prelimbic prefrontal cortex (plPFC) played a key role in anxiety and freeze behaviors (Moberly et al., 2018).

Resistance Breathing Resistance to the flow of breath can be created by pursed lips, tightened muscles in the upper back of the throat, use of tongue positions, or clenched teeth. Breathing against airway resistance (resistive loading), such as during ujjayi (Ocean Breath), increases intrathoracic pressure. The authors hypothesized that this slight increase in pressure could further activate the vagal nerves (Brown & Gerbarg, 2005; Calabrese et al., 2000). Studies have shown that practicing ujjayi breathing for 6 to 12 weeks significantly improved parasympathetic function, HRV, RSA, and other cardiovascular parameters: heart rate, systolic blood pressure, diastolic blood pressure, pulse pressure, mean arterial pressure (Lathadevi et al., 2012; Mahour & Verma, 2017). In an open trial with 17 healthy, yoga naïve adults, slow breathing at 6 bpm with similar inspiration and expiration times was an effective and simple way to increase BRS and improve oxygen saturation. The addition of ujjayi breath conferred a limited additional benefit over slow breathing done at 6 bpm during normoxia (Mason et al., 2013). Ujjayi can be more difficult to learn, particularly on inhalation. Due to the short training period for learning ujjayi (10 minutes), the study subjects, may not have had time to become proficient enough in this form of resistance breathing. The added effort to perform the ujjayi may have counteracted the potential benefits. More research on resistance breathing is needed with attention to the comfort of participants with resistance breath practices.

Breath Holds As breath holds alter oxygen and carbon dioxide levels, they engender marked effects. Of note, there is a paucity of research on post-exhalation breath holds, hence, this section will focus on physiological changes that occur with breath holding after inhalation. These effects vary considerably based on the length of the breath hold. One study found that breath holds of 4 seconds increased alveolar-capillarymembrane-conductance, briefly increasing the rate of oxygen diffusion into the blood stream (Jorgenson et al., 2018). Long breath holds are unlikely to have a similar effect because oxygen is rapidly used by the body, hence, oxygen levels fall when not quickly replenished when inhalation is delayed too long. Rather, during longer holds of at least 29-52 seconds, the body diverts blood away from most organs towards the main cerebral artery ensuring a healthy supply of oxygen to the brain (Eichhorn et al., 2018). The natural increase in CO2 that occurs when it is not expelled during breath holds, supports this transient neuro adaptation via cerebral vascular dilation, enhancing oxygen delivery to neurons (Fierstra et al., 2013). Intriguingly, many yogic traditions teach hyperventilation methods (rapid forceful breathing) followed by breath holds, which we speculate may support increased neurophysiological and cardiorespiratory flexibility, unless rapid breath techniques are contra-indicated (please see rapid and forceful breathing section above). Hyperventilation induces hypocapnia, associated with the vasoconstriction of cerebral vasculature. When this is followed by a long breath hold with hypercapnia (CO2 retention) and the associated vasodilation, the net result could increase cerebral arterial reactivity, a marker of neurological health. As research has reported, when the natural hyperventilation that normally follows a long breathhold is prevented, blood flow in the main cerebral artery remains high and the brain is likely to sustain a state of increased vasodilation (Eichhorn et al., 2018). This confluence of factors may also support 40

 Psychophysiology

increased oxygenation. When these techniques are done in succession (hyperventilation followed by a long breath hold), the possibility of increased oxygenation and arterial cerebral reactivity could be of physiological value. It is important to note that prolonged breath holds increase stress on the heart and are contraindicated in individuals with cardiovascular disease (e.g., hypertension or cardiac arrhythmias), respiratory disease (e.g., asthma or chronic obstructive pulmonary disease, COPD), or if they are a regular smoker. Breath holds are also contraindicated during pregnancy, following recent surgery, and in people with aneurysms, uncontrolled hypertension, stroke or any other condition wherein decreased oxygenation or increased intrathoracic or intra-abdominal pressure could be harmful (Eichhorn et al., 2018; Skow et al., 2015; Brown et al., 2013). Given the possible changes in neural signaling and sympathetic activation associated with shifts in cerebral blood flow and partial pressure of O2 and CO2 (McKay et al., 2003), we advise individuals with heightened sensitivity to such changes, such as those with bipolar disorder or schizophrenia, to avoid breath holding. The length of breath holds also appears to affect metabolic rate. In one small study of experienced yoga practitioners, researchers found that short breath holds (10-12 percent of the entire breath period) led to a 52 percent increase in O2 consumption. When breath holding increased to 25 percent of the breath period, O2 consumption dropped by 19 percent, possibly in response to the body’s need to restore vital oxygen reserves (Telles & Desiraju, 1991). This shift in O2 consumption suggests a modification in cellular mechanisms that may be a short-term adaptation. As described above short-term hypoxia stimulates HIF-1, reducing the cellular need for oxygen and consequently the rate of consumption. During short breath holds, HIF-1 helps reduce oxidative stress and possibly reverse pre-existing damage, as it upregulates glutathione and other endogenous antioxidants (Samanta & Semenza, 2017). Unlike the healthy subjects in many studies of VRBPs, individuals with cardiorespiratory impairments may not be able to hold their breath for more than 6 or 10 seconds without becoming uncomfortable or symptomatic. Just as short periods of reduced oxygen consumption may increase allostasis, the physiological capacity to remain stable under conditions of stress, as develops with the regular practice of breath holds, improves tolerance of higher levels of CO2 and exercise (Spicuzza et al., 2000; Gabutti et al., 2001). Increased CO2 tolerance may also be helpful in the treatment of anxiety. Although the literature offers conflicting findings, some trials report a correlation between increased levels of CO2 and panic, anxiety and fear (Katzman et al., 2002; Taugher et al., 2014). When a person with panic disorder is ready to engage in breath-holding, something that can be initially anxiety provoking, the gradual increased tolerance may help reduce fear and over reactivity.

Public Health Prospects for Breath Practices Breath practices are garnering widespread attention in health care and entering the realm of public health. Breath awareness and VRBPs are now integrated into the third wave of cognitive behavioural therapy. The National Center for Complementary and Integrative Health (2007) reported that 38 percent of Americans used complementary alternative medicine with 12.7 percent of the total practicing deep breathing techniques. A 2012 survey found that 23 percent of all insomnia sufferer in the USA used relaxation techniques, deep breathing being the most prevalent. In the UK, the West London Clinical Commissioning Group commissioned Mason to teach coherent breathing to their staff for personal wellbeing and to teach to patients. The UK National Health Service recommends slow deep breathing for the management of depression and anxiety, as does the Western Australian Department of Health (https://www.nhs.uk/ 41

 Psychophysiology

conditions/stress-anxiety-depression/ways-relieve-/stress/). The Indian Ministry of Ayurveda, Yoga & Naturopathy, Unani, Siddha, Sowa Rigpa and Homoeopathy (AYUSH) has been preparing to promote yogic techniques throughout society as part of their National Health Strategy (2017). Breath practices may take center stage in this initiative, as the country is committed to medical pluralism. Breath practices can be taught in a secular manner and are readily accepted by diverse cultural groups. A study of 17,000 patients, conducted at Harvard, documented substantial cost savings associated with mind-body practices (Stahl et al., 2017). In one year, individuals who engaged in mind-body practices reduced health care utilization by 43 percent, saving over $2,300 per person. Considering that slow VRBPs are simple, safe, effective, versatile, and inexpensive, they are prime candidates for international health initiatives and should yield significant reductions in health expenditures. There are many pathways to disseminate breath practices throughout societies, including health campaigns to engage national audiences. Also, train-the-trainer can prepare community leaders, health professionals, first responders, and school teachers to learn how to instruct others in the art of breathing. Sustainability by trainers embedded in communities would be vital to success. This could be achieved by community groups meeting regularly to plan and practice VRBPs together. Obviously, resources would be needed to achieve country-wide integration of VRBPs into healthcare, school systems, and workplaces, but the cost/benefit ratio should be quite favorable. Increasingly robust evidence about the therapeutic effects of VRBPs should guide the development of effective interventions for individual, public, and global health.

REFERENCES Acker, T. (2004). Cellular oxygen sensing need in CNS function: Physiological and pathological implications. The Journal of Experimental Biology, 207(18), 3171–3188. doi:10.1242/jeb.01075 PMID:15299039 Adriaensen, D., Brouns, I., Pintelon, I., De Proost, I., & Timmermans, J. (2006). Evidence for a role of neuroepithelial bodies as complex airway sensors: Comparison with smooth muscle-associated airway receptors. Journal of Applied Physiology, 101(3), 960–970. doi:10.1152/japplphysiol.00267.2006 PMID:16741263 Bates, M., Lesnewich, L., Uhouse, S., Gohel, S., & Buckman, J. (2019). Resonance-paced breathing alters neural response to visual cues: Proof-of-concept for a neuroscience-informed adjunct to addiction treatments. Frontiers in Psychiatry, 10, 10. doi:10.3389/fpsyt.2019.00624 PMID:31543840 Bernardi, L., Gabutti, A., Porta, C., & Spicuzza, L. (2001). Slow breathing reduces chemoreflex response to hypoxia and hypercapnia and increases baroreflex sensitivity. Journal of Hypertension, 19(12), 2221–2229. doi:10.1097/00004872-200112000-00016 PMID:11725167 Bernardi, L., Porta, C., Spicuzza, L., Bellwon, J., Spadacini, G., Frey, A., Yeung, L. Y. C., Sanderson, J. E., Pedretti, R., & Tramarin, R. (2002). Slow breathing increases arterial baroreflex sensitivity in patients with chronic heart failure. Circulation, 105(2), 143–145. doi:10.1161/hc0202.103311 PMID:11790690 Bernardi, L., Spadacini, G., Bellwon, J., Hajric, R., Roskamm, H., & Frey, A. (1998). Effect of breathing rate on oxygen saturation and exercise performance in chronic heart failure. Lancet, 351(9112), 1308–1311. doi:10.1016/S0140-6736(97)10341-5 PMID:9643792

42

 Psychophysiology

Brown, R., & Gerbarg, P. (2005). Sudarshan Kriya yogic breathing in the treatment of stress, anxiety, and depression: Part I—neurophysiologic model. Journal of Alternative and Complementary Medicine (New York, N.Y.), 11(1), 189–201. doi:10.1089/acm.2005.11.189 PMID:15750381 Brown, R., & Gerbarg, P. (2012). The healing power of the breath. Simple techniques to reduce stress and anxiety, enhance concentration and balance your emotions. Shambhala Publications. Brown, R., & Gerbarg, P. (2017). Breathing techniques in psychiatric treatment. In P. L. Gerbarg, R. P. Brown, & P. R. Muskin (Eds.), Complementary and integrative treatments in psychiatric practice (pp. 241–250). American Psychiatric Association Publishing. Brown, R., Gerbarg, P., & Muench, F. (2013). Breathing practices for treatment of psychiatric and stressrelated medical conditions. The Psychiatric Clinics of North America, 36(1), 121–140. doi:10.1016/j. psc.2013.01.001 PMID:23538082 Calabrese, P., Perrault, H., Dinh, T., Eberhard, A., & Benchetrit, G. (2000). Cardiorespiratory interactions during resistive load breathing. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology, 279(6), R2208–R2213. doi:10.1152/ajpregu.2000.279.6.R2208 PMID:11080087 Carter, M., de Lecea, L., & Adamantidis, A. (2013). Functional wiring of hypocretin and LC-NE neurons: Implications for arousal. Frontiers in Behavioral Neuroscience, 7, 43. doi:10.3389/fnbeh.2013.00043 PMID:23730276 Cutz, E., Pan, J., Yeger, H., Domnik, N., & Fisher, J. (2013). Recent advances and contraversies on the role of pulmonary neuroepithelial bodies as airway sensors. Seminars in Cell & Developmental Biology, 24(1), 40–50. doi:10.1016/j.semcdb.2012.09.003 PMID:23022441 Descilo, T., Vedamurtachar, A., Gerbarg, P., Nagaraja, D., Gangadhar, B., Damodaran, B., Adelson, B., Braslow, L. H., Marcus, S., & Brown, R. P. (2010). Effects of a yoga breath intervention alone and in combination with an exposure therapy for post-traumatic stress disorder and depression in survivors of the 2004 South-East Asia tsunami. Acta Psychiatrica Scandinavica, 121(4), 289–300. doi:10.1111/j.16000447.2009.01466.x PMID:19694633 Eichhorn, L., Doerner, J., Luetkens, J., Lunkenheimer, J., Dolscheid-Pommerich, R., Erdfelder, F., Fimmers, R., Nadal, J., Stoffel-Wagner, B., Schild, H. H., Hoeft, A., Zur, B., & Naehle, C. P. (2018). Cardiovascular magnetic resonance assessment of acute cardiovascular effects of voluntary apnoea in elite divers. Journal of Cardiovascular Magnetic Resonance, 20(1), 40. doi:10.118612968-018-0455-x PMID:29909774 Gabutti, A., Spicuzza, L., Porta, C., & Bernardi, L. (2001). Functions and changes of chemoreflex in physiological and pathological conditions. Recenti Progressi in Medicina, 92(7-8), 433–445. PMID:11475784 Gerbarg, P., & Brown, R. (2011). Mind-body practices for recovery from sexual trauma. In T. BryantDavis (Ed.), Surviving sexual violence: A guide to recovery and empowerment (pp. 199–216). Rowman & Littlefield Publishers. Gerbarg, P., & Brown, R. (2016). Neurobiology and neurophysiology of breath practices in psychiatric care. The Psychiatric Times, 33(11), 22–25.

43

 Psychophysiology

Gerbarg, P., Brown, R., Streeter, C., Katzman, M., & Vermani, M. (2019). Breath practices for survivor and caregiver stress, depression, and post-traumatic stress disorder: Connection, co-regulation, compassion. OBM Integrative and Complementary Medicine, 4(3), 1–29. doi:10.21926/obm.icm.1903045 Gerbarg, P. L. (2008). Yoga and neuro-psychoanalysis. In F. S. Anderson (Ed.), Bodies in treatment: The unspoken dimension (pp. 127–150). The Analytic press, Inc. Gerbarg, P. L., & Brown, R. P. (2015). Yoga and neuronal pathways to enhance stress response, emotion regulation, bonding, and spirituality. In E. G. Horowitz & S. Elgelid (Eds.), Yoga therapy: theory and practice (pp. 49–64). Routledge Taylor and Francis Group. Godoy, D., Seifi, A., Garza, D., Lubillo-Montenegro, S., & Murillo-Cabezas, F. (2017). Hyperventilation therapy for control of post traumatic intracranial hypertension. Frontiers in Neurology, 8, 250. doi:10.3389/fneur.2017.00250 PMID:28769857 Gupta, R., Agnihotri, S., Telles, S., & Balkrishna, A. (2019). Performance in a Corsi block-tapping task following high-frequency yoga breathing or breath awareness. International Journal of Yoga, 12(3), 247. doi:10.4103/ijoy.IJOY_55_18 PMID:31543634 Hammer, P., & Saul, J. (2005). Resonance in a mathematical model of baroreflex control: Arterial blood pressure waves accompanying postural stress. American Journal of Physiology. Regulatory, Integrative and Comparative Physiology, 288(6), R1637–R1648. doi:10.1152/ajpregu.00050.2004 PMID:15718393 Hayashi, K., Fujikawa, M., & Sawa, T. (2008). Hyperventilation-induced hypocapnia changes the pattern of electroencephalographic bicoherence growth during sevoflurane anaesthesia. British Journal of Anaesthesia, 101(5), 666–672. doi:10.1093/bja/aen269 PMID:18791187 Herrero, J., Khuvis, S., Yeagle, E., Cerf, M., & Mehta, A. (2018). Breathing above the brain stem: Volitional control and attentional modulation in humans. Journal of Neurophysiology, 119(1), 145–159. doi:10.1152/jn.00551.2017 PMID:28954895 Jalife, J., Slenter, V., Salata, J., & Michaels, D. (1983). Dynamic vagal control of pacemaker activity in the mammalian sinoatrial node. Circulation Research, 52(6), 642–656. doi:10.1161/01.RES.52.6.642 PMID:6861283 Jandackova, V., Scholes, S., Britton, A., & Steptoe, A. (2016). Are changes in heart rate variability in middle‐aged and older people normative or caused by pathological conditions? Findings from a large population‐based longitudinal cohort study. Journal of the American Heart Association, 5(2), e002365. doi:10.1161/JAHA.115.002365 PMID:26873682 Jorgenson, C., Coffman, K., & Johnson, B. (2018). Effects of intrathoracic pressure, inhalation time, and breath hold time on lung diffusing capacity. Respiratory Physiology & Neurobiology, 258, 69–75. doi:10.1016/j.resp.2018.06.001 PMID:29885374 Joseph, C., Porta, C., Casucci, G., Casiraghi, N., Maffeis, M., Rossi, M., & Bernardi, L. (2005). Slow breathing improves arterial baroreflex sensitivity and decreases blood pressure in essential hypertension. Hypertension, 46(4), 714–718. doi:10.1161/01.HYP.0000179581.68566.7d PMID:16129818

44

 Psychophysiology

Katzman, M., Struzik, L., Vijay, N., Coonerty-Femiano, A., Mahamed, S., & Duffin, J. (2002). Central and peripheral chemoreflexes in panic disorder. Psychiatry Research, 113(1-2), 181–192. doi:10.1016/ S0165-1781(02)00238-X PMID:12467957 Kemp, A., & Quintana, D. (2013). The relationship between mental and physical health: Insights from the study of heart rate variability. International Journal of Psychophysiology, 89(3), 288–296. doi:10.1016/j. ijpsycho.2013.06.018 PMID:23797149 Khalsa, S., Adolphs, R., Cameron, O., Critchley, H., Davenport, P., Feinstein, J., ... Zucker, N. (2018). Interoception and mental health: A roadmap. Biological Psychiatry, 3(6), 501–513. PMID:29884281 Lathadevi, G., Maheswari, T., & Nagashree, R. (2012). Modulation of cardiovascular response after Ujjayi pranayama and Shavasana. Training in normal human volunteers. http://www.jcdr.net//back_issues.asp?issn=0973-709x&year=2012&month=May&volume=6&issue=4&page=571&id=2140 Lehrer, P., Vaschillo, E., & Vaschillo, B. (2000). Resonant frequency biofeedback training to increase cardiac variability: Rationale and manual for training. Applied Psychophysiology and Biofeedback, 25(3), 177–191. doi:10.1023/A:1009554825745 PMID:10999236 Lehrer, P., Vaschillo, E., Vaschillo, B., Lu, S., Eckberg, D., Edelberg, R., Shih, W. J., Lin, Y., Kuusela, T. A., Tahvanainen, K. U. O., & Hamer, R. M. (2003). Heart rate variability biofeedback increases baroreflex gain and peak expiratory flow. Psychosomatic Medicine, 65(5), 796–805. doi:10.1097/01. PSY.0000089200.81962.19 PMID:14508023 Mahour, J., & Verma, P. (2017). Effect of Ujjayi pranayama on cardiovascular autonomic function tests. National Journal of Physiology, Pharmacy and Pharmacology, 7(4), 391–395. doi:10.5455/ njppp.2017.7.1029809122016 Mason, H., Vandoni, M., deBarbieri, G., Codrons, E., Ugargol, V., & Bernardi, L. (2013). Cardiovascular and respiratory effect of yogic slow breathing in the yoga beginner: What is the best approach? EvidenceBased Complementary and Alternative Medicine, 2013, 1–7. doi:10.1155/2013/743504 PMID:23710236 McKay, L., Evans, K., Frackowiak, R., & Corfield, D. (2003). Neural correlates of voluntary breathing in humans. Journal of Applied Physiology, 95(3), 1170–1178. doi:10.1152/japplphysiol.00641.2002 PMID:12754178 Michiels, C. (2004). Physiological and pathological responses to hypoxia. American Journal of Pathology, 164(6), 1875–1882. doi:10.1016/S0002-9440(10)63747-9 PMID:15161623 Moberly, A., Schreck, M., Bhattarai, J., Zweifel, L., Luo, W., & Ma, M. (2018). Olfactory inputs modulate respiration-related rhythmic activity in the prefrontal cortex and freezing behavior. Nature Communications, 9(1), 1528. doi:10.103841467-018-03988-1 PMID:29670106 Nyer, M., Gerbarg, P., Silveri, M., Johnston, J., Scott, T., Nauphal, M., Owen, L., Nielsen, G. H., Mischoulon, D., Brown, R. P., Fava, M., & Streeter, C. C. (2018). A randomized controlled dosing study of Iyengar yoga and coherent breathing for the treatment of major depressive disorder: Impact on suicidal ideation and safety findings. Complementary Therapies in Medicine, 37, 136–142. doi:10.1016/j. ctim.2018.02.006 PMID:29609926

45

 Psychophysiology

Ogoh, S., Hayashi, N., Inagaki, M., Ainslie, P., & Miyamoto, T. (2008). Interaction between the ventilatory and cerebrovascular responses to hypo- and hypercapnia at rest and during exercise. The Journal of Physiology, 586(17), 4327–4338. doi:10.1113/jphysiol.2008.157073 PMID:18635644 Peper, E., Kawakami, M., Sata, M., Franklin, Y., Gibney, K., & Wilson, V. (2005). Two breaths per minute yogic breathing. In M. Kawakami (Ed.), The theses of mitsumasa kawakami II: The theory of yoga-based good health (pp. 483–493). Samskara. Pomidori, L., Campigotto, F., Amatya, T., Bernardi, L., & Cogo, A. (2009). Efficacy and tolerability of yoga breathing in patients with chronic obstructive pulmonary disease. Journal of Cardiopulmonary Rehabilitation and Prevention, 29(2), 133–137. doi:10.1097/HCR.0b013e31819a0227 PMID:19305239 Porges, S. (2001). The polyvagal theory: Phylogenetic substrates of a social nervous system. International Journal of Psychophysiology, 42(2), 123–146. doi:10.1016/S0167-8760(01)00162-3 PMID:11587772 Porges, S. (2003). Social engagement and attachment. Annals of the New York Academy of Sciences, 1008(1), 31–47. doi:10.1196/annals.1301.004 PMID:14998870 Porges, S., & Carter, C. (2017). Polyvagal theory and the social engagement system neurophysiological bridge between connectedness and health. In P. L. Gerbarg, R. P. Brown, & P. R. Muskin (Eds.), Complementary and Integrative Treatments in Psychiatric Practice (pp. 221–240). American Psychiatric Association Publishing. Porges, S., & Furman, S. (2010). The early development of the autonomic nervous system provides a neural platform for social behaviour: A polyvagal perspective. Infant and Child Development, 20(1), 106–118. doi:10.1002/icd.688 PMID:21516219 Raghuraj, P., & Telles, S. (2008). Immediate effect of specific nostril manipulating yoga breathing practices on autonomic and respiratory variables. Applied Psychophysiology and Biofeedback, 33(2), 65–75. doi:10.100710484-008-9055-0 PMID:18347974 Rosengård-Bärlund, M., Bernardi, L., Holmqvist, J., Debarbieri, G., Mäntysaari, M., af Björkesten, C., Forsblom, C., & Groop, P.-H. (2011). Deep breathing improves blunted baroreflex sensitivity even after 30 years of type 1 diabetes. Diabetologia, 54(7), 1862–1870. doi:10.100700125-011-2164-y PMID:21538174 Russo, M., Santarelli, D., & O’Rourke, D. (2017). The physiological effects of slow breathing in the healthy human. Breathe (Sheffield, England), 13(4), 298–309. doi:10.1183/20734735.009817 PMID:29209423 Samanta, D., & Semenza, G. (2017). Maintenance of redox homeostasis by hypoxia-inducible factors. Redox Biology, 13, 331–335. doi:10.1016/j.redox.2017.05.022 PMID:28624704 Siebenmann, C., Ryrsø, C., Oberholzer, L., Fisher, J., Hilsted, L., Rasmussen, P., Secher, N. H., & Lundby, C. (2019). Hypoxia-induced vagal withdrawal is independent of the hypoxic ventilatory response in men. Journal of Applied Physiology, 126(1), 124–131. doi:10.1152/japplphysiol.00701.2018 PMID:30496709 Skow, R., Day, T., Fuller, J., Bruce, C., & Steinback, C. (2015). The ins and outs of breath holding: Simple demonstrations of complex respiratory physiology. Advances in Physiology Education, 39(3), 223–231. doi:10.1152/advan.00030.2015 PMID:26330043

46

 Psychophysiology

Spicuzza, L., Gabutti, A., Porta, C., Montano, N., & Bernardi, L. (2000). Yoga and chemoreflex response to hypoxia and hypercapnia. Lancet, 356(9240), 1495–1496. doi:10.1016/S0140-6736(00)02881-6 PMID:11081541 Stahl, J., Dossett, M., LaJoie, A., Denninger, J., Mehta, D., Goldman, R., Fricchione, G. L., & Benson, H. (2017). Correction: Relaxation response and resiliency training and its effect on healthcare resource utilization. PLoS One, 12(2), e0172874. doi:10.1371/journal.pone.0172874 PMID:28222185 Steffen, P., Austin, T., DeBarros, A., & Brown, T. (2017). The impact of resonance frequency breathing on measures of heart rate variability, blood pressure, and mood. Frontiers in Public Health, 5, 222. doi:10.3389/fpubh.2017.00222 PMID:28890890 Streeter, C., Gerbarg, P., Whitfield, T., Owen, L., Johnston, J., Silveri, M., Gensler, M., Faulkner, C. L., Mann, C., Wixted, M., Hernon, A. M., Nyer, M. B., Brown, E. R. P., & Jensen, J. E. (2017). Treatment of major depressive disorder with Iyengar yoga and coherent breathing: A randomized controlled dosing study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 23(3), 201–207. doi:10.1089/acm.2016.0140 PMID:28296480 Strigo, I., & Craig, A. (2016). Interoception, homeostatic emotions and sympathovagal balance. Philosophical Transactions of the Royal Society B: Biological Sciences, 371(1708). Taha, B., Simon, P., Dempsey, J., Skatrud, J., & Iber, C. (1985). Respiratory sinus arrhythmia in humans: An obligatory role for vagal feedback from the lungs. Journal of Applied Physiology, 78(2), 638–645. doi:10.1152/jappl.1995.78.2.638 PMID:7759434 Taugher, R., Lu, Y., Wang, Y., Kreple, C., Ghobbeh, A., Fan, R., Sowers, L. P., & Wemmie, J. A. (2014). The bed nucleus of the stria terminalis is critical for anxiety-related behavior evoked by CO2 and acidosis. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 34(31), 10247–10255. doi:10.1523/JNEUROSCI.1680-14.2014 PMID:25080586 Telles, S., & Desiraju, T. (1991). Oxygen consumption during pranayamic type of very slow-rate breathing. The Indian Journal of Medical Research, 94, 357–363. PMID:1794892 Telles, S., Nagarathna, R., & Nagendra, H. (1994). Breathing through a particular nostril can alter metabolism and autonomic activities. Indian Journal of Physiology and Pharmacology, 38, 133–137. PMID:8063359 Telles, S., & Singh, N. (2013). Science of the mind: Ancient yoga texts and modern studies. The Psychiatric Clinics of North America, 36(1), 93–108. doi:10.1016/j.psc.2013.01.010 PMID:23538080 Thayer, J., Hansen, A., Saus-Rose, E., & Johnsen, B. (2009). Heart rate variability, prefrontal neural function, and cognitive performance: The neurovisceral integration perspective on self-regulation, adaptation, and health. Annals of Behavioral Medicine, 37(2), 141–153. doi:10.100712160-009-9101-z PMID:19424767 Thomas, L., & Ashcroft, M. (2019). Exploring the molecular interface between hypoxia-inducible factor signalling and mitochondria. Cellular and Molecular Life Sciences, 76(9), 1759–1777. doi:10.100700018019-03039-y PMID:30767037

47

 Psychophysiology

van de Vooren, H., Gademan, M., Swenne, C., TenVoorde, B., Schalij, M., & Van der Wall, E. (2007). Baroreflex sensitivity, blood pressure buffering, and resonance: What are the links? Computer simulation of healthy subjects and heart failure patients. Journal of Applied Physiology, 102(4), 1348–1356. doi:10.1152/japplphysiol.00158.2006 PMID:17185500 Vaschillo, E., Lehrer, P., Rishe, N., & Konstantinov, M. (2002). Heart rate variability biofeedback as a method for assessing baroreflex function: A preliminary study of resonance in the cardiovascular system. Applied Psychophysiology and Biofeedback, 27(1), 1–27. doi:10.1023/A:1014587304314 PMID:12001882 Wang, D., Clouter, A., Chen, Q., Shapiro, K., & Hanslmayr, S. (2018). Single-trial phase entrainment of theta oscillations in sensory regions predicts human associative memory performance. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 38(28), 6299–6309. doi:10.1523/ JNEUROSCI.0349-18.2018 PMID:29899027 Zhong, W., Ciatipis, M., Wolfenstetter, T., Jessberger, J., Müller, C., Ponsel, S., … Draguhn, A. (2017). Selective entrainment of gamma subbands by different slow network oscillations. Proceedings of the National Academy of Sciences, 114(17), 4519-4524. 10.1073/pnas.1617249114 Zope, S., & Zope, R. (2013). Sudarshan Kriya yoga: Breathing for health. International Journal of Yoga, 6(1), 4. doi:10.4103/0973-6131.105935 PMID:23440614

48

49

Chapter 3

Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama Ramakrishnan Angarai Ganesan https://orcid.org/0000-0002-3646-1955 Indian Institute of Science, Bangalore, India

ABSTRACT The benefits of pranayama for positive health are well known. Even though there are many studies published on the effectiveness of pranayama, there are very few papers that actually have systematically studied the physiological mechanisms involved, causing the benefits of pranayama, especially with respect to the cardiac function. This chapter attempts to have a detailed look at the physiology behind deep breathing. The chapter also conjectures that voluntary, deep breathing with attention may have a role to play in faster recovery from surgeries and prevent or delay the onset of Alzheimer’s disease, Parkinson’s disease, and maybe even cancer. Extended, carefully controlled, and detailed studies are needed to prove or disprove these conjectures.

INTRODUCTION Indian tradition has always extolled the practice of different types of deep yogic breathing techniques, generally known as pranayama, for better health of a human being. Taylor et al. (2010) attempted a framework to integrate the mechanisms of the bidirectional interaction of mind and body that mediate efficacy of mind-body therapies. Telles et al. (2014) have proposed that yoga regulated breathing is a bidirectional mind-body practice. A number of studies (Pramanik et al., 2009; Dhungel et al., 2008; Russo et al., 2017) have shown reduction of heart rate and blood pressure after different types of pranayama practices. However, not many have looked into the physiological and neural mechanisms of how yoga breathing is intimately connected to the cardiovascular, chemoreflex and endocrine systems of the human body, thus leading to positive health, especially that of the heart. A variety of deep yoga breathing exercises are in practice, such as one involving a sound coming from a constricted throat (ujjayi), alterDOI: 10.4018/978-1-7998-3254-6.ch003

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

nate nostril breathing (nadi shuddhi), forceful abdominal breathing (kapalabhati), cooling pranayama (sheetali), humming bee breathing (bhramari) and forced respiration with movement (bhastrika), which are performed at varying rates. In this article, the word pranayama is used in a very generic sense to denote a wide spectrum of many different types of deep yoga breathing practices and it needs to be emphasized that it clearly is not just one specific thing. Fast and deep breathing is activated by the autonomic nervous system during strenuous exercises such as aerobic exercise, climbing a steep mountain or sprinting. This is because the increased muscle activity demands increased oxygen and the cells need more oxygen to burn more calories to release energy. However, interestingly, it is the brain, rather than the skeletal muscles, which consume more oxygen in general. Since many of human occupations today involve brain work, rather than physical work, it is important to delve deeper into the physiological and control mechanisms of voluntary, deep breathing and the role of the latter in ensuring efficient functioning of the brain. This article carefully studies the respiratory physiology and the cardio-respiratory control circuits and identifies possible mechanisms behind the effect of yoga breathing in enhancing cardiac health (Ramakrishnan, 2019). In this article, deep yoga breathing refers to voluntarily taking deep breaths in a regular, systematic and near-periodic manner. It is proposed that voluntary, attentive, deep yoga breathing is a very effective cardiac exercise, while also having the potential to prevent many other ailments. While the article attempts to unravel some of the neural, physiological, chemical and hormonal connections, there may be other mechanisms playing a role in its effectiveness in assisting the overall health – including that of the brain. For example, Jerath et al. (2006) advance the hypothesis that only the stretch receptors are responsible for the effects observed.

NEURAL CONTROL AND PHYSIOLOGICAL PROCESSES BEHIND RESPIRATION Oxygen is a Critical Nutrition for the Brain and the Body The human body has approximately 130 trillion cells, working as a single unit, with total co-operation and understanding. Everyone starts her/his life as a single cell in their mother’s womb and the rest of the cells are put together systematically over a period of time by restructuring the food one consumes and the oxygen one breathes in. Almost 99 percent of the mass of the human body is made up of six elements, namely oxygen, carbon, hydrogen, nitrogen, calcium, and phosphorus. Only about 0.85 percent is composed of another five elements: potassium, sulphur, sodium, chlorine and magnesium. Oxygen is the most abundant element contained within living organisms, constituting about 65 percent of the human body (Emsley & John, 2011). Oxygen is required by each and every cell of the body for its metabolic activity. Even though the adult brain weighs less than 2 percent of the body weight, it consumes about 20 percent of the oxygen inhaled by the human beings. This makes it clear how crucial the level of oxygen in the body is in order that the brain and the body function efficiently. Carbon dioxide (CO2) is the waste gas produced when carbon is combined with oxygen as part of the body’s energy-making processes (metabolism) in each of the body’s cells. Oxygen is supplied and CO2 is removed from every cell by the circulation of blood. Respiration exchanges oxygen from the environment for carbon dioxide from the body’s cells. Thus, respiration is fundamental to the health of human brain. Brain cells start dying, if devoid of oxygen for more than 4 minutes. This is why people lose 50

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

control of speech or a part of the body after a stroke. Strokes happen when blood flow to a part of one’s brain stops, which in turn is caused by a blood clot that blocks a blood vessel in the brain. The nature of the loss of function is purely decided by which part of one’s brain was affected by the loss of oxygen.

What is involved in Oxygen Delivery to the Whole Body? Assuming a mean adult breathing rate of 14 cycles per minute, an adult human being breathes about 20,000 times every day, exchanging about 10,000 litres of air. Oxygen forms one of the principal reactants of aerobic respiration, which is the process used by all the cells to turn fuel, such as fats and sugars, which contain stored energy, into a usable form. In an adult an average of 20 trillion red blood cells carry the oxygen molecules from the lung to the individual cells of the body. Each red blood cell contains about 270 million oxygen-carrying- molecules or hemoglobin. These blood cells in turn are created by the bone marrow from the food that one consumes, after it has been digested in the alimentary canal, and absorbed through the small intestine. Hematopoietin secreted by the kidney triggers haemoglobin production in the cavities of the bone marrow. The output of respiration is the adenosine triphosphate molecule, which supplies energy to the cell for the chemical reactions it needs to survive. Bone marrow cavities in a human body produce approximately 2.5 million red blood cells per second. Brainstem, through the cervical nerves, triggers the muscles of respiration in the chest. Oxygen enters the haemoglobin cells through diffusion across the two-cell wall between the capillaries and the hundreds of millions of tiny sacs (alveoli) in the lung, with a total surface area of 100 square meters. The erythrocytes carry the oxygen through the cardiovascular system; the total length of the capillaries in a human body is several times the circumference of the earth. Assuming an average heart rate of 70 beats per minute, the heart beats about 100,000 times a day. Thus, the brain, lungs, blood vessels, bones, stomach, intestines and heart work constantly to deliver oxygen to the cells throughout the body to keep the human machinery going.

Oxygen Intake Can Be Increased by Deep, Yoga Breathing In normal (involuntary, passive) human breathing, the volume of air displaced between inhalation and exhalation (called the tidal volume) is about 500 ml or 7 mL/kg of body mass for a healthy, young adult (Beardsell et al., 2009). With intentional deep breathing with maximal effort, one can reach the vital capacity (maximum amount of air that can be moved in and out of the lungs in a respiratory cycle) of about 5 litres or 70 ml/kg of body mass (Family Practice Notebook, 2020) per respiration. This is almost 10 times the tidal volume or a 900 percent increase in the effectiveness of breathing OR the volume of air displaced per breathing cycle. Together with holding the breath after inhalation taught as part of advanced pranayama practices, this can ensure extremely effective gas exchange, better eliminating the carbon-dioxide from the body Bernardi et al. (1998) report significant increase in mean SaO2 at breathing rates of 15, 6 and 3 breaths per minute in both controls and chronic heart failure patients. Westerdahl et al. (2005) have reported that patients performing deep-breathing exercises after coronary artery bypass graft surgery had significantly smaller atelectatic areas and better pulmonary function on the fourth postoperative day compared to a control group performing no exercises.

51

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

Number of Muscles Involved in Inspiration Varies with the Type of Breathing In the literature, yoga breathing has never been explicitly considered as an exercise. However, this is because of a lack of focused study or observation of the physiology involved. Normal, involuntary inspiration is primarily caused by the contraction of the eleven pairs of external intercostal muscles and the diaphragm, which add up to 23 muscles. The left and right sternocleidomastoid, serratus anterior, pectoralis major and minor, the scalene (anterior, middle and posterior scalene) and latissimus dorssi muscles assist in elevating the rib cage and are considered accessory muscles of breathing. These nine muscles get involved depending upon the level of effort put in by the subject during inspiration. Thus, deep inspiration in yoga breathing can recruit up to 32 muscles. Specific pranayama practices involve voluntary movement of fingers, hands, abdomen etc. In certain schools of yoga such as the Art of Living’s Sudarshana Kriya, bhastrika pranayama is performed by lifting both the arms along with opening of all the fingers during inspiration and bringing the arms down to a position, where the elbows are bent and a fist is made by both the palms (Art of Living, 2020). This involves 8 muscles of each hand during inspiration. Infraspinatus is a rotator cuff muscle that helps raise the upper arm and triceps brachii straightens the arm. The extensor muscles found in the posterior side of the forearm extend the fingers. The extensor pollicis brevis, extensor pollicis longus and abductor pollicis longus muscles help extend the thumb. The extensor indicis (index finger), extensor digitorum (middle and ring fingers), and extensor digiti minimi (little finger) muscles extend the digits to open the hand. Thus, raising both the arms along with forceful inspiration during bhastrika takes the total number of muscles activated to 48.

Number of Muscles involved in Expiration Varies Widely from Passive to Active Breathing Normal expiration is passive, accomplished by the recoil of the thoracic cage and may hardly involve any muscle contraction. However, conscious, deep expiration is accomplished by the contraction of the eleven pairs of innermost intercostal muscles. The eleven pairs of internal intercostal muscles can add force to the exhalation process by lowering the rib cage. In addition, the four abdominal muscles (rectus abdominus, internal and external obliques and transverse abdominus) can press the abdominal organs upward into the diaphragm, thus diminishing the volume of the thoracic cavity. These abdominal muscles are the main ones involved, when one performs the quick exhalations in kapalabhati. Kapalabhati is a specific type of breathing suggested in Gheranda Samhita (Vasu, 2012), where the exhalation is active and the inhalation is passive, the opposite of normal breathing. Thus, deep, forceful expiration in pranayama can recruit up to 48 muscles. In the case of forced exhalation during bhastrika, the arm is brought down and the hand is made into a fist, which involves five more muscles, taking the total to 53 muscles. The muscles flexor digitorum superficialis, flexor digitorum profundus, and flexor pollicis longus, which extend from the bones of the arm and forearm and insert into the phalanges of the hand, flex the fingers and thumb and the brachialis brachioradialis and biceps brachii muscles bend the elbow (Barclay, 2020). In addition to the above muscles, the following 8 muscles have also been observed to contribute to respiration: trapezius, erector spinae, iliocostalis lumborum, quadratus lumborum, serratus posterior superior, serratus posterior inferior, levatores costarum, transversus thoracis, subclavius (Kendall et al., 2005). Thus, by forced inhalation and exhalation with movement of the shoulder, chest, arms and fingers, one can activate up to 109 muscles in all, forming about 16 percent of all the muscles in the body. 52

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

Control of Involuntary and Voluntary Respiration by the Central Nervous System The principal control center responsible for involuntary respiration is the medulla oblongata. The muscles that cause inspiratory and expiratory movements are directly controlled by it (Lumencandela, 2019). Coughing and sneezing are non-respiratory air movements, which are reflexes controlled by the medulla. The other respiratory center is the pons, which primarily decides the rate of involuntary respiration. When breathing occurs under conscious control, such as in any pranayama practice, the voluntary respiration is under the control of the primary motor cortex. However, chemoreceptor signals from involuntary respiration can override the signals from the motor cortex. In addition, whenever one perceives any danger or experiences emotional stress, the hypothalamus takes control and the respiratory rate is increased to handle the fight or flight response. Muscles involved in respiration are controlled by eighteen (five cervical, twelve thoracic and one lumbar) pairs of the total thirty one pairs of the spinal nerves. The accessory muscles are controlled by the three pairs of cervical nerves C1 to C3, and the diaphragm is controlled by C3 to C5. The eleven pairs of thoracic nerves T1 to T11 are responsible for the contraction of the 11 pairs each of the external, internal and the innermost intercostal muscles. The abdominal muscles are innervated by the thoracic nerves T6 to T12 and the lumbar nerve L1. This involvement of 36 spinal nerves in breathing clearly shows the importance given by nature to the process of respiration.

Acid-alkaline Balance in the Body and Pranayama The pH value of a solution is a measure of the acidity or alkalinity of a solution on a logarithmic scale, defined as, pH = −log10 CH, where CH is the hydrogen ion concentration in moles per litre. A pH value of 7 is neutral; lower values are more acidic and higher values, more alkaline. Carbon-dioxide dissolved in the tissue fluids increases the H-ion concentration as follows. H2O + CO2 ➔ H2CO3 ➔ H+ + HCO3- (bicarbonate) As the level of CO2 increases in the body, so does the H+ concentration, and the pH reduces. In other words, the acidity increases. The normal range of pH of blood is 7.35-7.45 (alkaline) and acidic pH is bad for health. A continuous blood pH below 7.0 can be fatal; however, the body tightly regulates the arterial blood pH levels. Thus, flushing out CO2 by active respiration can support this mechanism, which is better for the overall health of the body.

Role of Chemoreceptors, Baroreceptors and Glands in Cardiorespiratory Control The central chemoreceptors in the medulla detect the pH level of the spinal fluid as an indirect measure of the level of carbon-dioxide in the blood. The medulla and pons integrate the signals from these chemosensors and change the ventilation volume and rate in order to increase or decrease the removal of carbon dioxide. The peripheral chemoreceptors in the aortic and carotid bodies also perform a similar function. There are also baroreceptors in the carotid sinus and the aortic arch, which constantly sense the changes in the blood pressure by sensing the changes to the tension of the arterial walls.

53

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

Daly and Bondurant (1962) showed that breathing pure oxygen decreases the heart rate. Deep breathing naturally results in increased oxygen and reduced carbon-dioxide levels in the blood and consequently, reduced heart rate and pumping force of the heart due to the feedback given by the chemoreceptors to the cardiorespiratory control centres located in the pons and medulla oblongata in the brainstem. Pinna et al. (2000) have given a simplified model, showing the interactions between the cardiorespiratory systems, the neural control and the chemosensors. The hormones epinephrine and norepinephrine released by the adrenal glands help balance the heart rate, blood pressure and stress response, including the fight-or-flight response. The hormone aldosterone released by the adrenal glands affects the body’s ability to regulate blood pressure, by controlling the amount of sodium released by the kidney and colon into the bloodstream or potassium released in the urine. The thyroid hormones, released by the thyroid gland increase the heart rate, cardiac contractility and cardiac output. They also promote vasodilation, which leads to enhanced blood flow to many organs. The adrenal and thyroid glands are controlled by the pituitary, which in turn is controlled by the hypothalamus. Thus, the whole process of control of heart rate and blood pressure includes the complex feedback circuitry involving the baroreceptors, chemoreceptors and the level of hormones. Thus, it can be argued that deep breathing is an indirect exercise, activating all these circuitry.

Pranayama can be Considered a Cardiorespiratory Exercise Respiratory sinus arrhythmia is a physiological phenomenon reflecting respiratory-circulatory interactions universally observed among vertebrates (Anrep et al., 1936). The heart rate goes up every time one inhales, and goes down every time one exhales, and the variation can easily go up to 10–20 beats/ min in children and teenagers. It is thought that nature evolved this mechanism to increase the effectiveness of perfusion during inhalation and save energy expenditure during exhalation. However, recently it has been observed that the systematic fluctuations in heart rate occur even during long breath holds after inhalation (Ramakrishnan and Adarsh, 2020). Thus, it can be said that because of the respiratory sinus arrhythmia alone, active, deep yoga respiration constantly exercises the cardiac muscle, alternately increasing and decreasing the heart rate. It must be borne in mind that the circulatory system takes the increased oxygen to each of the body’s 130 trillion cells in every cardiac cycle. Thus, deep inhaling and exhaling pranayama practices such as nadi suddhi and bhastrika can be considered good cardiorespiratory exercises.

Possible Role of Additional Nitric Oxide Released from the Sinuses During Bhramari Nitric oxide (NO) can be produced on demand by the lining of every artery called the endothelium. NO is a pluripotent gaseous messenger with potent vasodilating and antimicrobial activity (Lundberg, 2008). However, paranasal sinus epithelium expresses an inducible nitric oxide synthase that continuously generates large amounts of NO (Lundberg et al., 1995), a highly bioactive signalling molecule. NO gas from the nose and sinuses is inhaled with every breath and reaches the lungs in a more diluted form to enhance pulmonary oxygen uptake via local vasodilation. In this sense, NO may be regarded as an ‘aerocrine’ hormone that is produced in the nose and sinuses and transported to a distal site of action with every inhalation. 54

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

Lundberg (Weitzberg & Lundberg, 2002; Lundberg, 2008) showed that humming increases the release of NO by a factor 15. The bhramari pranayama is exactly based on humming. Practice of bhramari has been suggested (Kuppusamy et al., 2016) and shown (Pramanik et al., 2020) to reduce blood pressure. It requires detailed studies to look at whether the additional NO reaches the blood and the cerebrospinal fluid (CSF). Increase of NO in the circulating blood is bound to cause reduction in the blood pressure through its vasodilation effects on the epithelium of all the capillaries in the body. If the NO reaches the CSF, it can cause additional benefits in the brain, including participation in neurogenesis in structures such as hippocampus (Shen et al., 2019).

POSSIBLE BENEFICIAL ROLE OF PRANAYAMA IN SPECIFIC AILMENTS Can Pranayama Accelerate Recovery from Surgery? Wounds need oxygen to heal. Continuous supply of oxygen to the tissue through microcirculation is vital for the healing process and for resistance to infection (Health, 2020). One of the biggest factors that can inhibit the body’s ability to recover is low oxygen flow to the affected area. Hence, the evaluation of tissue perfusion and oxygenation is important in all types of wounds. The Johns Hopkins hospital uses hyperbaric oxygen therapy (HBOT) for wound healing (Health, 2020). One normally thinks only of cuts, falls and accidents when one hears the word ‘wound’. However, all surgeries result in wounds that take time to heal. Thus, it is the conjecture of the author that recovery from surgeries may be accelerated by deep breathing.

Role of Pranayama in Cancer Care The Nobel-prize winning biochemist (The Noble Prize, 2020) showed that cancer basically needs a low oxygen environment to survive. He suggested that cancer cells “live in hypoxic, very low oxygen, and acidic conditions and derive energy from sugars by fermenting them the way yeast does.” From this, he theorized that these low-oxygen and highly-acidic conditions caused cancer. Recently, scientists at The Johns Hopkins University, based on their experiments conducted on human breast cancer cells and mice, explain how certain cancer stem cells thrive in low oxygen conditions (Johns Hopkins Medicine, 2016). Thus, it is the author’s conjecture that when the body cells have sufficient oxygen, cancer may be prevented or at least delayed. Raa and colleague (2007) have reported that tumor growth was significantly reduced after hyperbaric oxygen treatment compared to control and even more than chemotherapy in mammary tumors inducted in rats. Recent review by Danhauer and colleague (2017) of studies on effect of yoga interventions including breathing exercises on cancer patients during treatment by chemotherapy concludes that there is evidence to recommend yoga for improving psychological outcomes (such as reduction of anxiety, depression and distress), with potential for also improving physical symptoms.

The Promise of Epigenetic Changes Due to Deep, Yoga Breathing Practices The recent major meta-analysis of 569 different genome-wide association studies (reported over a period of two decades) by Patron et al. (2019) from University of Alberta has revealed that not more than 5 to 55

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

10 percent of risk in many cancers, diabetes and Alzheimer’s disease can be attributed to the contributions of common gene mutations (single nucleotide polymorphisms). The authors suggest that the risks for a majority of diseases arise from one’s metabolism, environment, lifestyle, or exposure to different nutrients, chemicals, microorganisms or sub microscopic infectious agents. Thus, epigenetics offers new possibilities to the humanity, which earlier believed that one’s destiny is fully controlled by one’s genes that completely decided whether one would develop a disease or not. Using a controlled study of ten subjects, Qu and colleague (2013) showed that two hours of Sudarshan Kriya and related practices can cause changes in the expression of 111 genes in the circulating immune cells (peripheral blood mononuclear cells) in healthy people, as compared to only 38 genes in the case of the control subjects, who had a nature walk and listened to relaxing music. Buric and others (2017) based on their review of eighteen studies on mind-body interventions (MBI) such as mindfulness, yoga and breath regulation, report that these practices are associated with a downregulation of nuclear factor kappa B pathway, which is the opposite of the effects of chronic stress on gene expression and hence, MBI practices may lead to a reduced risk of inflammation-related diseases. Thus, while the DNA is not a good predictor of one’s health, practices such as pranayama hold promise for improvement of human health through epigenetic changes. Definitely, many more such studies are needed to explore and confirm the exciting possibilities that epigenetics, induced by breathing practices, promise. It goes without saying that it is ultimately important to look at the physiological and the molecular mechanisms that lead to such alterations in gene expression profiles in the circulating immune cells.

Alzheimer’s Disease and Modulation of the Circulation of CSF by Breathing Beta amyloid is produced as a waste product of metabolic activity in the brain. It is a toxic protein associated with Alzheimer’s disease (National Institute on Aging, 2017). In Alzheimer’s disease, beta-amyloid clumps together to form amyloid plaques, negatively impacting communication between neurons. It is required to study whether improved circulation of CSF in the brain may aid in the effective elimination of metabolic waste products such as beta-amyloids from the brain, thus reducing the risk of Alzheimer’s disease. Interestingly, Heck and others (2017) have shown that neuronal oscillations follow the respiratory rhythm in the sense that the increase in gamma power is phase-locked to breathing. They propose that the electrical activity of different functional areas of the cortex are modulated by the breathing-driven proprioceptive and interoceptive inputs to the brain. A very recent study by Delaidelli (2017) has shown that breathing modulates the circulation of CSF in the brain. This is a finding with far-reaching implications. Focused research is needed to see whether, with regular practice of deep breathing, one may be able to improve the circulation of CSF in the brain and either prevent or delay the onset of Alzheimer’s disease.

Pranayama may Delay the Onset of Parkinson’s: A Conjecture Parkinson’s disease involves loss of dopaminergic neurons in substantia nigra pars compacta (SNc) (Olanow & Tatton, 1999). Due to their structural and functional properties, neurons in the SNc are one of the most vulnerable and energy consuming neurons. It has been proposed that the primary factor that causes the degeneration of SNc neurons is their high metabolic requirements (Pissadaki & Bolam, 2013). A very recent paper (Muddapu et al., 2019) based on computational modelling of basal ganglia,

56

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

suggests that this process of degeneration of SNc neurons might be initiated by weak-excitotoxicity mediated by energy deficit. It is well known that if the human body is furnished with adequate nutrients, then it has the necessary adaptive mechanisms in terms of feedback and preferential supply, to make sure that the necessities of tissues that require enhanced nutrition are met. Thus, once again, controlled studies are needed to investigate if regular practice of deep yoga breathing can boost oxygen delivery to the brain and thus result in delay or prevention of the onset of Parkinson’s disease.

CONCLUSION Good breathing is an essential part of nutrition for the brain and the body. Deep yoga breathing is a cardiorespiratory exercise, with a potential to recruit over 100 muscles, 36 spinal nerves and a number of peripheral and central chemoreceptors, volume receptors, besides the adrenal and thyroid glands. Further, it can increase the oxygen intake considerably. With increased oxygen supply in the blood, the heart needs to work less and thus its health improves; the heart rate and the blood pressure ought to reduce with longitudinal practice of pranayama. If the recent findings of CSF flow modulation by respiration are confirmed by further studies, then regular yoga breathing has the potential to provide sufficient supplies and also delay or prevent the accumulation of beta amyloids in the brain, thus delaying and may be preventing Parkinson’s and Alzheimer’s diseases. Finally, it may be good to mention here that there is some belief that breath retention, especially after exhalation, if not practiced properly, may lead to some health issues. However, this is more a rumour and has not been proved by any studies. Nevertheless, it may be advisable to practise the breath retention techniques only after systematic learning from a yoga expert.

REFERENCES Anrep, G., Pascual, W., & Rössler, R. (1936). Respiratory variations of the heart rate - I—The reflex mechanism of the respiratory arrhythmia. Proceedings of the Royal Society of London. Series B, Biological Sciences, 119(813), 191–217. doi:10.1098/rspb.1936.0005 Barclay, T. (2019). Muscles of the hand and the wrist. https://www.innerbody.com/image_skel13/ligm27. html Beardsell, I., Bell, S., Robinson, S., & Rumbold, H. (2009). MCEM Part A: MCQs. Royal Society of Medicine Press. Bernardi, L., Spadacini, G., Bellwon, J., Hajric, R., Roskamm, H., & Frey, A. W. (1998). Effect of breathing rate on oxygen saturation and exercise performance in chronic heart failure. Lancet, 351(9112), 1308–1311. doi:10.1016/S0140-6736(97)10341-5 PMID:9643792 Buric, I., Farias, M., Jong, J., Mee, C., & Brazil, I. A. (2017). What is the molecular signature of mindbody interventions? a systematic review of gene expression changes induced by meditation and related practices. Frontiers in Immunology, 8, 670. doi:10.3389/fimmu.2017.00670 PMID:28670311

57

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

Daly, W. J., & Bondurant, S. (1962). Effects of oxygen breathing on the heart rate, blood pressure, and cardiac index of normal men—Resting, with reactive hyperemia, and after atropine. The Journal of Clinical Investigation, 41(1), 126–132. doi:10.1172/JCI104454 PMID:13883265 Danhauer, S. C., Addington, E. L., Sohl, S. J., Chaoul, A., & Cohen, L. (2017). Review of yoga therapy during cancer treatment. Supportive Care in Cancer, 25(4), 1357–1372. doi:10.100700520-016-3556-9 PMID:28064385 Delaidelli, A., & Moiraghi, A. (2017). Respiration: A New Mechanism for CSF Circulation? The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 37(30), 7076–7078. doi:10.1523/ JNEUROSCI.1155-17.2017 PMID:28747391 Dhungel, Upadhyay, Malhotra, Sarkar, & Prajapati. (2008). Effect of alternate nostril breathing exercise on cardiorespiratory functions. Nepal Medical College Journal, 10, 25–27. PMID:18700626 Emsley, J. (2011). Nature’s building blocks: An a-z guide to the elements. Oxford University Press. Family Practice Notebook. (2020). Vital capacity. https://fpnotebook.com/Lung/Lab/VtlCpcty.htm Health. (2020). Hyperbaric oxygen therapy. https://www.hopkinsmedicine.org/health/treatment-testsand-therapies/hyperbaric-oxygen-therapy Heck, D. H., McAfee, S. S., Liu, Y., Babajani-Feremi, A., Rezaie, R., Freeman, W. J., Wheless, J. W., Papanicolaou, A. C., Ruszinkó, M., Sokolov, Y., & Kozma, R. (2017). Breathing as a fundamental rhythm of brain function. Frontiers in Neural Circuits, 10, 115. doi:10.3389/fncir.2016.00115 PMID:28127277 Jerath, R., Edry, J. W., Barnes, V. A., & Jerath, V. (2006). Physiology of long pranayamic breathing: Neural respiratory elements may provide a mechanism that explains how slow deep breathing shifts the autonomic nervous system. Medical Hypotheses, 67(3), 566–571. doi:10.1016/j.mehy.2006.02.042 PMID:16624497 Johns Hopkins Medicine. (2016). How cancer stem cells thrive when oxygen is scarce. Science Daily. https://www.sciencedaily.com/releases/2016/03/160328100159.htm Kendall, F., McCreary, E., Provance, P., Rodgers, M., & Romai, W. (2005). Muscles testing and function with posture and pain. Lippincott Williams & Wilkins. Kuppusamy, M., Kamaldeen, D., Pitani, R., & Amaldas, J. (2016). Immediate effects of bhramari pranayama on resting cardiovascular parameters in healthy adolescents. Journal of Clinical and Diagnostic Research: JCDR, 10, CC17–CC19. doi:10.7860/JCDR/2016/19202.7894 PMID:27437210 Lumencandela. (2019). Respiratory control. https://courses.lumenlearning.com/boundless-ap/chapter/ respiration-control/ Lundberg, J. O. (2008). Nitric oxide and the paranasal sinuses. The Anatomical Record, 291(11), 1479–1484. doi:10.1002/ar.20782 PMID:18951492 Lundberg, J. O., Farkas-Szallasi, T., Weitzberg, E., Rinder, J., Lidholm, J., Änggåard, A., Hökfelt, T., Lundberg, J. M., & Alving, K. (1995). High nitric oxide production in human paranasal sinuses. Nature Medicine, 1(4), 370–373. doi:10.1038/nm0495-370 PMID:7585069

58

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

Muddapu, V. R., Mandali, A., Chakravarthy, V. S., & Ramaswamy, S. (2019). A computational model of loss of dopaminergic cells in Parkinson’s disease due to glutamate-induced excitotoxicity. Frontiers in Neural Circuits, 13, 11. doi:10.3389/fncir.2019.00011 PMID:30858799 National Institute on Aging. (2017). What happens to the brain in Alzheimer’s disease. https://www.nia. nih.gov/health/what-happens-brain-alzheimers-disease Olanow, C. W., & Tatton, W. G. (1999). Etiology and pathogenesis of Parkinson’s disease. Annual Review of Neuroscience, 22(1), 123–144. doi:10.1146/annurev.neuro.22.1.123 PMID:10202534 Patron, J., Serra-Cayuela, A., Han, B., Li, C., & Wishart, D. S. (2019). Assessing the performance of genome-wide association studies for predicting disease risk. PLoS One, 14(12), e0220215. doi:10.1371/ journal.pone.0220215 PMID:31805043 Pinna, G. D., Maestri, R., Mortara, A., & La Rovere, M. T. (2000). Cardiorespiratory interactions during periodic breathing in awake chronic heart failure patients. American Journal of Physiology. Heart and Circulatory Physiology, 278(3), H932–H941. doi:10.1152/ajpheart.2000.278.3.H932 PMID:10710362 Pissadaki, E. K., & Bolam, J. P. (2013). The energy cost of action potential propagation in dopamine neurons: Clues to susceptibility in Parkinson’s disease. Frontiers in Computational Neuroscience, 7, 13. doi:10.3389/fncom.2013.00013 PMID:23515615 Pramanik, T., Pudasaini, B., & Prajapati, R. (2010). Immediate effect of a slow pace breathing exercise Bhramari pranayama on blood pressure and heart rate. Nepal Medical College Journal, 12, 154–157. PMID:21446363 Pramanik, T., Sharma, H. O., Mishra, S., Mishra, A., Prajapati, R., & Singh, S. (2009). Immediate effect of slow pace bhastrika pranayama on blood pressure and heart rate. Journal of Alternative and Complementary Medicine (New York, N.Y.), 15(3), 293–295. doi:10.1089/acm.2008.0440 PMID:19249921 Qu, S., Olafsrud, S. M., Meza-Zepeda, L. A., & Saatcioglu, F. (2013). Rapid gene expression changes in peripheral blood lymphocytes upon practice of a comprehensive yoga program. PLoS One, 8(4), e61910. doi:10.1371/journal.pone.0061910 PMID:23613970 Raa, A., Stansberg, C., Steen, V. M., Bjerkvig, R., Reed, R. K., & Stuhr, L. E. (2007). Hyperoxia retards growth and induces apoptosis and loss of glands and blood vessels in DMBA-induced rat mammary tumors. BMC Cancer, 7(1), 23. doi:10.1186/1471-2407-7-23 PMID:17263869 Ramakrishnan, A. G. (2019). Physiological mechanisms behind the effectiveness of pranayama. OSF Preprint: https://osf.io/eufqp/ Ramakrishnan, A. G., & Adarsh, A. (2020). R-wave amplitude changes and atypical heart rate changes accompanying breath hold during low breathing rates. In Proceedings of the 42nd Annual International Conference of the IEEE Engineering in Medicine and Biology Society. EMBS Virtual Academy. Russo, M. A., Santarelli, D. M., & O’Rourke, D. (2017). The physiological effects of slow breathing in the healthy human. Breathe (Sheffield, England), 13(4), 298–309. doi:10.1183/20734735.009817 PMID:29209423

59

 Cardiorespiratory and Endocrine Mechanisms Behind the Effectiveness of Pranayama

Shen, J., Wang, D., Wang, X., Gupta, S., Ayloo, B., Wu, S., Prasad, P., Xiong, Q., Xia, J., & Ge, S. (2019). Neurovascular coupling in the dentate gyrus regulates adult hippocampal neurogenesis. Neuron, 103(5), 878–890.e3. doi:10.1016/j.neuron.2019.05.045 PMID:31257104 Taylor, A. G., Goehler, L. E., Galper, D. I., Innes, K. E., & Bourguignon, C. (2010). Top-down and bottomup mechanism in mind-body medicine: Development of an integrative framework for psychophysiological research. Explore (New York, N.Y.), 6(1), 29–41. doi:10.1016/j.explore.2009.10.004 PMID:20129310 Telles, S., Singh, N., & Balkrishna, A. (2014). Role of respiration in mind-body practices: Concepts from contemporary science and traditional yoga texts. Frontiers in Psychiatry, 5, 167. doi:10.3389/ fpsyt.2014.00167 PMID:25505427 The Art of Living. (2020). Meditation: What is sudarshan kriya. https://www.artofliving.org/in-en/ what-sudarshan-kriya The Nobel Prize. (2020). Otto Warburg Biographical. https://www.nobelprize.org/prizes/medicine/1931/ warburg/biographical/ Vasu, S. C. (2012). The Gheranda sanhita: A treatise on hatha yoga [Translated]. Newman Press. Weitzberg, E., & Lundberg, J. O. (2002). Humming greatly increases nasal nitric oxide. American Journal of Respiratory and Critical Care Medicine, 166(2), 144–145. doi:10.1164/rccm.200202-138BC PMID:12119224 Westerdahl, E., Lindmark, B., Eriksson, T., Hedenstierna, G., & Tenling, A. (2005). Deep-breathing exercises reduce atelectasis and improve pulmonary function after coronary artery bypass surgery. Chest, 128(5), 3482–3488. doi:10.1378/chest.128.5.3482 PMID:16304303

60

61

Chapter 4

Neurobiology of Meditation Danilo Forghieri Santaella Sports Center, University of São Paulo, Brazil

ABSTRACT Meditation should not be considered a simple activity that is performed with focused attention; this is concentration. When practicing concentration correctly, with a good “anchor” for attention, a specific state of mind takes place, in which logic relaxation happens, and there is a relative freedom from selfidentification. Such states of mind are to be experienced and cannot be practiced; thus, meditation techniques (concentrations) are the means to reach this goal. Those who achieve such a state experience positive neurophysiological effects, which have been studied for decades, such as increased functionality and connectivity of the brain, and also increased gray matter volume in specific cortical areas, whether in the young or in the elderly. Meditation has, thus, a proven potential role to help one maintain a healthy cognition and should be included in daily life routines of everybody who wishes for it.

INTRODUCTION Even though meditation is a millennial practice such as yoga itself, during the past two decades it has underwent a huge growth of interest amongst both Easterners and Westerners. This has brought up many different approaches towards meditation and meditation-like states and techniques, as well as the interest of the scientific society in investigating its effects both on psychological and physiological variables. Nevertheless, although the growing interest and practice of meditation may be beneficial, it also creates the necessity of some methodological definitions and guidelines for the sake of increasing its positive effects and also to foster possible comparisons of results amongst papers and investigations. In this chapter we will elaborate on the neurobiological effects of meditation, after we come to a definition of what we consider it to be.

DOI: 10.4018/978-1-7998-3254-6.ch004

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Neurobiology of Meditation

WHAT IS MEDITATION? Exactly in the second sutra given by Sri Patañjali – who supposedly lived in the 2nd century B. C. (Taimni, 2014), the great yoga master states that yoga is meditation, since it is said that “Yoga is the cessation of the agitations of the mind.” In fact, the whole Yogasutras are meant to be a practical guidebook towards achieving the state of advaita (non-duality), or the experience of interconnectivity amongst all living beings, which may only be lived and not explained, and needs the absence of the self to occur. The path for reaching this state is then, described into eight limbs (Ashtanga Yoga), all of each pointed to the same aim, samadhi (the permanent state of utmost multifactorial health of the living being). Although not necessarily sequential, practices tend to be conducted in such a way in the Hatha Yogic approach (yogic approach which developed many techniques to help practitioners achieve what Sri Patanjali prescribes; it describes many practices such as postures, breathing exercises and meditations); thus, yamas (restrictions of behavior) and niyamas (observances of behavior) are learned and applied to asanas (postures); a comfortable and steady posture which means no violence nor self deceive or attachment for the practitioner is necessary and prepares for the good practice of pranayama (breathing practices); when practiced respecting and keeping the previous components applied to it, within the orientations of the most experienced and well-intentioned master, the correct practice of pranayama brings the sadhaka (practitioner) and his/her mind to the state of pratyahara (abstraction of the senses), in which the most tenacious and richer inner experiences of yoga begin; away from external interferences, untouched by sensorial inputs, the practice of dharana (concentration) is made possible, and one focuses his/her mind in a given object (what we have called “anchor”). In the state/practice of dharana, the self or individual existence/experience still remains – there is the observer and the object of observation; mind and attention come and go, as the waves in the shore; although fluctuations of the mind, these waves of thoughts and/or perceptions and memories are much subtler than those experienced in ordinary states of mind. The deepening of possible states of consciousness goes further, and Patanjali shows that the next possible achievable state is called dhyana (meditation), before the utmost one named samadhi. In a dhyana state there is supposed to be a fusion of the observer and the object observed; a state of momentary disintegration of individual identity in which one observes the absence of thoughts, since thinking is a function of the identified mind. Such a state and experience may not be practiced; conditions for its manifestation must be built, and it is what Sri Patanjali does in his Yogasutras. Thus, the mere translation of dhyana with the Western term meditation is a reduction of its meaning and reach and makes it possible for people who have experienced and practiced dharana to misunderstand one for the other. The dictionary definition of “to meditate” is “to think about a specific subject”, what may resemble Patañjali’s definition of dharana, but not, by far, dhyana. Then, some of the so-called meditation techniques which are practiced worldwide are in fact concentration exercises, which may help the practitioner to achieve the state of dhyana. In a modern approach and written in a scientific language, some authors (Cardoso et al., 2004) have contributed with an elegant definition of the called “meditation techniques”, or the means to create a meditative experience, in order to try and contribute to a better methodological control of variables, so results might be better compared and studies reproduced and replicated by others. According to the authors, it is an ancient technique used to induce a modified state of mind, consciousness and physical relaxation, and should follow an operational guideline as such: its technique should be clearly specified; there should be an emphasis on muscle relaxation during the process; a “logic relaxation” should occur; it should be a self-induced state; there should be a self-focusing object termed as “anchor”. The 62

 Neurobiology of Meditation

term “anchor” seems to be an accurate metaphor, since it refers to a device used by boats and its crew to avoid boats to get drifted by the ties – the same being applied to the mind arrested by perceptions and/or thoughts. Thus, the anchor is the object in which one tries to focus his/her attention during the concentration process that is called meditation. Accepting that definition may allow one to conclude (correctly) that asanas are a type of meditation in which the anchor is the stretched area of the body; that pranayama is also a sort of meditation in which the anchor is the breath; that in general, yoga and meditation are the same, but with anchors that vary from the denser and external to the subtler and internal. A further analysis of the meditation techniques available at present allows one to make some kind of needed differentiation. The same authors state that there may be active, passive and mixed meditation techniques (Cardoso et al., 2004). Passive or perceptive/concentrative are those in which the practitioner tries not to get involved with the anchor, with thoughts that use to appear and get to a relaxation of logic state – examples are Za Zen, Transcendental Meditation, and Vipasana. Active techniques may be exemplified by Osho’s meditations and Kinhin (walking meditation) from Budhism, or even the practice of Tai Chi Chuan. Here it is important to stress that according to the rationale presented of Patañjali’s dhyana, we consider all those “meditation techniques”, sorts of dharana, or concentration, what in no way diminish their importance, since they pave the way to the dhyana experience, which is lived and not practiced, since there is no logic involved, and also no individual present/active. This very exquisite state of mind brings about some other very specific physiological responses, which will be now addressed.

PHYSIOLOGICAL RESPONSES TO MEDITATION In the beginning of last century, much attention was driven to the fight or flight response, described by Walter Cannon (Cannon, 1932), in which the animal organism adapts to a treat and one chooses weather to confront or to avoid the imminent conflict, at many specified costs, such as increased: sympathetic drive; oxygen consumption; heart rate; blood pressure; ventilation; cortisol secretion; muscular tension and so on. In the last quarter of that century however, great attention was given to the opposite possible response. Wallace, Benson and Wilson described the “Relaxation Response” in 1971 - a wakeful hypometabolic state possible to be achieved from the practice of meditation (Wallace et al., 1971). This response elicits physiological adaptations opposite to those of the fight or flight response, such as reduction of heart rate; blood pressure; oxygen consumption; ventilation and muscle relaxation. The possible applications of such a broad effect intervention as meditation are innumerous, and fascinating. Since de description of the relaxation response, and in this century, from the event of mindfulness as a method (Kabat-Zinn, 2003). Many researchers have proven meditation to have positive effects on mental health, such as those found in an African-American sample with coronary heart disease which showed reductions in depressive and anxiety symptoms, stress management, reduced cardiovascular mortality, myocardial infarction and stroke (Schneider et al., 2012), high blood pressure control as a non-pharmacological treatment (Schneider, 2005), reduction of cortisol levels (Tang et al., 2007), reduction in neuroticism and increase perceived psychological stress control associated to greater telomerase expression and activity (Hoge et al., 2013), besides improving sleep quality and increasing melatonin secretion (Tooley et al., 2000). However, since the focus of this chapter is on neurobiology, we will not address deeply those studies in order to have a closer look into nervous adaptations to meditation and its possible therapeutic uses as an integrative practice.

63

 Neurobiology of Meditation

NEUROBIOLOGICAL EFFECTS OF MEDITATION According to Koike and Cardoso (2014), meditation has measurable effects both in the Central Nervous System (CNS) as well as in the Autonomic Nervous System (ANS), opinion which is corroborated by a very elegant paper by Gard and colleagues (2014) in which the authors describe the neuronal mechanisms of yoga (seen here as a synonymous to meditation), stating that it has both a top-down component, present and motivated by the conscious and self-induced detachment from thought and memories, as well as a Bottom-Up component present on the integration process of observing body parts and/or breath or any given anchor. Similar statement is made by Esch (2014), when he affirms that the mechanisms of meditation are divided into four areas: attention regulation, emotion regulation, body awareness and self-perception. In terms of the CNS, effects of meditation may be further divided into structural (in which we will include tissue variables – gray and white matter, vessels, blood flow) and functional (brain activation, connectivity and neurotransmitters).

FUNCTIONAL INFLUENCES OF MEDITATION Since the end of last century, the influences of Meditation on brain function is being studied (Davidson, 1998), and he states that the shift of cerebral activation to a frontal left predominance may indicate a state of greater calm and happiness. In fact, there is overall reduction of brain activation during Meditation (Lazar et al., 2000), with specific shifts of activity. Davidson and colleagues (2003) have found in another study an increased activation of the left frontal area of the cortex, which was also associated to antibody titer after a flu injection at the end of a mindfulness retreat when compared to a passive control group; such findings are corroborated by other authors (Brefczynski-Lewis et al., 2007), which found increased activation of the prefrontal cortex and of the limbic system in long term meditators. Besides, other authors also found increased activation of areas related to positive reward and motivation, attention and emotional responses to bodily stimuli (Critchley et al., 2001), while others found increased activity of areas related to controlling and evaluating emotions, as well as relating them to memory areas such as the hippocampus both during and after the practice of Meditation (Esch et al., 2004). Furthermore, increased activation of the anterior cingulated gyrus and of the corpus callosum (Tang et al., 2010), also found during sort-term meditation, which may indicate a facilitated communication between brain hemispheres, and also indicate at least in part the mechanisms trough which meditation helps controlling blood pressure and heart rate, since the anterior cingulate plays an important role in modulating autonomic functions, and on top of it, it helps decision-making and attention allocation, also under the influence of meditation. As related to stress management and brain activity, meditation and increased reported mindfulness are proven to reduce the activation of the amygdala and increase the activation of the prefrontal cortex during emotions recollection/naming (Creswell et al., 2007); in addition, Hölzel et al. (2009) has also stated that meditation helps one coping better with stress. In a recent and very elegant study (Kozasa et al., 2018) a group of meditators was immersed in a 7-day meditation retreat in the Buddhist traditional Zen meditation, undergoing structural magnetic resonance imaging prior and right after it. Nineteen experienced meditators and 14 meditation-naïve were compared. Subjects performed an attention task while being examined. There was no difference between meditators and non-meditators for the number of the correct answers and response-time during the Stroop Word Color Test performed 64

 Neurobiology of Meditation

before and after the training, however, many brain areas were less activated in the meditators already in the pre-training condition: anterior cingulate, ventromedial prefrontal cortex/anterior cingulate, caudate/ putamen/pallidum/temporal lobe (center), insula/putamen/temporal lobe (right). Authors state that the meditators training in attention during the practice of meditation might have been responsible for such a result even in the contrast incongruent/neutral condition of the Stroop test. This difference disappeared after the meditation training, since activation decreased in the control group and increased in the meditators one, indicating effects of meditation on neural function and also plasticity. Not only stress, but also the aging process has a strong effect in reducing brain functional connectivity. In fact, some authors who study the aging processes have shown functional connectivity losses, which may include decreased resting state connectivity of the hippocampus (Salami et al., 2014), connectivity of the default mode network (DMN) (Mowinckel et al., 2012); working memory network (DeCarli et al., 2012) and salience network (He et al., 2014). Besides that, antero-posterior connectivity also suffers the effects of aging, which is also associated to senescence cognitive declines, (Vidal-Piñeiro et al., 2014; Toussaint et al., 2014) which also found smaller deactivation of the DMN when facing attentional stimuli (superior frontal gyrus, posterior cingulated gyrus are some examples of areas affected) (Damoiseaux et al., 2007). Other effects of aging are decreased amplitude and increased latency in P300 component of evoked related potential, indicating a decline in cognition related to smaller neuronal availability when facing memory and selective attention tasks (Dinteren et al., 2014). In an opposite direction to these declining events and facts present in the aging process, meditation has been proven to also restore partially the decreased brain functional connectivity. In fact, Gard and others (2014) verified increased fluid intelligence and greater functional connectivity resilience during steady state functional magnetic resonance imaging acquisition in elderly which practiced yoga or meditation. In another very elegant paper, the same group of investigators (Gardet et al., 2015), the same authors found increased degreecentrality of the caudate, which indicates greater resting state functional connectivity of an area related to motor and attentional processes - in a sample of elderly practitioners of yoga/meditation; increases were significant between the caudate and the para hippocampal gyrus and with the left inferior temporal gyrus, which indicate effects of meditation in motor-related activations; inhibitory control of action and reward system; and associative and procedural learning. Regarding the DMN (network which remains active during self-referenced thoughts and planning, memory retrieval and rumination – negative aspects associated to depression), meditators have decreased DMN resting activation, during meditation, and also an increased and/or preserved connection between anteroposterior structures of this network (posterior cingulate and dorsolateral prefrontal cortex), what may indicate cognitive preservation (Brewer et al., 2011). Meditation has its effects on the neuro-endocrine system as well, since it has been shown to be associated to reduced secretion of thyrotropin releasing hormone, growth hormone releasing factor and corticotrophin releasing hormone (Bevan et al., 1979), while other authors have seen increased secretion of endorphin (Infante et al., 1998), serotonin (Walton et al., 1995) and melatonin (Tooley et al., 2000) as a result of the practice of meditation. All those put together may point to a neuro-endocrine regulating effect of meditation. In fact, in a recent systematic review (Moraes et al., 2018), authors have stated that the effects of meditation and meditation-like techniques on the psycho-neuro-immune system are so pronounced that they may be used as interventions to help attenuate disease progression and/or sideeffects of drug treatment. In the 42 studies included in the review (published between 2005 and 2015), the authors found association of the practice of meditation and yoga to decreased levels of cortisol, epi-

65

 Neurobiology of Meditation

nephrine and norepinephrine; other positive associations were also found between those practices and sleep disorder, depression, anxiety, reductions of inflammatory processes, and symptoms of fibromyalgia. A preservation of the species related part of the nervous system is the ANS. Even in a sub-cortical organization, when a person feels challenged by any given situation, a non-conscious ANS activation occurs, and a very intricate chain of neurochemical events take place. The more self-identified, the more present the roots of all sufferings are (raga and dvesa – attachment to life and fear of death) as stated by Sri Patañjali. These seeds of suffering are, obviously possible due to an a priori fundamental misunderstanding, maya, or illusion – which makes one to believe he/she is different and separated from others and has rights and needs superior to those of others, giving rise to conflicts and sufferings. It is reasonable to state that the more attached to life and possessions, the more threatened one may feel when a given challenge appears and thus, a sympathetic over-reaction may take place, increasing the neurophysiological consequences of the fight or flight response. Since meditation in a process through which the sadhaka takes him/herself towards the dissolution of the individuation process, it may have very positive effects in regulating the activities of the ANS. In fact, some authors have reported increased heart rate variability of yoga elderly practitioners who practiced bhastrika pranayama for 4 months, shifting the sympathovagal balance towards a parasympathetic predominance (Santaella et al., 2011); corroborating those findings, Peng et al. (1999) also found a significant decrease of the sympathovagal balance with increased heart rate variability during meditation, which also indicate a predominance of the parasympathetic branch of the heart rate autonomic modulation. Some others have also found reduction of the sympathetic modulation, associated with reduction of cortisol (Jevning et al., 1978) and epinephrine and norepinephrine levels (Infante et al., 2001); all these changes indicate an increase in the parasympathetic activation of the ANS. Taken together the presented facts about the influences and effects of meditation on brain functionality, connectivity and neuroendocrine function, it is possible to infer that meditation may have a real and positive influence on cognition wetter in young or elderly practitioners. Important to point out is the influence it may have in decreasing DMN activation during resting state, whilst preserving its functionality, since the over-activity of such a network indicates an increased value attribution to the self/ personal reference, which is diminished during the higher and deeper states of the mind experienced during meditation.

STRUCTURAL INFLUENCES OF MEDITATION Many studies have now come to strengthen the body of evidence that meditation does have a significant effect not only on brain functionality, but on its structure as well. Studies with magnetic resonance are available from different groups and done for quite some time. Interesting starting information was provided by Lazar and colleagues (2005), who investigated cortical thickness of 20 volunteers experienced in insight meditation and compared to matched controls. Some brain regions were significantly thicker in the Meditation group than in controls, such as prefrontal cortex and right anterior insula, brain regions associated to interoception, attention, and sensory perception. Specifically, in prefrontal cortex thickness, differences were greater in older and more experienced participants, which suggest, according to the authors, that meditation can have age-related cortical thinning preventive effect. Besides, a previous work of our laboratory done through structural brain magnetic resonance showed increased thickness of the left prefrontal lobe comprised by anterior and lateral portions of the middle and superior frontal 66

 Neurobiology of Meditation

gyri in women over 60 years of age who practiced yoga for at least 8 years when compared to a matched control yoga naïve group. Important to point out is the fact that although this study was conducted and focused on yoga, since groups were matched for age, physical activity, handedness and years of formal education, it is feasible to state that the physical aspect of the yoga practice, such as the asanas that resemble physical exercises as stretches may have had little influence on the results, and greater relevance for them must be addressed to the attentional component of yoga, present in all practices as stated in the beginning of the chapter, from yamas and niyamas to asanas and pranayamas, which, as such, put yoga and meditation yet another time as similar practices (Afonso et al., 2017). Taking into consideration that meditation might influence cortical thickness, some authors (Sato et al., 2012) used structural magnetic resonance to compare brains of experienced meditators and non-meditators with the hypothesis that it could be possible to predict by the mere analysis of brain structure weather a person was or was not a meditation practitioner without having previous information about it. Classification of regular meditators and non-meditators was done through support vector machines, which found some areas as containing important data, which were used to predict and discriminate groups with 95 percent of accuracy: right pars opercularis cortex; right precentral gyrus; left entorhinal cortex; thalamus; and right basal putamen). Another study that corroborates the statement that Meditation practice influences cortical volume and thickness is given by Hernandes and colleagues (2016) who compared 23 experienced Sahaja yoga meditation practitioners with 23 matched non-meditators (age, gender and education level), through structural magnetic resonance and found greater gray matter volume in areas spread through the whole brain, with predominance for the right hemisphere: ventromedial orbitofrontal cortex, insula, inferior temporal and parietal cortices, but also in left ventrolateral prefrontal cortex and left insula. Authors state that this effect of meditation is related to its training component of focusing and sustaining attention, interoceptive perception, compassion and self-control. Further evidence of greater gray matter volume is given by Hölzel and colleagues (2011) who showed that 8 weeks of mindfulness training led to positive effects on psychological well-being increases in regional brain gray matter density of sixteen healthy volunteers, who had never practiced meditation when compared to a control group of 17 individuals, also meditation-naïve, as a waiting list. Authors found increased gray matter density of the left hippocampus, cerebellum, temporo-parietal junction, and posterior cingulate cortex, areas associated to learning and memory, emotion regulation, perspective taking, and self-referential processing.

CONCLUSION Although all studies presented may have investigated meditation techniques, and not necessarily the state of meditation or dhyana itself, it seems reasonable to state that those techniques are effective in bringing positive effects to the brain. There is still some work to be done, especially in determining if there is a dose-response relationship, and possible similarities and/or differences amongst types of meditation techniques and their effects. Nevertheless, investigative facts reported here make it possible to state that meditation techniques, leading to the state of meditation, in which the self is not the main focus, are furthermost positive practices which might be present in the daily routines of those who wish to preserve cognition, as present in immediate and working memory, associative capacity, emotions control and regulation, attention and many other variables. No side-effects have been reported, and all practices are among the cheapest possible to encounter; many have a well-defined methodology, and beginners should

67

 Neurobiology of Meditation

choose those to start with. Following the methods tested whether by tradition or by science and in the best cases by both, the sadhaka will certainly contribute for his/her best possible state of mental health.

REFERENCES Afonso, R. F., Balardin, J. B., Lazar, S., Sato, J. R., Igarashi, N., Santaella, D. F., & Kozasa, E. H. (2017). Greater cortical thickness in elderly female yoga practitioners—A cross-sectional study. Frontiers in Aging Neuroscience, 9, 201. doi:10.3389/fnagi.2017.00201 PMID:28676757 Bevan, A. J. W., Symons, R. G., Beng, C. G., & Welby, M. L. (1979). Short-term endocrine changes in transcendental meditation. Proceedings of the Endocrine Society of Australia, 2, 56. Brefczynski-Lewis, J. A., Lutz, A., Schaefer, H. S., Levinson, D. B., & Davidson, R. J. (2007). Neural correlates of attentional expertise in long-term meditation practitioners. Proceedings of the National Academy of Sciences of the United States of America, 104(27), 11483–11488. doi:10.1073/pnas.0606552104 PMID:17596341 Brewer, J. A., Worhunsky, P. D., Gray, J. R., Tang, Y. Y., Weber, J., & Kober, H. (2011). Meditation experience is associated with differences in default mode network activity and connectivity. Proceedings of the National Academy of Sciences of the United States of America, 108(50), 20254–20259. doi:10.1073/ pnas.1112029108 PMID:22114193 Cannon, W. B. (1932). The wisdom of the body. W.W. Norton & Company. doi:10.1097/00000441193212000-00028 Cardoso, R., De Souza, E., Camano, L., & Leite, J. R. (2004). Meditation in health: An operational definition. Brain Research Protocols, 14(1), 58–60. doi:10.1016/j.brainresprot.2004.09.002 PMID:15519952 Creswell, J. D., Way, B. M., Eisenberger, N. I., & Lieberman, M. D. (2007). Neural correlates of dispositional mindfulness during affect labeling. Psychosomatic Medicine, 69(6), 560–565. doi:10.1097/ PSY.0b013e3180f6171f PMID:17634566 Critchley, H. D., Melmed, R. N., Featherstone, E., Mathias, C. J., & Dolan, R. J. (2001). Brain activity during biofeedback relaxation: A functional neuroimaging investigation. Brain, 124(5), 1003–1012. doi:10.1093/brain/124.5.1003 PMID:11335702 Damoiseaux, J. S., Beckmann, C. F., Arigita, E. S., Barkhof, F., Scheltens, P., Stam, C. J., & Rombouts, S. A. R. B. (2007). Reduced resting-state brain activity in the “default network” in normal aging. Cerebral Cortex (New York, N.Y.), 18(8), 1856–1864. doi:10.1093/cercor/bhm207 PMID:18063564 Davidson, R. J. (1998). Anterior electrophysiological asymmetries, emotion, and depression: Conceptual and methodological conundrums. Psychophysiology, 35(5), 607–614. doi:10.1017/S0048577298000134 PMID:9715104 Davidson, R. J., Kabat-Zinn, J., Schumacher, J., Rosenkranz, M., Muller, D., Santorelli, S. F., & Sheridan, J. F. (2003). Alterations in brain and immune function produced by mindfulness meditation. Psychosomatic Medicine, 65(4), 564–570. doi:10.1097/01.PSY.0000077505.67574.E3 PMID:12883106

68

 Neurobiology of Meditation

DeCarli, C., Kawas, C., Morrison, J. H., Reuter-Lorenz, P. A., Sperling, R. A., & Wright, C. B. (2012). Session II: Mechanisms of age-related cognitive change and targets for intervention: neural circuits, networks, and plasticity. Journals of Gerontology Series A: Biomedical Sciences and Medical Sciences, 67(7), 747–753. doi:10.1093/gerona/gls111 PMID:22570135 Esch, T. (2014). The neurobiology of meditation and mindfulness. In S. Schmidt & H. Walach (Eds.), Meditation–neuroscientific approaches and philosophical implications (pp. 153–173). Springer. doi:10.1007/978-3-319-01634-4_9 Esch, T., Guarna, M., Bianchi, E., & Stefano, G. B. (2004). Meditation and limbic processes. Biofeedback, 32, 22–32. Gard, T., Noggle, J. J., Park, C. L., Vago, D. R., & Wilson, A. (2014). Potential self-regulatory mechanisms of yoga for psychological health. Frontiers in Human Neuroscience, 8, 770. doi:10.3389/ fnhum.2014.00770 PMID:25368562 Gard, T., Taquet, M., Dixit, R., Holzel, B. K., de Montjoye, Y. A., Brach, N., & Lazar, S. W. (2014). Fluid intelligence and brain functional organization in aging yoga and meditation practitioners. Frontiers in Aging Neuroscience, 6, 76. doi:10.3389/fnagi.2014.00076 PMID:24795629 Gard, T., Taquet, M., Dixit, R., Holzel, B. K., Dickerson, B. C., & Lazar, S. W. (2015). Greater widespread functional connectivity of the caudate in older adults who practice kripalu yoga and vipassana meditation than in controls. Frontiers in Human Neuroscience, 9, 137. doi:10.3389/fnhum.2015.00137 PMID:25852521 He, X., Qin, W., Liu, Y., Zhang, X., Duan, Y., Song, J., & Yu, C. (2014). Abnormal salience network in normal aging and in amnestic mild cognitive impairment and Alzheimer’s disease. Human Brain Mapping, 35(7), 3446–3464. doi:10.1002/hbm.22414 PMID:24222384 Hernandez, S. E., Suero, J., Barros, A., Gonzalez-Mora, J. L., & Rubia, K. (2016). Increased grey matter associated with long-term sahaja yoga meditation: A voxel-based morphometry study. PLoS One, 11(3), e0150757. doi:10.1371/journal.pone.0150757 PMID:26938433 Hoge, E. A., Chen, M. M., Orr, E., Metcalf, C. A., Fischer, L. E., Pollack, M. H., & Simon, N. M. (2013). Loving-Kindness Meditation practice associated with longer telomeres in women. Brain, Behavior, and Immunity, 32, 159–163. doi:10.1016/j.bbi.2013.04.005 PMID:23602876 Holzel, B. K., Carmody, J., Evans, K. C., Hoge, E. A., Dusek, J. A., Morgan, L., & Lazar, S. W. (2009). Stress reduction correlates with structural changes in the amygdala. Social Cognitive and Affective Neuroscience, 5(1), 11–17. doi:10.1093can/nsp034 PMID:19776221 Holzel, B. K., Carmody, J., Vangel, M., Congleton, C., Yerramsetti, S. M., Gard, T., & Lazar, S. W. (2011). Mindfulness practice leads to increases in regional brain gray matter density. Psychiatry Research: Neuroimaging, 191(1), 36–43. doi:10.1016/j.pscychresns.2010.08.006 PMID:21071182 Infante, J. R., Peran, F., Martinez, M., Roldan, A., Poyatos, R., Ruiz, C., & Garrido, F. (1998). ACTH and β-endorphin in transcendental meditation. Physiology & Behavior, 64(3), 311–315. doi:10.1016/ S0031-9384(98)00071-7 PMID:9748098

69

 Neurobiology of Meditation

Infante, J. R., Torres-Avisbal, M., Pinel, P., Vallejo, J. A., Peran, F., Gonzalez, F., & Roldan, A. (2001). Catecholamine levels in practitioners of the transcendental meditation technique. Physiology & Behavior, 72(1-2), 141–146. doi:10.1016/S0031-9384(00)00386-3 PMID:11239991 Jevning, R., Wilson, A. F., & Davidson, J. M. (1978). Adrenocortical activity during meditation. Hormones and Behavior, 10(1), 54–60. doi:10.1016/0018-506X(78)90024-7 PMID:350747 Kabat-Zinn, J. (2003). Mindfulness‐based interventions in context: Past, present, and future. Clinical Psychology: Science and Practice, 10(2), 144–156. doi:10.1093/clipsy.bpg016 Koike, M. K., & Cardoso, R. (2014). Meditation can produce beneficial effects to prevent cardiovascular disease. Hormone Molecular Biology and Clinical Investigation, 18(3), 137–143. doi:10.1515/ hmbci-2013-0056 PMID:25390009 Kozasa, E. H., Balardin, J. B., Sato, J., Chaim, K., Lacerda, S., Radvany, J., & Amaro, E. Jr. (2018). Effects of a seven-day meditation retreat on the brain function of meditators and non-meditators during an attention task. Frontiers in Human Neuroscience, 12, 222. doi:10.3389/fnhum.2018.00222 PMID:29942255 Lazar, S. W., Bush, G., Gollub, R. L., Fricchione, G. L., Khalsa, G., & Benson, H. (2000). Functional brain mapping of the relaxation response and meditation. Neuroreport, 11(7), 1581–1585. doi:10.1097/00001756200005150-00042 PMID:10841380 Lazar, S. W., Kerr, C. E., Wasserman, R. H., Gray, J. R., Greve, D. N., Treadway, M. T., & Rauch, S. L. (2005). Meditation experience is associated with increased cortical thickness. Neuroreport, 16(17), 1893–1897. doi:10.1097/01.wnr.0000186598.66243.19 PMID:16272874 Moraes, L. J., Miranda, M. B., Loures, L. F., Mainieri, A. G., & Marmora, C. H. C. (2018). A systematic review of psychoneuroimmunology-based interventions. Psychology Health and Medicine, 23(6), 635–652. doi:10.1080/13548506.2017.1417607 PMID:29262731 Mowinckel, A. M., Espeseth, T., & Westlye, L. T. (2012). Network-specific effects of age and in-scanner subject motion: A resting-state fMRI study of 238 healthy adults. NeuroImage, 63(3), 1364–1373. doi:10.1016/j.neuroimage.2012.08.004 PMID:22992492 Peng, C. K., Mietus, J. E., Liu, Y., Khalsa, G., Douglas, P. S., Benson, H., & Goldberger, A. L. (1999). Exaggerated heart rate oscillations during two meditation techniques. International Journal of Cardiology, 70(2), 101–107. doi:10.1016/S0167-5273(99)00066-2 PMID:10454297 Salami, A., Pudas, S., & Nyberg, L. (2014). Elevated hippocampal resting-state connectivity underlies deficient neurocognitive function in aging. Proceedings of the National Academy of Sciences of the United States of America, 111(49), 17654–17659. doi:10.1073/pnas.1410233111 PMID:25422457 Santaella, D. F., Devesa, C. R., Rojo, M. R., Amato, M. B., Drager, L. F., Casali, K. R., & LorenziFilho, G. (2011). Yoga respiratory training improves respiratory function and cardiac sympathovagal balance in elderly subjects: A randomised controlled trial. BMJ Open, 1(1), e000085. doi:10.1136/ bmjopen-2011-000085 PMID:22021757

70

 Neurobiology of Meditation

Sato, J. R., Kozasa, E. H., Russell, T. A., Radvany, J., Mello, L. E., Lacerda, S. S., & Amaro, E. Jr. (2012). Brain imaging analysis can identify participants under regular mental training. PLoS One, 7(7), e39832. doi:10.1371/journal.pone.0039832 PMID:22802944 Schneider, R. H., Alexander, C. N., Staggers, F., Rainforth, M., Salerno, J. W., Hartz, A., & Nidich, S. I. (2005). Long-term effects of stress reduction on mortality in persons 3 55 years of age with systemic hypertension. The American Journal of Cardiology, 95(9), 1060–1064. doi:10.1016/j.amjcard.2004.12.058 PMID:15842971 Schneider, R. H., Grim, C. E., Rainforth, M. V., Kotchen, T., Nidich, S. I., Gaylord-King, C., & Alexander, C. N. (2012). Stress reduction in the secondary prevention of cardiovascular disease: Randomized, controlled trial of transcendental meditation and health education in Blacks. Circulation: Cardiovascular Quality and Outcomes, 5(6), 750–758. doi:10.1161/CIRCOUTCOMES.112.967406 PMID:23149426 Taimni, I. K. (2014). The science of yoga: Commentary of the patanjali yogasutras. Theosophical Ed. Tang, Y. Y., Lu, Q., Geng, X., Stein, E. A., Yang, Y., & Posner, M. I. (2010). Short-term meditation induces white matter changes in the anterior cingulate. Proceedings of the National Academy of Sciences of the United States of America, 107(35), 15649–15652. doi:10.1073/pnas.1011043107 PMID:20713717 Tang, Y. Y., Ma, Y., Wang, J., Fan, Y., Feng, S., Lu, Q., & Posner, M. I. (2007). Short-term meditation training improves attention and self-regulation. Proceedings of the National Academy of Sciences of the United States of America, 104(43), 17152–17156. doi:10.1073/pnas.0707678104 PMID:17940025 Tooley, G. A., Armstrong, S. M., Norman, T. R., & Sali, A. (2000). Acute increases in night-time plasma melatonin levels following a period of meditation. Biological Psychology, 53(1), 69–78. doi:10.1016/ S0301-0511(00)00035-1 PMID:10876066 Toussaint, P. J., Maiz, S., Coynel, D., Doyon, J., Messe, A., de Souza, L. C., & Benali, H. (2014). Characteristics of the default mode functional connectivity in normal ageing and Alzheimer’s disease using resting state fMRI with a combined approach of entropy-based and graph theoretical measurements. NeuroImage, 101, 778–786. doi:10.1016/j.neuroimage.2014.08.003 PMID:25111470 Van Dinteren, R., Arns, M., Jongsma, M. L., & Kessels, R. P. (2014). P300 development across the lifespan: A systematic review and meta-analysis. PLoS One, 9(2), e87347. doi:10.1371/journal.pone.0087347 PMID:24551055 Vidal-Piñeiro, D., Valls-Pedret, C., Fernández-Cabello, S., Arenaza-Urquijo, E. M., Sala-Llonch, R., Solana, E., & Bartres-Faz, D. (2014). Decreased default mode network connectivity correlates with ageassociated structural and cognitive changes. Frontiers in Aging Neuroscience, 6, 256. PMID:25309433 Wallace, R. K., Benson, H., & Wilson, A. F. (1971). A wakeful hypometabolic physiologic state. American Journal of Physiology-Legacy Content, 221(3), 795–799. doi:10.1152/ajplegacy.1971.221.3.795 PMID:5570336 Walton, K. G., Pugh, N. D., Gelderloos, P., & Macrae, P. (1995). Stress reduction and preventing hypertension: Preliminary support for a psychoneuroendocrine mechanism. Journal of Alternative and Complementary Medicine (New York, N.Y.), 1(3), 263–283. doi:10.1089/acm.1995.1.263 PMID:9395623

71

72

Chapter 5

The Neurophilosophy of Meditation Vinod D. Deshmukh University of Florida, USA

ABSTRACT Meditation is the art and science of optimal self-integration. We are naturally nested in Nature. Meditation is being at-peace with oneself and at-home in the world. It is the art of self-observation, learning, and developing presence of mind and selfless love. It is a unique mode of being present in the moment. It is a natural state of health and happiness. We should let go of the vicious cycle of negative feelings and discover the virtuous cycle of positive feelings like joy, cheerfulness, equanimity, friendliness, compassion, and love. This is possible through meditative practices like yoga, mindfulness, flow-state, and nature-experience. With advancement in neurobiology, we can understand how these meditative skills are learned, developed, evolved, and mastered. These meditative skills and lifestyle are the key to positive psychology and mental health. These secular meditative practices are being recommended in our education and health care systems.

EXISTENTIAL UNITY, NONDUALISM AND MENTAL EQUANIMITY “When, to one who knows (oneness), all beings have, verily, become one with his own self, then what delusion and what sorrow can be to him, who has seen the Oneness”, translated by Radhakrishnan (1953; Isha Upanishad (800 B.C.). “What delusion, what sorrow is there for the wise person who sees the Unity of Existence and perceives all beings as his own self”, translated by Sharvananda (1958). “To the illumined soul, who sees everything as a manifestation of his own self, how can there be delusion or grief since he sees only oneness?”, translated by Krishnamurthy (2019). Nondualism means not two or one whole reality, undivided without a second. Nondualism primarily refers to a mature state of consciousness, in which the dichotomy of I-Other dissipates, disappears, or it is transcended, and the spontaneous nondual awareness is described as the selfless or centerless presence without dichotomy. DOI: 10.4018/978-1-7998-3254-6.ch005

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 The Neurophilosophy of Meditation

“Being steadfast in yoga, O Dhanajaya, perform actions, abandoning attachment, remaining unconcerned as regards to success or failure. This evenness of mind is known as yoga, translated by Swarupananda (1909; Bhagavad-Gita 2:48). “Focused in Yoga, do thy work, O Winner of wealth (Arjun), abandoning attachment, with an even mind in the success or failure, for the evenness of mind is called Yoga, translated by Radhakrishnan (1948; Bhagavad-Gita 2:48). These three terms point to the same truth of the unitive state of essential nondual awareness and the existential holistic being.

VEDIC-UPANISHADIC MEDITATION The non-REM wake state of Turiya or Chaitanyam, is the fourth state of consciousness-as-such which is non-narrative, non-cognitive, non-conative awareness. It is nondual, non-plural, spontaneous, effortless, selfless, egoless, primal, timeless, eternal, sentient arousal-aware-being. It is the sacred, divine perfection. There is nothing beyond it. It is the Ground of all, the Source of all. WE are THAT! Let all REM dreams, wandering thoughts, mental travel, mind-agitations, ruminations of the past, and simulations of the future, dissipate freely, completely, and naturally into the immeasurable ocean or the sky-like space of Life’s self-experience and existence. It is the Consciousness-as-such, Life-as-such. It is the spontaneous, still, silent, serene, supreme energy-awareness-being. It is the ultimate existential feeling or qualia of existential presence” (Deshmukh, 2020). Meditation according to the yoga tradition has an exceedingly long history. In fact, the original lineage goes back to the Vedic-Upanishadic period. Most of the Vedic-Upanishadic literature is devoted to spiritual growth, Atmonnati, and actualization of one’s essential being with clear self-understanding. This leads to complete calmness, natural peace of mind and a well-composed being in reality. Vedanta described such a state as sat-chit-anand or blissful conscious being. What one personally experiences from moment to moment depends on one’s self-perspective or the mindset. One can look at oneself as an isolated living being, a living organism in the environmental niche, a passionate human being, a cognitive intellectual individual, or a holistic human being embedded in its natural surroundings. The extraordinary self-perspective of this existential being is possible only after a profound self-understanding and self-realization. This wise stage in human life has been described in the Upanishads and BhagavadGita as the supreme self-integration, unity and equanimity called Turiya, Sthita-prajña, Atma-prasād, and Brahmi-sthiti (Deshmukh, 2011). Turiya, the fourth state of consciousness as described in Mandukya Upanishad, is an extraordinarily inspired state of calm, alert, and intuitive consciousness with self-integration. It can be understood within the framework of modern neuroscience. The author proposed that Turiya is a unique conscious state of non-rapid eye movement wakefulness (NREM-W) characterized by a) an optimal attentiveness with freedom from day dreaming and intrusive thoughts, b) behavioral quiescence with no internal speech or restless eye movements, c) heightened perceptual clarity from moment to moment with perceptualconceptual excellence and d) an enhanced sense of well-being, curiosity, serenity, compassion, and creativity. Many exceptionally gifted people throughout human history seem to have verified the truth of such a unique state. Just as we stand on the shoulders of past intellectual giants to understand and discover what is new in science, so also, we may be guided by the genius and experiences of the ancient seers and philosophers to understand and discover the neuroscience of self-consciousness (Deshmukh, 2004). The most ancient description of meditation to achieve such self-realization and self-freedom is in the Mandukya Upanishad. It has only 12 mantras to explain this complex subject and the self-integrative 73

 The Neurophilosophy of Meditation

process. Two of those mantras, the seventh and the twelfth, are the most meaningful. Here is a translation of those two mantras by Chinmayānanda (1983). “It is not that which is conscious of internal subjective world, nor that which is conscious of the external world, nor that which is conscious of both, nor that which is a mass of consciousness, nor that which is simple consciousness, nor is it unconsciousness; It is unseen by any sense organ, beyond empirical dealings (interactions) incomprehensible by the mind, uninferable, unthinkable, indescribable; essentially it is realized by the peaceful, auspicious, and the nondual Self alone, negating all experiential phenomena (Neti, Neti: Not this, Not this). This is what is considered as Turiya, the fourth state of consciousness. This is also Atman (the essential Being); it is to be realized”. That which has no parts (which is whole), the soundless (silent), the incomprehensible, beyond all the senses, the cessation of all phenomena, all-blissful and nondual AUM, is the fourth state of consciousness, and verily it is the same as Atman. He who knows this, merges his self (ego, sense of agency, location, and ownership) in the supreme Self – the individual (being) in the Total (the Universal Being). Turiya or Chaitanyam, the fourth state of consciousness, which is non-REM wakefulness, or bare awareness was formally described in an article (Deshmukh, 2004). A general theory of relaxation was proposed by Kokoszka (1992). He also proposed four states of consciousness on the basis of basic rest activity cycle (BRAC), REM and non-REM sleep-wake cycle, and active and passive modes of wake behavior, and the ordinary versus extraordinary or altered meditative states. He called the fourth state of consciousness a differentiated waking state of consciousness, which seems to be analogous to the state of Turiya from Vedic-Upanishadic literature. He also mentioned ergo-stasis as a dynamic equilibrium of energy (sentient arousal), which is analogous to the metabolic homeostasis, and behavioral allostasis.

INTEGRATIVE AND INSIGHTFUL MEDITATIVE PRESENCE “The observed ever changes. The observable observer also changes. The integrative presence is changefree. That is our true nature. Discover it, (meditative) presence is the instantaneous and simultaneous consciousness. Presence is current consciousness. There is no constructed self-image in presence. Such cognitive construction is usually based on autobiographical memories. How can past be present? Self-free presence is blissful. It cannot be precisely defined as it has no fines or limits. It implies limitless silence, freedom, energy, intelligence, and love. It is undivided awareness. It cannot be stored or accumulated. It is a unique mode of being and a great way of living. Presence is the ground state of attention. It is undirected attentiveness. It can be directed anywhere anytime, if intended. Like superfluid, it is the most adaptive consciousness” (Deshmukh, 1990). What is integrative presence? It is a spontaneous and effortless feeling of being mindfully present in the present moment. Presence or present mind has the limitless potential. Our mental activity is an expression of that immeasurable present mind. It is not a thought or memory. It is the feeling of being the whole, as one is, at this moment. I had described it as the experience of no-experience. It is a holistic and primal optimal feeling of being in existence now. It is ever-fresh and new. As the saint Jñāneshwara described it, “Atma-prabha nich navi”. The spontaneously self-effulgent arousal (sentient life-energy), awareness, and being are ever-fresh and new. It serves as an inner light or guide, like the sky-lamp that illumines the dark path at night. One who abides in such a serene inner glow, realizes the truth of absolute unity and freedom (Dandekar, 1953).

74

 The Neurophilosophy of Meditation

What is insightful presence? It is the intuitive deeper understanding of the self and surroundings at the present moment. Such an insight is possible only with full conscious awareness and a feeling for the wholeness of the present reality. This is what I have called the Insightful Presence. Think of the three aspects of our mind: hindsight, insight, and foresight. Hindsight is recalling the past based on our multiple memories and reconstructing a conceptual model of reality including past self-surroundings autobiography. Foresight is predictive processing of the future. For foresight, one carries a self-constructed internal model of the surrounding world, and predictively anticipates the future and plans for the action. Based on the outcome of one’s action, one updates the internal model for the next predictive action. It is a trial-and-error learning and acting. Machine-learning in the field of artificial intelligence is based on similar predictive processing. Insight is being fully present and aware of the present self-surrounding situation. The present experiential reality is fluid, dynamic, and ever-changing; yet it appears to be the same. It is like the tip of a water fountain, where the water is continuously renewed. That is why insightful presence is ever-fresh and new. It can be full of surprises. It reduces the momentary uncertainty, which is the phenomenal unit of conscious information-processing. Meditation is self-integration and self-harmony. It is a unique spontaneous feeling of being at-peace with self and at-home in the Nature. It implies a restful, refreshed body-brain, a non-wandering, focused mind, a self-satiated self, living freely, joyously, and fearlessly, in the endlessly beautiful, benevolent, and complex natural world. It is a genuine feeling of being the life-itself with a sense of awe, oneness, or unity with the natural existence. It is an affective holistic self-aware being, not just a narrative and cognitive self, interacting with the world. It is an adaptive synchrony, synergy and harmony with Nature. It is nondual integrated awareness. It is an ecosystems perspective with affective awareness without any existential conflict, anxiety, or fear, including the fear of personal death. It is true freedom from oneself; when one is free, it is natural to be happy. It is living a life with curiosity, creativity, and compassion for all. It loves the timeless presence. The final state of self-freedom or Kaivalya is ineffable; it is a matter of direct personal experience.

MEDITATION AND THE ORIGINAL GROUND OF EXISTENTIAL EXPERIENCE Meditation is returning to the ever-present ground of our being or existing. The ground of conscious and subconscious energy-being-awareness includes conscious arousal, awareness, attention, and agency. These are the four essential elements of conscious living and existing which are effulgent and pyramidally projected neural-behavioral self-organizing system. Specific neural networks synchronize with each other for all bio-psycho-social-existential processes. These foundational neural networks are 1) reticulo-limbic network for arousal and emotion, 2) precuneus-cingulate network for intrinsic self-awareness and intentionality, 3) fronto-parietal network for attentive interactions with the surrounding experiential world, and 4) temporo-parietal-hippocampal network for autobiographical memory (sense of agency, location and ownership), thinking, context and consequences. The whole experienced reality is grounded in and is dependent upon these four components of consciousness, the subconscious background activity, and the synchronized neural networks operating in the real world. We are always grounded in reality, which is both imminent in the experience now, and transcendent as the whole beyond. The individuals are always grounded in the experiential, and existential world. Jiva or the sense of self is ever grounded in the Atman-Brahman, which is the ultimate ground of all, being infinite and ineffable. The intrinsic feeling of holistic consciousness is nondual and timeless. It is the 75

 The Neurophilosophy of Meditation

wholeness of existence, and not a limited part of our narrow personal experience. One feels grounded in the wholeness of the present moment. One feels that I am the whole person (Poornatma). The natural wholeness includes all things and all living beings. In spontaneous stillness, silence, and serenity, one returns to a feeling of oneness, or the existential unity. Krishnamurti (1978, 1979, 1983) on meditation: “This transformation is not the “me” becoming the greater “me” but the transformation of the content of consciousness; consciousness is its content. The consciousness of the world is your consciousness; you are the world and the world is you. Meditation is the complete transformation of thought and its activities. Harmony is not the fruit of thought; it comes with the perception of the whole”. “Meditation is the emptying of the content of consciousness”. “The death that meditation brings about is the immortality of the new”. “Meditation is the freedom from all measure”. “To find out whether the brain can operate completely, wholly, all the senses must operate at the same moment, at the same level, with the same intensity. Then you will find with the total movement of all the senses, there is no centre. It is only when partial sensory responses take place that there is a centre. That is the beginning of the ego, the ‘me’, the self”. “Where there is silence there is space – not from one point to another point as we usually think of it. Where there is silence there is no point but only silence. And that silence has that extraordinary energy of the universe (apogee). There is an origin, an original ground, from which all things arise, and that original ground is not the word. The word is never the thing. And meditation is to come upon that ground which is the origin of all things and which is free from all time. This is the way of meditation. And blessed is he who finds it”. Why is meditative freedom beyond all measures? The answer is that, the very process of human observation and measurement is a deliberate, and an intentional act with observer-observed duality and memory-based cognition, conceptualization, and thinking. True freedom is a spontaneous feeling of being fully alive, aware, and one with all that is. Such a self-integrated feeling of oneness has been described as the infinite space-like (Akashavat), oceanic (Samudravat), non dual awareness (Advaitabhava) and consciousness as such (Chaitanyam). It is the direct meditative feeling expressed in Advaita Vedanta, Buddhism and other meditative disciplines. Physical science depends on the measurement of an observed phenomenal event, whereas, meditation is about one’s present mode of being – existing, being synchronized with the natural reality with a spontaneous feeling of self-integration and ego-transcendence.

SELF-INTEGRATION, NONDUAL AWARENESS AND SELF-FREEDOM Historically, the state of experiential unity, nondual awareness, and self-freedom or Kaivalya has been well described in the Vedic-Upanishadic and yogic literature. It is also described in the buddhist literature. The supreme existential unity and multiplicity was beautifully described in the Isha Upanishad (mantra 4-5), “the self (Atman-Brahman, Universal Spirit) is one. Unmoving, it is faster than the mind. Having preceded the mind, it is beyond the reach of the senses. Ever steady, it outstrips all that run. By its very presence, it enables the cosmic energy (Natural Circadian Energy, prana, maatarishwa, sutratman) to sustain the activities of (all) living beings. It (atman, being, self, spirit) moves and It moves not. It is far, and it is near. It is within all this, and it is also outside (beyond) all this (experience-existence)”. Radhakrishnan (1953) commented on this very clearly: “It is that whose activity sustains all life, on which all causes and effects depend and in which all these are here, which is called the thread which supports all the worlds (through which it runs). The whole world has the supreme self as its basis. The supreme is one essence but has two natures, an eternal immutability, and an unceasing change. It is still76

 The Neurophilosophy of Meditation

ness and movement. Immovable in itself, all things are moved from It. The unity and manifoldness are both aspects of the (same) life divine. Unity is the truth and multiplicity is its manifestation. The former is the truth, vidya, the latter is ignorance, avidya. The latter is not false except when it is viewed in itself, cut off from the eternal unity. Unity constitutes the base of multiplicity and upholds it, but multiplicity does not constitute and uphold unity”. The state of nondual self-awareness and absolute self-freedom or Kaivalya were summarized recently (Deshmukh, 2020). Advaita-bhava, non-dual awareness, or consciousness as such is finally being recognized in the neuroscience literature especially in the writings of a few neuroscientists (Josipovic, 2019). Josipovic proposed a dynamic functional neural network with its main node in the central area of the precuneus, and its main axis with another node in the dorso-lateral prefrontal cortex, as the likely neural correlate of nondual awareness and dual consciousness. He described ten self-evident properties of non-dual awareness (advaita-bhava) as follows: 1. Presence or Being—this is perhaps the most obvious of all dimensions of consciousness-as-such. It is self-evident existence. 2. Emptiness—non-reification, ineffability. This is, both, an absence of any phenomenal content other than itself, and an absence of any categorization or reification about itself. 3. Non-representational reflexive cognizance—nondual awareness that knows itself to be conscious or aware directly, without relying on conceptual and symbolic representations, and without subjectobject structuring. 4. Luminosity or radiance—nondual awareness appears as if lit from within itself by its own clear transparent light. Clarity here refers to how unobscured by the substrate this awareness is. 5. Bliss—a silent contentment of being complete in itself; no sense of anything lacking in any way. 6. Nondual—without subject-object structuring; not taking itself as either subject or object of its knowing. 7. Infinite—nothing outside of it, it is experienced as infinite in that it does not have any boundaries or edges. 8. Singularity or unity—homogenous, unity of all dimensions; not constructed or created through altering mind or body but merely recognized (abhijnaana) or realized. 9. Continuous—unchanging self-sameness; upon clearly self-recognizing, it also recognizes itself as that awareness which was always present in all experiences (Singh, 1963). 10. Self—self as aware presence, singular and continuous. This is not a self in the usual sense of constructed self. Yet, nondual awareness is not someone else’s awareness, or an object fundamentally different from oneself. Rather, it is who one is, and has always been, as the conscious presence in all of one’s experiences, an intrinsic reflexive knowing. These ten properties of nondual awareness or advaita-bhava remind me of the beautiful Sanskrit verses from Adi Shankaracharya’s Aparokshānubhuti, (1938) the direct experience of self-realization and freedom. Nondual and dual awareness are two modes of our being-in-existence. In the dual cognitive mode, we interact with the other person, an object, a mental idea, a concept, or a thought (Quesque & Brass, 2019). The dualistic response and reactive behavior are hierarchically organized at the brainstem tectum level, the subcortical limbic level, and finally at the neocortical level. 1) At the tectal level, the response can be either approach or avoid depending on whether the object or person that one is facing is inviting or threatening (Cisek, 2019). 2) At the subcortical limbic level, the ventro77

 The Neurophilosophy of Meditation

medial prefrontal cortex and the anterior cingulate cortex are involved in self-monitoring, interoception, and the approach response, whereas, the insular cortex and amygdala are involved in the exteroception, and the avoidance, aggressive, or defensive response. 3) At the neocortical level, the complex response is much slower and more deliberate. It would involve the consideration of the context of the situation, as well as the consequences of one’s actions and behavior. This would involve self-consciousness with value assessment in terms of risk: reward ratio, via anterior cingulate, prefrontal, precuneus networks and the fronto-parietal executive network. In nondual awareness, there is no interaction with the other, because there is no “other” in the unique feeling of our primordial and unitive self-aware being. Therefore, there is no sense of the behavioral center, agency, location, or ownership. It is spontaneous, intrinsic, complete, and blissful being-in-itself. Kaivalya or true self-freedom is freedom from one’s self-centered mindset and behavior. When one is free, it is natural to be happy, cheerful, curious, creative, and compassionate. Such an absolute mental freedom is possible. All of us have the potential for it. It is a matter of our self-observation, self-learning, self-regulation, self-perspective, and the essential feeling of self-sentient existence and holistic awareness.

MIND-BODY MEDICINE Mind-body medicine-practices include yoga, mindfulness, and other meditations, which can help us to develop mindful presence, serenity, empathy, and benevolence. They can also lead to a stress-free personal wellbeing, emotional resilience, prosocial behaviour, and an adaptive personality. These practices should be used more often and taught widely and regularly in our education and health care systems (Dossett et al., 2020). Telles et al. (2015) proposed the following neurobiological mechanisms for the positive effects of mind-body practices: (a) re-patterning of primary interoceptive and higher order homeostatic mechanisms; (b) improved central regulation of autonomic, psychologic, neurologic, immunologic, cardiorespiratory, and gastrointestinal functions; (c) reorganization within cortical and subcortical structures, interconnectivity adjustments among central regulatory networks, neurotransmitter changes, improved emotion regulation by higher centers, better interhemispheric balance, and enhanced cognitive function; and (d) modulation of epigenetic factors, such as growth factors or hormones, as well as extensive up and down regulation of genes. Lutz and colleague (2019) proposed a predictive coding or processing perspective on understanding the effects of meditation with an emphasis on non-action, non-engagement and disengagement from the habitual mindset and experience. They summarized that the increased capacity of both non-engagements with, and active disengagement from the automatized and predetermined courses of mental activity, seems not only to make the meditator realize the extent to which he/she is usually driven by such habitual schemes unawares, but also to endow him/her with an increased capacity to counteract their attractive power. Thus, meditation may facilitate cognitive flexibility and creativity. The impact of meditation practices is mainly on downregulating affective and cognitive habits. Schoenberg and Vago (2019) proposed a developmental-stage model of meditation classification on the basis of electrophysiological studies. They proposed a pyramid of five meditative states namely 1) relaxation practice with neuro-visceral integration, 2) concentration, attention practice, 3) insight practice, 4) non-dual awareness, and 5) unified compassion or unconditional love. From this article, it is not clear how such a classification could be developed entirely on the basis of human electrophysiology 78

 The Neurophilosophy of Meditation

including EEG, evoked potentials, and the readiness potential studies. The pyramid seems to be based more on the review of the literature on meditation than electrophysiology.

Ashtanga Yoga Meditation Patañjali’s yoga sutra is a classic text on the philosophy and practice of yoga. The philosophy of yoga describes two purposes for the practice of yoga namely a) living a healthy stress-free life (klesha-tanukarana), and b) to realize an intuitive, selfless, holistic consciousness (samādhi bhāvanā). These two goals are achieved by persistent learning and practice (abhyāsa) and a detached mental attitude toward transient phenomena (vairāgya). It described eight steps (ashta-angās). The first five steps are external called bahiranga. They consist of 1) yama (social conduct), 2) niyama (personal lifestyle), 3) āsana or steady and comfortable posture, 4) prānāyāma or a disciplined effortless breathing, 5) pratyāhāra or self-restraining of multi-modal sensory-motor inputs and outputs. The last three steps of the yoga sutra are internal, called antaranga. They consist of 6) dhāranā or focused attention-meditation, 7) dhyāna or calm and sustained attentiveness or presence, and 8) samādhi or selfless presence, and holistic, intuitive consciousness with spontaneous insight-wisdom (Ritambharā Prajñā). These three steps together are also called sanyama. In the yogic tradition, the ascent and descent of Kundalini shakti and the ways to achieve its arousal-awareness-being have been described. The passage of Kundalini energy-arousal through Kundalini chakras (centers or neural networks) has been described. There is a considerable symbolism in these descriptions resulting in some misunderstanding, inaccuracies, speculations, and mysticism (Deshmukh, 2012a). Dharana, dhyana, samadhi, and sanyama can be understood as spontaneous excellence with optimal self-regulation, effortless performance, joyous experience, and the natural self-absorptive being in the present situation. Here are a few descriptions of the four terms. Deshmukh (1990) suggested that the focusing of conscious arousal (awareness) to a point or event in the environment is focal attention. Total undivided attention to an event or object for a long period of time is vigilant meditation. In the advanced stages of vigilance, we are so completely engrossed in the present totality that we forget ourselves as separate entities: ego-transcendent meditation. These three skills together constitute perfected self-regulation and self-excellence. Kak (2016) described concentration or dharana as the steadfastness of the mind. Continued focusing of mental-effort on the (same) place is meditation. Meditation, shining forth as the object alone, is trance or samadhi. In samadhi, the individual does not interpose himself between the object and its cognition. In other words, the individual loses his sense of (a separate) self or being and he flows with the activity. In samadhi, man and nature is one. The three together constituting restraint. There is a lot of similarity between these three mental skills and the flow state as described by Csikszentmihalyi (1990). The neurocognitive mechanisms underlying the flow state were summarized by Harris and colleagues (2017). He described the flow state in attentional terms—focused concentration and task absorption—specific cognitive mechanisms have received limited interest. They proposed that an attentional explanation provides the best way to advance theoretical models and produce practical applications, as well as providing potential solutions to core issues such as how an objectively difficult task can be subjectively effortless. Patañjali’s yogasutra has been translated by many authors including (Woods, 1914; Jha, 1934; Taimni, 1961; Mukerji, 1963; Satchidananda, 1978; Deshmukh, 1990; Leggett, 1992; Kak, 2016).

79

 The Neurophilosophy of Meditation

Prana-Dhyana Yoga Prana-Dhyana Yoga is a Sanskrit word for the ancient discipline of meditation, as a means to samadhi or enlightenment. Samadhi is a self-absorptive, fully integrated state with the realization of one’s existential being in harmony with the natural reality. It is unitive, undifferentiated, reality-consciousness, and an essential being, which can only be experienced by spontaneous feeling, intuition, and self-understanding. Modern neuroscience can help us to better understand Prana-Dhyana Yoga. Deshmukh (2006) discussed topics including brain-mind-reality, consciousness, attention, emotional intelligence, sense of self, meditative mind, and meditative brain. Meditation is an art of being serene and alert in the present moment, instead of constantly struggling to change or to become someone else and being somewhere else. It is an art of adaptive management of attentional energy with total engagement (poornata, presence, mindfulness) or restful disengagement (shunyata, silence, emptiness, freshness). In both states, there is an experience of spontaneous restful unity with no sense of a separate situational self and personal time. It is a simultaneous, participatory consciousness rather than a dualistic, sequential attentive activity. There is a natural sense of wellbeing with self-understanding, spontaneous joy, serenity, freedom, and self-fulfillment. One realizes the truth of one’s harmonious being in nature and nature’s being in oneself. It is being alive and aware to the fullest when each conscious moment becomes a dynamic process of self-discovery and continuous learning about the ever-new and freshly unfolding creative reality (Deshmukh, 2012b). Herrero and colleagues (2018) described the use of direct intracranial recordings in humans, and the correlated cortical and limbic neuronal activity as measured by the intracranial electroencephalogram (iEEG) with the breathing cycle. They showed this effect to be the direct result of neuronal activity, as demonstrated by both the specificity of the finding to the cortical gray matter and the tracking of breath by the gamma-band (40–150 Hz) envelope in these structures. During volitionally paced breathing, iEEG-breath coherence increases in a fronto temporal-insular network, and during attention to breathing, there is increased coherence in the anterior cingulate, premotor, insular, and hippocampal cortices. They suggested that breathing can act as an organizing hierarchical principle for neuronal oscillations throughout the brain.

NEUROBIOLOGY OF YOGA MEDITATION Gard and colleagues (2014) proposed an excellent multi-systems model for the development of top-down and bottom-up self-regulatory skills in Patanjali’s Ashtanga-Yoga. After describing ethics, postures and breath-regulation, they define yoga meditation into four components: 1) Pratyahara or sensory withdrawal including relaxation techniques, inward-mindedness, and minimizing sensory input; 2) Dharana or concentration including single-pointed focused attention (object-based) and effortful; 3) Dhyana or meditation including unbroken flow of attention (object-based or not), open monitoring and effortless; and 4) Samadhi or self-Integration including merging of the subject and object, a transcendental consciousness and self-realization. Their schematic diagram of the systems network model of yoga for optimizing self-regulation is detailed and meaningful. They proposed that the yoga practice facilitates self-regulation via an ethically motivated monitoring and control process that involves initiation and maintenance of behavioral change as well as inhibiting undesired output by both higher level and lower level brain networks in the face of stress related to physical and emotional challenge. The high level 80

 The Neurophilosophy of Meditation

brain networks are involved in intention, motivational goal setting, attentional control, meta-awareness, response-inhibition, working memory, and cognitive reappraisal. The mechanisms from low level brain networks are involved in parasympathetic control, improved baroreceptor functioning, increased vagal tone, strengthening of the diaphragm, extinction learning, and early forms of attentional orienting and engagement. Telles and colleagues (2014) explained the neurophysiological mechanisms underlying the benefits of slow deep breathing prescribed by yoga texts and tradition. They differentiated between metabolic and behavioral breathing. Yoga regulated breathing acts as both a top down and bottom up mind-body practice. There is sufficient neuroanatomical evidence to support the idea that apart from the metabolic (chemoreceptor based) regulation of breathing internal and external factors influence breathing, which has been called behavioral breathing. Connections between cortical regions and brainstem respiratory neurons indicate that influences from higher centers can modify metabolic breathing. Slow breathing had a balancing effect on the autonomic nervous system via enhanced parasympathetic activation. Slow and deep breathing stimulates inhibitory signals induced by stretch and hyperpolarizes cells, leading to synchronization of neural elements in the heart, lungs, limbic system, and cortex. Slow breathing also enhances vagal activity leading to optimal psychophysiological arousal, decreased sympathetic activity, and stress responses. The three stages of inward meditation include a) Dharana meditation, focused attention with effort, b) Dhyana meditation, sustained effortless attentiveness or mindful presence, and c) Samadhi meditation, self-integrated, nondual state of awareness. Neurophysiological effects of dharana and dhyana were compared by (Telles et al., 2016). In dharana with effort, there is relative activation of the sympathetic nervous system, whereas, in the effortless dhyana, there is relative parasympathetic dominance. In samadhi, there is optimization of conscious arousal, awareness, and the self-integrated experience. Deepeshwar et al. (2019) described four different mental states of consciousness including cancalata (random thinking), ekagrata (non-meditative focusing), dharana (focused meditation), and dhyana (meditation) as defined in yoga texts. Meditation is a self-regulated mental process associated with deep relaxation and increased internal attention. These authors described three higher states of meditative consciousness namely a) Samadhi as the merging of the seer-seen duality with a three dimensional awareness; b) higher Samadhi with all-pervasive awareness; and c) the highest state of consciousness with a feeling of oneness with reality, and self-freedom, which is called Kaivalya, the infinite awareness with absolute self-absorption and freedom from the meditator with his/her self-centered motivations, biases, and limited perspectives. Gabois and Boudrias (2019) reviewed the effects of yoga on motor performance, body awareness and the experience of pain. They found that yoga has a positive effect on learning rate, speed and accuracy of a motor task by increasing attention and decreasing stress through a better control of sensorimotor rhythms. Yoga also seems to improve sensory awareness and interoception, increase the sense of embodiment (and embeddedness) and consequently to decrease fear. It was also found to regulate autonomic input, increase parasympathetic activity, and promote associated self-regulation. These effects were associated with increased theta power in the ACC and high-frequency heart rate variability. Yoga was also shown to reduce the threat signal, increase pain tolerance, decrease pain unpleasantness, and decrease the anxiety and distress associated with pain. Those effects on pain signals were primarily associated with the recruitment of specific brain areas such as the insula, the amygdala, and the hippocampus. Yoga could improve physical and mental health through down-regulation of the Hypothalamo-Pituitary-Adrenal (HPA) axis and the sympathetic nervous system (overactivation). 81

 The Neurophilosophy of Meditation

REFERENCES Adi, S., & Swami, V. (1938). Aparokshānubhuti. Advaita Ashrama. Chinmayānanda, S. (1983). Mandukya upanishad with karika. Central Chinmaya Missin Trust. Cisek, P. (2019). Resynthesizing behaviour through phylogenetic refinement. Attention, Perception & Psychophysics, 81(7), 2265–2287. doi:10.375813414-019-01760-1 PMID:31161495 Csikszentmihalyi, M. (1990). Flow: The psychology of optimal experience. Harper Row. Dandekar, S. V. (1953). Sārtha jñāneshwari. Prasaad Prakashan. Deepeshwar, S., Nagendra, H. R., Rana, B. B., & Visweswaraiah, N. K. (2019). Evolution from the four mental states to the highest state of consciousness: A neurophysiological basis of meditation as defined in yoga texts. Progress in Brain Research, 244, 31–83. doi:10.1016/bs.pbr.2018.10.029 PMID:30732843 Deshmukh, V. D. (1990). Presence: The key to mental excellence. Sunanda V. Deshmukh. Deshmukh, V. D. (2004). Turiya: The fourth state of consciousness and the STEP model of self-consciousness. Journal of the Interdisciplinary Crossroads, 1, 551–560. Deshmukh, V. D. (2006). Neuroscience of meditation. TheScientificWorldJournal, 6, 2239–2253. doi:10.1100/tsw.2006.353 PMID:17370019 Deshmukh, V. D. (2011). Vedic psychology: A science of wisdom. Journal of Alternative Medical Research, 3, 29–43. Deshmukh, V. D. (2012a). Patanjali’s ashtanga yoga: a neurologist’s perspective. Retrieved from www. ancientindianwisdom.com Deshmukh, V. D. (2012b). Breath-meditation: Prana-dhyana. Journal of Alternative Medical Research, 4, 163–169. Deshmukh, V. D. (2020). Jala-Bhramari, OM chanting, and kaivalya: A neuroscience perspective. Yoga Mimamsa, 52, 38–44. Dossett, M. L., Fricchione, G. L., & Benson, H. (2020). A new era for mind-body medicine. The New England Journal of Medicine, 382(15), 1390–1391. doi:10.1056/NEJMp1917461 PMID:32268025 Gadbois, E. R., & Boudrias, M. H. (2019). What are the known effects of yoga on the brain in relation to motor performances, body awareness, and pain? A narrative review. Complementary Therapies in Medicine, 44, 129–142. doi:10.1016/j.ctim.2019.03.021 PMID:31126545 Gard, T., Noggle, J. J., Park, C. L., Vago, D. R., & Wilson, A. (2014). Potential self-regulatory mechanisms of yoga for psychological health. Frontiers in Human Neuroscience, 8, 770. doi:10.3389/ fnhum.2014.00770 PMID:25368562 Harris, D. J., Vine, S. J., & Wilson, M. R. (2017). Neurocognitive mechanisms of the flow state. Progress in Brain Research, 234, 221–243. doi:10.1016/bs.pbr.2017.06.012 PMID:29031465

82

 The Neurophilosophy of Meditation

Herrero, J. L., Khuvis, S., Yeagle, E., Cerf, M., & Mehta, A. D. (2018). Breathing above the brainstem: Volitional control and attentional modulation in humans. Journal of Neurophysiology, 119(1), 145–159. doi:10.1152/jn.00551.2017 PMID:28954895 Jha, G. (1934). Yoga-darshana: sutras of patanjali with bhashya of vyasa. Theosophical Publishing House. Josipovic, Z. (2019). Non-dual awareness: Consciousness-as-such as non-representational reflexivity. Progress in Brain Research, 244, 273–298. doi:10.1016/bs.pbr.2018.10.021 PMID:30732841 Kak, S. (2016). Mind and self: Patanjali’s yoga sutra and modern science. Mount Meru Publishing. Kokoszka, A. (1992). Relaxation as an alternate state of consciousness: A rationale for a general theory of relaxation. International Journal of Psychosomatics, 39, 4–9. PMID:1428618 Krishnamurthy, V. (2019). Meet the ancient scriptures of Hinduism. Notion Press. Krishnamurti, J. (1978). The Wholeness of life. Harper & Row Publishers. Krishnamurti, J. (1979). Meditations. Harper & Row Publishers. Krishnamurti, J. (1983). The Flame of attention. Harper & Row, Publishers. Leggett, T. (1992). Shankara on the yoga-sutras. Motilal Banarasidass Publishers. Lutz, A., Mattout, J., & Pagnoni, G. (2019). The epistemic and pragmatic value of non-action: A predictive coding perspective on meditation. Current Opinion in Psychology, 28, 166–171. doi:10.1016/j. copsyc.2018.12.019 PMID:30711914 Mukerji, P. N. (1963). Yoga philosophy of patanjali by Swami Hariharananda Aranya. State University of New York Press. Quesque, F., & Brass, M. (2019). The role of the tempo parietal junction in self-other distinction. Brain Topography, 32(6), 943–955. doi:10.100710548-019-00737-5 PMID:31676934 Radhakrishnan, S. (1948). Bhagavad-gita. George Allen & Unwin Ltd. Radhakrishnan, S. (1953). The principal upnisads. Harper Collins Publishers. Satchidananda, S. (1978). The yoga-sutras of patanjali. Integral Yoga Publications. Schoenberg, P. L., & Vago, D. R. (2019). Mapping meditative states and stages with electrophysiology: Concepts, classification, and methods. Current Opinion in Psychology, 28, 211–217. doi:10.1016/j. copsyc.2019.01.007 PMID:30785068 Sharvananda, S. (1958). The vedas and their religious techniques. In K. Bhattacharyya (Ed.), The cultural heritage of India. Ramakrishna Mission. Singh, J. (1963). Praty-abhijnaana-hridayam: The secret of self-recognition. Motilal Banarasidass Publishers. Swarupananda, S. (1909). Srimad-bhagavad-gita. Kessinger Publishing Co. Taimni, I. K. (1961). The science of yoga. The Theosophical Publishing House.

83

 The Neurophilosophy of Meditation

Telles, S., Gerbarg, P., & Kozasa, E. H. (2015). Physiological effects of mind and body practices. BioMed Research International, 2015, 1–2. doi:10.1155/2015/983086 PMID:26171397 Telles, S., Singh, N., & Balkrishna, A. (2014). Role of respiration in mind-body practices: Concepts from contemporary science and traditional yoga texts. Frontiers in Psychiatry, 5, 167. doi:10.3389/ fpsyt.2014.00167 PMID:25505427 Telles, S., Singh, N., Gupta, R. K., & Balkrishna, A. (2016). A selective review of dharana and dhyana in healthy participants. Journal of Ayurveda and Integrative Medicine, 7(4), 255–260. doi:10.1016/j. jaim.2016.09.004 PMID:27889426 Woods, J. H. (1914). The yoga-system of patanjali. Motilal Banarasidass Publishers.

84

85

Chapter 6

The Yoga of the Bhagavad Gita: Spirituality, Meditation, and the Rise of a New Scientific Paradigm Rubens Turci Universidade Estadual do Rio de Janeiro, Brazil

ABSTRACT This chapter explores the fact that śraddhā (truth-force or heart force; love-in-action, self-reliance, faithin-oneself, and also enthusiasm) may be said to represent the key noticeable outcome of spirituality, yoga, and meditation, and also that śraddhā can be seen as the main category to bridge the gap between science and spirituality. Being itself the truth-force that is common to scientists (atheists or not) and people of different faiths, śraddhā represents a conceptual category that helps us to explain how science evolves shifting from one paradigm to another.

INTRODUCTION This chapter explores a series of thoughts on meditation, spirituality, and science in a manner that does not necessarily equal a journal article. It takes advantage of a wider space for reflection on broad topics and a license to be less systematic. For a start, it considers the current discussion brought by Adluri and Bagchee (2014a, 2014b) about the concept of science suitable to the humanities and introduces a discourse on method. Then, it addresses the candidate proposed by Turci (2015) to fulfill the role of a new scientific paradigm, capable of bridging the gap between science and spirituality. It suggests that the Bhagavad Gītā (referred to below as Gītā) provides the grounds for such a paradigm when it introduces the art and science of meditation. Ultimately, it shows that this new paradigm can be understood and translated in Western grounds as soon as one realizes that śraddhā (truth-force or heart force; love-inaction, self-reliance, faith-in-oneself, and also enthusiasm) accounts for the unity of the Gītā as well as for the two following corollaries of Turci’s (2007) Ph. D. dissertation “Śraddhā in the Bhagavad Gītā:” (Bhagwad Gita = the Song of the Lord, Circa 200 B.C.).

DOI: 10.4018/978-1-7998-3254-6.ch006

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 The Yoga of the Bhagavad Gita

(1) the fact that śraddhā may be said to represent the key noticeable outcome of spirituality, yoga, and meditation, and also; (2) that śraddhā can be seen as the main conceptual category to bridge the gap between science and spirituality. The discussion about the centrality of śraddhā in the Gītā adds to the discussion, for instance, how powerful technology available today used to investigate brain imaging, can provide enough data to explain the material and/or spiritual nature of all human experiences. On the one hand, scientists are trying to show that brain chemistry and the functioning of neurotransmitters through all neural networks can explain our spiritual feelings. Their underlying hypothesis is mostly based on the belief that in explaining the operations of the brain one will find proof for the non-existence of the soul. Since the results of Newberg and d’Aquili (2008) many researchers have argued that all experiences attributed to the soul represent nothing but consequences of the way the brain operates. On the other hand, however, unless one believes that the brain produces the reality of this universe, one can never assign to the brain’s fantasy what may be out there in some metaphysical realm. Furthermore, it is not difficult to find in our days a researcher of the brain discussing areas that are activated during meditation. There has been a continuous and broad effort by scientists around the world to solve the mysteries involved in what one perceives as a religious experience. Increasing scientific evidence indicates that yoga and meditation are useful in mental health, bringing a combination of mental alertness with physiological rest (Kumar et al., 2010; Dentico et al., 2018). The Gītā summarises the entire philosophy of the Vedas (Circa 1500-1200 B.C.) and the Upaniṣads (Circa 800-300B.C.). The text, as it is well known, is an episode in the Mahābhārata when the climactic battle is about to begin. There, the protagonist, Arjuna, learns from Krishna the ancestral and lost science of pure yoga and meditation. In the first chapter Arjuna is without śraddhā (aśraddadhāna); in the tenth, he acquires brahmavidyā; and finally, in the seventeenth, he is fully endowed with pure śraddhā. In the Gītā śraddhā is ground in spirituality, meditation, wisdom, and science. Turci (2007) discusses how śraddhā appears in order with the Vedic usage and opens the person to contact with the soul-force, which drives the whole universe. Being itself the truth-force that is common to scientists (atheists or not) and people of different faiths, śraddhā represents a conceptual category that helps us to explain how science evolves shifting from one paradigm to another. The work of Turci (2015) shows that the ability to recognize truth and the laws of nature comes from śraddhā. The Latin term “credere”, which is usually rendered by the English term “creed”, can be traced back to the same Sanskrit root as the term “śraddhā” (Smith, 1979). Their etymology suggests the general meaning of “to place one’s heart on”. Both terms are rooted in the primitive Indo-European term “kred-dhe”. Furthermore, their metaphorical sense of “putting one’s heart on” suggest that they do not relate to any creed in particular, but to that truth-force which lies beyond all beliefs. The philosophical category defined by the term “śraddhā” helps us to develop a deeper understanding of the kind of experience people have, when they state that their “spiritual” experiences make them lose their sense of “I” and “feel” one with the Absolute. Through the lenses of śraddhā, we are able to understand the outcome of experiences such as the Buddha’s enlightenment, Gandhi’s display of truth-force (which he called satyāgragha), and the wholehearted speech of Martin Luther King, “I have a dream”. They all can be said to represent more than a flash parietal lobe quieting down. Even when, in terms of brain stimulation, the superficial enthusiasm involved in these processes can be artificially reproduced in laboratories, the real outcome of them, śraddhā, cannot. Śraddhā can be scientifically measured only 86

 The Yoga of the Bhagavad Gita

in terms of one’s gain of altruistic behavior1 and loss of egoism. It expresses the power to overcome the hurdles of life as well as the joy of accomplishing truth in science and all fields of knowledge. Śraddhā; therefore, helps us to evaluate the observed benefits of the spiritual practices, which are also associated with levels of stress, anxiety, and so on. According to the Gītā, the brain (buddhi) is like the moon, shining only because of the light which comes from the sun, which is a symbol for the soul (Ātman). This metaphor suggests that one cannot disprove the existence of the soul just by describing and explaining all the phenomena of the moonlight. Changes in the frontal lobe of the brain may be seen as the effects, not the cause, of one being religious and compassionate. Therefore, one cannot truly say one has had a relevant transcendental experience unless one is fully transformed and becomes capable to display this individual truth-force and capacity to overcome the hurdles of life named śraddhā. In other words, you cannot reduce a spiritual experience to a pure brain phenomenon, no matter one’s assumption about the existence of a god. The Gītā affirms in the third verse of chapter seventeen (BhG 17.3) that whatever amount of śraddhā one has, thus he or she is (BhG 17.3). Everyone is made out of his or her levels and quality of śraddhā2 in accordance with the theory of the three guṇas (strands), as will be shown in the next sections. These guṇas are at the core of the establishment of all minor and provisory truths. Error and illusion are relative and necessary steps, both being summed up in the ladder of truth. Śraddhā is always in accordance with the relative truth of each person. It does not require as a sine qua non that one believes in any God. Many practitioners of Buddhism do not believe in any God, yet they do so out of their śraddhā (Pali: saddhā). In Theravada Buddhism śraddhā means faith plus knowledge, trust, and devotion. It is regarded as the object of absolute confidence, recollection and contemplation, devotion and worship. Graeme MacQueen’s interesting but unpublished “The Concept of Faith in Early Buddhism” (1972), written for W. C. Smith, in a seminar at Harvard University, addresses the Dīgha and Majjhima Nikāyas of the Buddhist Pāli Canon and throws light over the meaning of saddhā (in Sanskrit, śraddhā). MacQueen (24) states: The Buddhists are forced to show that their own faith is properly grounded and acceptable. This is done in two ways: (i) it is said that before “putting one’s heart on” the Buddhist path, one should examine the evidence available to him, (ii) it is claimed that the Doctrine (Dhamma) given by the Buddha is ehipassika (“come-and-see-able”). MacQueen’s (28-32) conclusion, then, is that saddhā, a necessary condition for the winning of Nibbāna (in Sanskrit or Skt. nirvāṇa), is fulfilled by knowledge. For, the Buddhist “faith” actually involves what Westerners usually understand by “faith” plus knowledge. There are a couple of passages in the Yogasūtra of Patañjali: With Commentary of Vyasa that also suggests the centrality of the concept of śraddhā to understand and to access truth. When commenting on sūtra I.20, where the term śraddhā occurs, Vyāsa says, “samādhi is preceded by śraddhā” (Dwivedi, 1980). Further, he adds, “śraddhā is the firm and agreeable conviction of the mind regarding the efficacy of yoga.” Then, he concludes: Energy (vīrya, power, valor, strength, dignity, heroic deed; from “vīr, to overpower, to be valiant) is born in him who pursues knowledge with śraddhā. Memory comes to help when he is possessed of energy. On the appearance of memory, the mind ceases to be disturbed and passes into samādhi. When the mind is in samādhi, discrimination appears, by which the object is known as it is (Dwivedi, 1980).

87

 The Yoga of the Bhagavad Gita

Vyāsa’s discussion on sūtra I.20 points to the core questions addressed in this chapter: (1) how śraddhā is related to the mental energy associated with the concrete experience by which the object is known as it is; and (2) how spirituality and science can be seen as a unified field, capable to account for all medical results associated with yoga, spirituality and meditation, and, as consequence, for the establishment of a new scientific paradigm. According to the Gītā, one may say that reason alone is insufficient to lead us from the experience of the phenomenological world (saṃsāra) to that of the sacred realm (nirvāṇa). While the more inorganic level of reality can be known through the senses and reasoning, the organic and sentient world seems to contain secrets only accessible through intuition, mostly in a state of contemplation, or deep meditation, when one is endowed with the highest levels of pure (sattvic) śraddhā. How it is that the brain reflects this sense of soul-force named śraddhā is something still beyond the reach of natural science. It is something which remains in the realm of spirituality – a category whose main concrete feature is the universal śraddhā, which is at the core of all philosophical and religious insights. Spirituality, whatever its form, presupposes śraddhā, and śraddhā lights the path, developing itself within that life where one nurtures self-reliance3. Śraddhā is an outcome derived from certainty. No matter one’s dogma of faith, it is always possible to develop the universal truth-force called śraddhā. For, śraddhā expresses what a sincere practitioner of any creed and also the atheist manifest when they experience to their hearts the true nature of their beliefs. Even when the brain chemistry may reproduce states that constitute fake beliefs, it cannot truly account for the full experience śraddhā represents. It is śraddhā, which cannot be induced by any artificial means, that brings about an unbiased and positive perspective on life as a whole.

THE NAY SCIENCE The scope of this section is to discuss the alternative method of the whole-hearted textual criticism used here to offer one plausible and non-idiosyncratic interpretation of the Gītā among many4. This method is fully explained by Adluri and Bagchee (2014a). Those who are following the current debate on Indology5 are very aware of the breakthrough contribution for this field of their discussion about the concept of science suitable to the humanities. Their The Nay Science: A History of German Indology (2014b) raised very important questions. How should we read and interpret ancient texts? How do we approach these texts scientifically and critically? What is the influence of method on a non-Western episteme? Why we must question established paradigms? The textual interpretation of the Gītā that resulted in the present chapter has remained as close as possible to their findings. The Gītā was treated as a whole, rather than dissected in pieces, as advocated, for instance, by Paul Hacker (1961) and other Indologists fond of the hermeneutics of suspicion6 and its historical-critical method, which Adluri and Bagchee demonstrate are neither historical nor critical (Adluri & Bagchee, 2014a). Their book dedicates the first three hundred pages to investigate how scholars have approached the Mahābhārata and the Gītā. The significance of this task lies in its successful demonstration in the last hundred and fifty pages of the failure of the “historical-critical method,” rooted in an imagined layering of the text over time by competing authors from different schools of thought. The book undertook a comprehensive inquiry into the application of the Western text-historical method, used by those who claim that the works of non-conforming scholars, particularly in the Gītā, should be disregarded as non-scientific. The Nay Science opposed this method7. It suggested that Western Indologists’ deconstructions of the Gītā had produced poor results if any.

88

 The Yoga of the Bhagavad Gita

It is a well-known fact that the radical separation of theology and philosophy in the West was meant to safeguard the Christian understanding of religion as a dogma of faith (Sharma, 1987). The scientific paradigm had posed a serious threat to Christianity which could be counteracted only through a reiteration of faith in the Biblical God. When Indologists accuse Indian philosophy of being “religion,” they show the symptoms of their trauma. Their effort to separate the light of natural reason and the light of heart is a consequence of the fall of their dogmas about the natural world, wrongly entangled with their religious faith. Advocating that true philosophy cannot say anything about the soul and the universe, they remained attached to the Christian framework of Hegel’s secularism. Hegel (1826) wrote a treatise on Humboldt’s German translation of the Gītā and found there an opportunity to repress the romantic philosophers, which considered the Mahābhārata with the Gītā forming a unity. One may argue that Hegel feared the romantics because their willingness to acknowledge Indian thought represented a potential challenge to his Lutheran faith. Yet, philosophy can never be seen as an activity unrelated to myth and spirituality, as the ancient world had already demonstrated. It is not a surprise that The Nay Science ends with Gandhi’s interpretation8 of the Gītā. Gandhi’s hermeneutical method enriches his inherited tradition and makes various concepts of the Gītā relevant to his present. Gandhi’s allegorical method linked Arjuna’s story with his experiments with the truth. Gandhi’s satyāgraha (or śraddhā) represents his intuitive method, but he based it in the Gītā which he took to be a consistent scriptural treatise, whose main topic was the method he was learning to apply. Through his reading, he came to understand the best way to represent what truth (satyam) is and to act accordingly. Gandhi’s interpretation of the Gītā, which is not considered scholarly, preserves the core meaning of the text even though he has never been concerned with Indology at all. He approaches the text with the desire for the good, guided by love and the pursuit of wisdom. We must remember that the Gītā is above all, a piece of literature, which came out in verse form. It is the work of a philosopher who is also a poet. Its heart philosophy puts us in front of some perennial truths that can always be explored further. This means that the universal method, the new paradigm we are in the search for reuniting spirituality and science, must be rooted in our inner struggles to live a meaningful life. To what extent are we engaged in the pursuit of good, truth, and beauty? This is what śraddhā is all about, as we will see next.

DISCOURSE ON METHOD The Gītā is not a naive expression of ethics of intentions. It does not enforce any reward-punishment mechanism to develop or improve one’s “intentions”, but it rather casts light on the issue of human nature to help one to improve one’s motivation, or motives for acting. For, “intentional” mostly means “done with a selfish purpose”. “Motive”, however, as the reason for acting, represents that motivation or enthusiasm which urges a person to act in a virtuous way. Without, proper motivation, śraddhā becomes either rajasic, or tamasic, representing merely blind faith and dogmatism, whereas truth is nothing but an outcome of sattvic śraddhā. In the Gītā, Krishna provides strong reasons for Arjuna becoming himself motivated, despite Arjuna’s good intentions when considering not fighting. The implicit method that the Gītā teaches is that no depreciation of any opinion whatsoever should ever be seriously undertaken by the faithful seeker of truth. Indeed, it is a necessary corollary of the view embodied in its seventeenth chapter that every opinion, embodies one part of truth, and every one and all of the opinions that appear to be refuted in the text are held by logical thinking (anumāna) in 89

 The Yoga of the Bhagavad Gita

a transmuted form, in accordance with the triplet borne in the one and same consciousness: thesis or sañkalpa, antithesis or vikalpa, and synthesis or anukalpa. Vikalpa, or doubt, which is at the core of the scientific method, is the opposite of the initial cognition process of apprehension of truth, or sañkalpa. In the Gītā, the unfolding of knowledge, concealed because of Arjuna’s initial lack of śraddhā, appears as his enquiry and finding. Arjuna can ask about all things because he knows a little about them all and wishes to know more. As the Gītā says, only when the soul sees the Many rooted in the One and also branching out from that One, knowledge becomes complete and perfect, the Infinite becomes fulfilled and realized in that soul, the soul identifies itself with the All-Self, or Brahman (BhG 13.27). The inner Self (with capital letter) and the concrete self (lower case) are the two irreducible facts and factors of all Consciousness. The concrete self (An-Ātman) is held together only by the abstract Self, the two being always inseparable, though always distinguishable. The concrete forces a way through the senses outwards; hence the Jīva, or the person, sees the outward world. When one turns his or her gaze inwards, then he or she fulfills the Delphic commandment “know thyself!”, the abstract Ātman. The very being of the Ātman is the negation of all non-Self, all that appears as an object, and that can be known by the knower subject, the pure, universal, and abstract being of the “I” (Aham9) – the Supreme Self, also known as Ātman. The “I”, the highest nature of Brahman, the Self, is the origin and end of the entire universe; it is the womb of all beings (BhG 7.6). All multiple worlds are strung together on the “I”, even as jewels on a thread (BhG 7.7). The “I” and the “not-I” always imply each other and can never be separated. They carry out on each other, one another’s characteristics: the “I” becomes particularized into individuals, and the “Not-I” becomes generalized into all kinds of matter. The Gītā deals with the science of becoming10. It explains being (sat), not-being (asat), and becoming11 (sva-bhāva) through the analysis and synthesis of the constituents of AUM (generally written OṂ): the singularity (A), the duality (U), and the multiplicity (M). AUM is the most sacred syllable symbol and mantra of Brahman, the praṇava12 in which the whole art and science of meditation is hidden. “Know thyself!” is equivalent to the idea or consciousness “I-this-Not (am)” of the praṇava AUM, as discussed in the Atharva Veda, where “A” stands for the Self (Ātman), “U” for the non-Self (An-Ātman; Prakṛti) and M, for their interplay (Līlā), or relation of negation (Śakti13) which exists between them. There is nothing beyond and outside of this Primal Trinity, which constitutes the Para-Brahman. The simultaneous affirmation and negation involved in this relation constitute the triune Brahman, the Absolute, the Changeless accomplished in nirvāṇa, as well as the storehouse of all change, experienced in the world. The interplay between the Self (“A”) and the non-Self (“U”) gives rise to the Jīva and his/her different triads of attributes related to the psychology of cognition, desire, and action. As regards the inclusion of both Self (Ātman) and non-Self (prakṛti) in Brahman, the Gītā says that both the perishable non-Self and the unperishable Self are latent in Paramātma, or Brahman (BhG 15.16-7). It is a scientific fact that we begin with a piece of partial knowledge and end with a fuller one. None can turn attention to that of which he knows nothing at all. To start on any quest we must have at least heard of the subject as existing. This initial knowledge is very different in fullness from the knowledge we should acquire on that subject. So our method of investigation should take into account that every single piece of past knowledge is connected to the present knowledge even as jewels on a thread. The well-established conclusion (siddhānta) of the Gītā is that the view, the opinion, varies with the individual’s different levels and quality of śraddhā, which endows us with different angle of vision. Yet, no matter anyone’s point of departure, in each one’s final view, from the universal standpoint of śraddhā, he or she will arrive at a synthesis that includes and harmonizes all previous views. For the Gītā the impelling motive (prayojana) of the search for Truth is the heart-craving of the Jīva for knowing the 90

 The Yoga of the Bhagavad Gita

unified Self and finding peace. From such working of the Jīva’s doubts arise the multifarious forms of śraddhā, beginning with faith and belief and ending in trust, confidence, self-reliance, and certainty.

THE SEARCH FOR A UNIVERSAL METHOD AND THE RISE OF A NEW SCIENTIFIC PARADIGM Which comes first, the feeling14 (bhāva; sentiment, emotion) or the thought (cintā)? It is hard to answer, but this and the following sections will only make sense to the reader who assumes that feeling takes precedence over thought and even generates it. It is supposed here that feeling expresses the real necessity of the Self. The understanding of what necessity alone is, in any field to where one directs his or her attention, comes from dhyāna15. “Dheya”, the object of each science, is arrived at out of necessity. There is no real advance in science without necessity. Just by observing nature one could state that sunrise and sunset regularity expresses a general law of necessity, named Ṛta (law, order, truth) in the Vedic tradition. The concepts of dharma (that which follows from Ṛta) and karma (the spiritual principle of cause and effect) have come out of Ṛta. Ṛta is a natural law, which controls and directs the operation of the cosmos and everything within it. At the heart of the Gītā, there is this understanding of Ṛta in the form of the great spiritual science or Brahma-vidyā by which worlds are built. The lesson the Gītā offers seems to be that every action is motivated by a feeling or emotion that, after arising and being experienced, is registered in the mind. Experiencing Ṛta in meditation as a feeling triggers the logical thought about the natural cosmic order and promotes the development of all sciences. The intuition (arising as a feeling) about the proper development of any particular science (Ṛta) fulfills ones’ soul with certainty and self-reliance, or śraddhā. When a thought, a concrete hypothesis comes to one’s mind, it is because on a less conscious level he or she had already had the feeling that something new was in need to be expressed. According to the Gītā, both, the thrill caused by these feelings giving form to a new thesis, and the creative force that triggers thought and its scientific methods, come from buddhi16, the metaphorical heart. This same point of view is assumed in the thesis (sañkalpa) being established here. It negates (vikalpa) the current understanding that science and spirituality cannot be brought together under the same paradigms and offers a synthesis (anukalpa) in the form of a new paradigm, harmonizing science and spirituality. Throughout the whole Gītā, certainty arises from śraddhā and that is why śraddhā has sometimes been defined as āstikya-buddhi, or trusting judgment, or affirmative conviction. W.C. Smith touches on the issue of the universality of śraddhā in Faith and Belief (1979). He offers a thorough discussion on the view of śraddhā as āstikya-buddhi and makes it an orientation of yes-ness, an “awakeness to transcendence” that belongs to the higher visionary self in the “heart,” and not to the rational mind or manas. One may hint from there that the meaning śraddhā conveys implies the so-called reconciliation of science and spirituality. Arising from the vital stimuli received from the soul (Ātman) through its representative (i.e., buddhi, the power of forming and retaining conceptions; intellect, reason, discernment, and so on), śraddhā conveys a meaning impossible to be translated by a single term in any other language. And that is why it cannot be fully translated as “faith,” which comes from the Latin “fides.” The term “fides” fails17 to convey the meaning of unconditional freethinking, which is the characteristic mark of scientific thought. On the contrary, “fides” expresses the medieval thought, grounded in the Catholic Church Fathers’ paradigm “fides quaerens intellectum,” which defined for centuries how faith (fides) was conditioned by belief. If the term “fides” were a fit translation for the Sanskrit “śraddhā,” it 91

 The Yoga of the Bhagavad Gita

would be impossible disproving St. Anselm’s “fides quaerens intellectum,” as Turci (2015) pointed out. For, it is a well-established truth that “no belief can become knowledge without any evidential support.” (DuJardin, 2019) And that is precisely why the half-breed expression “śraddhā quaerens intellectum” coined by Turci (2015) was proved to pass both the scientific and spiritual requirements. Śraddhā reminds us that we must look inwards and work from the inside out, turning our eyes in a direction opposite to that in which we usually employ them in ordinary science, but without losing sight of our rationality. Since this new paradigm, rooted in a poetical reading of nature, bridges the gap between the realms of spirituality and science, it cannot be easily dismissed. For the most, successfully answering scientific questions is a matter of insight and a scientific method that connects us to our subject. On this matter, there is an old fashioned but interesting book, The Psychology of Philosophers (1929), written by Alexander Herzberg, which considers the personal characters and the views of many western thinkers. He manages to show that there would be less contradiction between philosophers and their philosophies if they had managed to restrain their selfish traits properly and were more willing to find truth than to claim originality. His findings point out the same direction as those of The Nay Science. As Adluri and Bagchee (2014b) suggest in the book all methodological approaches to human sciences should take into account the mastering of the self by following our heritage of integration of heart and reason.

BRIDGING THE GAP BETWEEN SCIENCE AND SPIRITUALITY The purpose of this section is to discuss the gap between empirical science and rational spirituality from the viewpoint of the spiritual laws that are said in the Gītā to govern the world-process evolving out of the sacred three-lettered word-sound AUM. Thoughtful reading is central to understanding that we can reach truth also through the language of myth and poetry18. This is the central point when it comes to the Gītā, which is an art object in itself, and that is why its subject is meaningful for the present. The Gītā is a complex literary and philosophical text that challenges our rationality. It reveals to us how to read the nuances of the world. It shows how science and spirituality are implied in the symbolic meaning of the sacred syllable (praṇava) AUM. It is implicit in the AUM the idea that śraddhā is the third constituent (M), which makes up every person (Jīva), out of the relation of the Self (A) and the non-Self (U). According to the Gītā, this universe can be described through the language of the triplets, grounded in the fundamental axioms about being, non-being, and becoming. Out of the relation of the Self to the non-Self, the triune nature of Saguṇa Brahman (AUM), manifested as Sat (universal being), Cit (universal consciousness) and Ānanda (universal bliss), imposes a triplet of attributes (guṇas19) in both. Inside the Jīvas, they appear as the faculties of activity (kriyā), cognition (jñāna), and desire (icchā) respectively. Furthermore, in the outside world these attributes are sattva (rhythm, harmony, goodness, truth, illuminating power, spirituality, being), rajas (mobility, passion, explosiveness, restlessness) and tamas20 (inertia, dullness, chaos, heaviness, substantiality, materiality, unconsciousness). The second and third triplets are nothing but the particular and manifested aspect of the universal and non-manifested first triplet. Here it will be of interest to focus on this third triplet. The three guṇas sattva, rajas, and tamas, are utterly inseparable though distinguishable. They are manifest in everyone’s nature and in all daily activities, with one of them preponderating and subordinating the other two, following one’s mood (BhG 18.19-40). In short, one may say that sattva corresponds to spirituality and science; rajas, to blind passion; and tamas, blind clinging and irrational carelessness that threatens life. 92

 The Yoga of the Bhagavad Gita

The trinity (AUM) is realized as the One (OṂ). “A” stands for Ātman, the Self. “U” stands for Anātman, the non-Self; and “M,” for the Niṣedha, the relation of negation that connects them both. So, the Self is present in everything, and without it, no manifestation can take place. In every person (Jīva), being and not-being are mutually immanent, as seed and tree. Thus, the Self may be seen as one with the world, and every relation among objects is essentially a denial of the difference between Self and world. Hence, the symbol AUM connotes Brahman, the state of unity of the Many. The One and the Many are the same, as we can infer from the many plants arising from one seed, and vice-versa. So, the seed is the denial of the plant in the same sense intended by Hegel is his Phenomenology of the Spirit (1807). The seed and the plant are both equal to and different from each other. Under the dialectical science of AUM, all different things are included, as well as their differences extinct. Despite Hegel’s critique of the Gītā, his most important contribution to philosophy seems to be a full development and application of the law that two opposites, a thesis and an antithesis,21 always find their reconciliation in a third something, a synthesis, which is neither the one nor the other exclusively but both taken together. Furthermore, modern science already acknowledges that its business is nothing but to trace unity in diversity, through its general laws of “becoming”, which is saṃsāra the endless combinations of the triplets above mentioned. The idea expressed by the Praṇava AUM is the Sūtrātman, the thread regulator of the modes of being, the beads, so to speak, strung thereon, which make up the rosary of the whole world process, or its web-soul. This idea of the thread-soul, the Sūtrātman, expresses the unity of concrete science and spirituality. When doing science, scientists concentrate their minds on the particular object of their attention with the hope to reach and contemplate the general law that rules their field of expertise. In doing so, they improve their faculty of abstraction, which is nothing but the counterpart of what we name “spirituality”, or the capacity to go from particulars to universals; from the multiple to the one.

THE ART AND SCIENCE OF MEDITATION REVEALED IN THE GĪTĀ Stating that “Yoga is skill in action” (BhG 2.50), Krishna teaches Arjuna not to be attached to the fruits of actions, or the Vedas (BhG 2.53), but to leave behind all personal desires (kāma), withdrawing the senses from the sense objects (BhG 2.58). Such a capacity of carrying out actions through the senses, mind, and reason, without desire (kāma), or selfish interest becomes possible through meditation. For, if an action is motivated by the impure purposes it cannot be said to be Ātman-para, or rooted in, or motivated by, Ātman, which is the real Self. This idea of “Self-interest” (capital letter) is not easy to grasp in English, as the expression “self-interest” (lower case) denotes those who are motivated by individual gain. The “Self” referred to here is the Ātman or divine essence, not the individual ego. Therefore, it is only altruistic actions, performed without selfish interest, that represent the interest of the real and universal Being or Self (Ātman). Having explained all that, Krishna equates himself with Ātman, states Ātman’s oneness, and explains how to act dwelling in the Ātman (BhG 6.30-1). As a result of Arjuna’s will to fit into the divine praxis (BhG 18.73), as well as his will to know about the divine (BhG 10.17-8), he reaches the promised (BhG 9.5) state of meditation (BhG 11.9-34) in which he is able to contemplate the universal form of the divine (viśvarūpa-darśana-yoga). We can explain the art and science of meditation revealed in the Gītā by a process of reflexing in the trinity of the praṇava AUM, observing how the infinite saṃsāra originates and is within Brahman. Insightful thinking (pratyakṣa) appears when the mind (chitta), its modifications (vṛtti) and their control (nirodha) are summed up in meditation. The yogi is he who always joins thesis (saṅkalpa) and antithesis 93

 The Yoga of the Bhagavad Gita

(vikalpa) into synthesis (anukalpa); for, he knows that all experiences lead to the understanding that the world is held within the Self, and the Self held within the world. In the Gītā, meditation means, in brief, this contemplation without thoughts of the continuous flow of life arising from the praṇava AUM. This means that meditation is of three kinds: Saguṇa Dhyāna (concrete meditation; associated to the letter “U,” which represents the non-Self); Nirguṇa Dhyāna (abstract meditation, associated to the letter “A,” which represents the Self); and Brahma Dhyāna (transcendental meditation), associated to the letter “M”, which represents the transcendental relation of negation and identity of the Self and the non-Self). Saguṇa Dhyāna leads to the contemplation of the true nature of all sciences; Nirguṇa Dhyāna, to the contemplation of the true nature of spirituality; and Brahma Dhyāna, to their immanent and transcendent organic relation of negation and identity. Meditation, therefore, is related exclusively to the mind. The Gītā is a representative of meditation in the form of the silent dialogue going on inside ourselves. When I ask, “who is this I?”, and also, “who is the witness of my thoughts?” I am acknowledging this double nature where one appears as Jīva, and the other is said to be Ātman. A careful examination of the questions Arjuna offered in the text gives us plenty of information about what meditation truly is: the contemplation of the praṇava OṂ. Meditating on OṂ means to develop the feeling (śraddhā) of identity with the triune AUM, as expressed by the main mahāvākyas (great aphorisms), for instance, those of the Chāndogya Upaniṣad: “sarvaṃ khalv-idaṃ brahma” “All of this is Brahman” (ChU 3.14.1) and “Tat tvam asi”, “Thou art Ātman” (ChU 6.8.7), both encapsulating the meaning suggested in the six metaphors on the art and science of meditation we will explore next.

THE MANTRA WHICH ENVISIONS THE MAIN METAPHORS ON MEDITATION Most forms of meditation begin by inviting us to pay attention to breathing and thinking. As we pay attention to our breath, we begin to witness thoughts. We realize that there are two voices within us as if they were two people: the one who is thinking and the other who is witnessing the flow of thoughts. So far as we know, this idea first appeared in the Ṛgveda, and it represents the first reference in history to the concept of meditation. The passage describes two birds abiding in the same tree: “Two birds of beautiful wings, friends and always united, cling to a common tree; one eats the sweet fruit; the other looks on without eating” (ṚV 1,164.20). The two voices inside us correspond to these two birds. One is the doer of all thinking and the other the witness (sākṣin 22) of the thinking process. This allegory is reproduced in some of the major Upaniṣads,23 for example, in the third chapter of Muṇḍaka Upaniṣad. The second important allegory of the process of meditation appears in the Māntrika Upaniṣad. The text gives an exposition on Brahman in accordance with its changeless (Self) and changing (non-Self) aspects of the one reality. The poetic structure of the text favors the understanding of Ātman as a Swan. Verses 1-2 runs like this: a swan bathes in the calm waters of a lake. The lake is our mind when it has its calming mental waves. The Swan (in Sanskrit; haṃsa) represents the unification process of the two birds of the Rigveda metaphor. Verses 8-13 recounts the metaphor of the two birds above mentioned. Haṃsa is the name of this bird species, which has the special quality of drinking only milk from mixed water. That is, Haṃsa separates the wheat from the chaff; drink the essence of knowledge. He knows, in everything, to seek the essence. Haṃsa is also a word that is the onomatopoeic sound of the breathing process24: I breathe in “ham;” I exhale “sa”. The very breath that gives us life is represented in this sound that designates the bird Haṃsa. Thus, the Haṃsa came to symbolize the prāṇa, the life force,

94

 The Yoga of the Bhagavad Gita

representing perfect balance and unity within life, also reminding us that meditation usually begins with our focus in our breath. The third and main metaphor of the process of meditation is represented by the episode of the Gītā itself. In the Gītā, Krishna and Arjuna symbolize the two birds, described in the Ṛgveda, which are unified in the Haṃsa metaphor. Arjuna achieves communion and unification with Krishna in his own heart. In this allegory, the chariot is the human body. The four horses correspond to the four ways in which the sāṃkhya epistemology conceives our mental structure: intellect, mind, organ of the five senses, and organ of the movement. It is these four components that would correspond to the control, the reins of the four horses. When these reins are under the command of Ātman, here represented by Krishna, we reach the final goal represented by the swan Haṃsa. Any meditation practice developed over time by the different traditions of thought, both religious and secular, derives from this fundamental understanding, exemplified in these three great allegories. Three other secondary metaphors help us understand the fundamentals and basic requirements for meditation practice. They all appear in the second chapter of the Gītā. The first of them is the tortoise metaphor: just as the tortoise collects its limbs into the shell, we must withdraw the senses from the sense objects (BhG 2.58). It means that to succeed in meditation we must redefine our way of acting in the world through the five senses. This reversal movement of the functioning of being is defined by Patañjali in his Yoga Sutra as pratyāhāra (mastery over external influences). Abandoning the functioning regulated by the external world of saṃsāra25, we seek Nirvāṇa. The second accessory metaphor says that an unruly mind is like a boat carried away by the wind (BhG 2.67). The third metaphor states that what is night for the average person is day for the wise; and vice versa26 (BhG 2.69). These six metaphors together are all implied in the mantra “Oṃ namo Nārāyaṇāya,” which has first appeared in the Vedas, being later differently explained in several Upaniṣads, such as the Nārāyaṇa Upaniṣad, the Ātmabodha Upaniṣad, and the Tārasāra Upaniṣad. It is a mantra whose ultimate meaning used to be secret, as acknowledged by Ramanuja, who is known to be initiated into this mantra by his guru. Yet, it may be argued, its secret meaning is fully disclosed in the Gītā, a text that can be read as its final interpretation with all its layers of meaning. “Nāra” means “person”; and “ayana, “shelter”. So Nārāyaṇa means that which is a shelter for all human beings. In the Gītā, quite obviously, Arjuna stands for Nāra and Krishna for Nārāyaṇa. And as such, the mantra “Oṃ namo Nārāyaṇāya” expresses the surrender (namo; namas) of Arjuna (Nāra; the archetypal of a human being) to (whose dative ending is “ya”) Krishna (Nārāyaṇa; the representative of Ātman). Arjuna and Krishna, therefore, may be seen as the two birds referred to in the Ṛgveda and unified as the bird Haṃsa in the Upaniṣads.

CONCLUSION One must not ignore the fact that the Gītā is not as much a sectarian work as it is a source of philosophic and theological inquiry. As it has been shown, the concept of śraddhā helps one to see the theological unity of the Gītā. Furthermore, śraddhā designates not a creed but what lies beyond every creed. It does not define a “religion” but tells the different paths that people can follow in their spiritual journey. It explains that the real practitioner of meditation is he or she who always joins thesis (saṅkalpa) and antithesis (vikalpa) into synthesis (anukalpa), with the understanding that the world is held within the Self, and the Self held within the world.

95

 The Yoga of the Bhagavad Gita

The universal and unifying art and science of meditation established in the Gītā consists of finding the summation so that trinity (AUM) is realized as the One (OṂ). It helps one to answer the questions, “who is this I?”, and also, “who is the witness of my thoughts?” It helps us to understand the interplay between the Self (“A”) and the non-Self (“U”), which gives rise to the Jīva. It is implicit in the AUM the idea that śraddhā is the third constituent (M), which makes up every person (Jīva), out of the relation of negation of the Self (A) and the non-Self (U). Meditation, then, arises as the simple relation of identity and denial of pure consciousness and impure consciousness. It leads to the understanding of the triplet borne in the one and same consciousness in the form of a thesis or sañkalpa, an antithesis or vikalpa, and synthesis or anukalpa. Meditation is explained in the Gītā as a form of cultivation of the mind where even though one is engaged with the objects of the senses, he or she remains under the guidance of Ātman (BhG 2.64). One may say that meditation means Ātman taking control of the mind by one shifting his or her mindset from its outward orientation, or the guṇa-para state, described in BhG 2.67, to its inward orientation, or Ātman-para state, described in BhG 2.58. This is discussed in the Gītā from the viewpoint of Arjuna’s difficulty to withdraw his senses from the objects and its delights (preyas) and to envision the best (śreyas) course of action to undertake. It is when one has one’s mind at the heart that the necessary action can be naturally performed in a spirit full of śraddhā. Ultimately, śraddhā arises as a synthesis that harmonizes the tension between our best spiritual motivations (śreyas) and our material needs (preyas). And that is why it represents a “material” phenomenon manifested as “spiritual” energy in our body through the modulations (guṇas) of prakṛti. From this beginning of the new millennium, meditation is not only accepted as a culture, but it has entered the environment of universities, including the health system. Furthermore, The Nay Science has managed to carefully unmask the pretentious critique of those who were unwilling to interpret the message of the Gītā in their unity. Initially considered countercultural in the West, meditation is gradually becoming a universal ground of all sciences, in accordance with the formula “śraddhā quaerens intellectum”. In conclusion, if we believe that the joy we experience in meditation is real, then this experience must redefine the way we see the world. The meditative discipline becomes something to be lived twentyfour hours per day. Our only concern becomes to act in this world of Saṃsāra, as a way to express it as Nirvāṇa. This is, in a nutshell, the core message of the Gītā.

ACKNOWLEDGMENT This research received no specific grant from any funding agency in the public, commercial, or not-forprofit sectors.

96

 The Yoga of the Bhagavad Gita

REFERENCES Adluri, V., & Bagchee, J. (2014a). Who’s zoomin’ who: Bhagavad gītā recensions in India and Germany. International Journal of Dharma Studies, 4, 1–41. Adluri, V., & Bagchee, J. (2014b). The Nay-Science: A history of german indology. Oxford University Press. doi:10.1093/acprof:oso/9780199931347.001.0001 Bolle, K. W., & Mahabharata, B. (1979). The Bhagavadgītā: A new translation. University of California Press. Butler, E. P., Lenz, J. R., Vargas, A. L. C., McGetchin, D., Sullivan, B. M., Long, J. D., & Bilimoria, P. (2016). Reviews of The Nay Science. Springer. doi:10.118640613-016-0033-9 Chapple, C. K. (2009). The bhagavad gita: twenty-fifth–anniversary edition. SUNY Press. Dentico, D., Bachhuber, D., Riedner, B. A., Ferrarelli, F., Tononi, G., Davidson, R. J., & Lutz, A. (2018). Acute effects of meditation training on the waking and sleeping brain: Is it all about homeostasis? The European Journal of Neuroscience, 48(6), 2310–2321. doi:10.1111/ejn.14131 PMID:30144201 Desai, M. (2009). The bhagavad gita according to gandhi. North Atlantic Books. DuJardin, T. A. (2019). Faith and evidentialism: the concept of faith and its epistemic implications [Doctoral dissertation]. Boston University. https://hdl.handle.net/2144/37985 Dwivedi, M. N. (1980). Yoga-sutras of patanjali: sanskrit text and english translation together with an introduction and notes. Satguru Publications. Easwaran, E. (2007). The bhagavad gita:(classics of indian spirituality). Nilgiri Press. Edgerton, F. (1944). The bhagavad gita, translated and interpreted. Motilal Banarsidass. Emerson, R. W. (1987). Self-Reliance. In J. Porte (Ed.), Essays and lectures (p. 261). The Library of America. Hacker, P. (1961). On the method of historical research into the anonymous Sanskrit literature of Hinduism. Journal of the German Oriental Society, 111, 483–492. Hegel, G. W. F., & Humboldt, W. (1995). On the Episode of the Mahābhārata known by the name Bhagavad-Gītā. Indian Council of Philosophical Research. Kumar, S., Nagendra, H. R., Manjunath, N. K., Naveen, K. V., & Telles, S. (2010). Meditation on OM: Relevance from ancient texts and contemporary science. International Journal of Yoga, 3(1), 2. doi:10.4103/0973-6131.66771 PMID:20948894 MacQueen, G. (1972). The Concept of Faith in Early Buddhism [Unpublished manuscript]. Harvard University. Newberg, A., & d’Aquili, E. G. (2008). Why God won’t go away: Brain science and the biology of belief. Ballantine Books. Olivelle, P. (1996). Upaniṣads: a new Translation. Oxford University Press.

97

 The Yoga of the Bhagavad Gita

Sharma, K. (1987). Bhakti and the Bhakti movement: a new perspective: a study in the history of ideas. Munshiram Manoharlal Publishers Pvt. Limited. Smith, W. C. (1979). Faith and belief. Princeton University Press. Turci, R. (2007). Śraddhā in the Bhagavad Gītā [Doctoral dissertation]. McMaster University, Hamilton, Canada. Turci, R. (2015). Śraddhā in the Bhagavad Gītā: An investigation on the primeval expressions of the contemporary paradigm on heart-philosophy. International Journal of Dharma Studies, 3(1), 2. doi:10.118640613-015-0013-5

ADDITIONAL READING Adluri, V., & Bagchee, J. (2016). Paradigm Lost: The application of the historical-critical method to the bhagavad gītā. International Journal of Hindu Studies, 20(2), 199–301. doi:10.100711407-016-9187-4 Bechert, H., & von Simson, G. (Eds.). (1979). Introduction to Indology: status, methods, tasks. Darmstadt Science Book Society. Bilimoria, P. (2004). Perturbations of desire: emotions disarming morality in the “Great Song” of the Mahabharata. In R. B. Solomon (Ed.), Thinking about feeling. Oxford University Press. Bilimoria, P. (2007). Dismantling normativity in Indian ethics-from Vedic altarity to the Gitas alterity. In Y. Wang (Ed.), The ethical dimension and deconstruction normative ethics in asian traditions (pp. 19–39). Routledge. Bilimoria, P. (2009). A report on indologism. South and Southeast Asia Culture and Religion, 3, 18–46. Bilimoria, P. (2015). Philosophical orientalism in comparative philosophy of religion: Hegel to Habermas (& Zîzêk). Cultura Oriental, 2, 47–63. Bilimoria, P., & D’Cruz, J. V. (1988). The Bhagavadgita: A perspective of text, tradition and transformation. Journal of Studies in the Bhagavadgita, 5, 51–82. Chakrabarty, D. (2008). Provincializing Europe: Postcolonial thought and historical difference. Princeton University Press. doi:10.1515/9781400828654 Diehl, C. (1978). Americans and German scholarship 1770-1870. Yale University Press. Figueira, D. (2015). The hermeneutics of suspicion: Cross-cultural encounters with India. Bloomsbury Publishing.

98

 The Yoga of the Bhagavad Gita

KEY TERMS AND DEFINITIONS Ātman: The inner Self and also the Supreme Self; the soul. As discussed in the Atharva Veda, it is represented by the letter “A” of the word AUM. AUM (OṂ): AUM is the most sacred syllable symbol in which the whole art and science of meditation is hidden. It is generally pronounced OṂ. It represents the science of being (sat), not-being (asat), and becoming: the singularity (A), the duality (U), and the multiplicity (M). “A” stands for the Self (Ātman), “U” for the non-Self (An-Ātman; Prakṛti) and M, for the relation of negation (Śakti) which exists between them. Buddhi: The spiritual dimension of the psyche where the experience of śraddhā takes place. It is associated with the mind or “manas”. The problem is that the subtle distinction between manas and buddhi is hard to express in the Western theological vocabulary. Guṇas: The triplet of attributes associated with nature. The theory of the three strands is at the core of the establishment of all minor and provisory truths. Error and illusion are relative and necessary steps, both being summed up in the ladder of truth. The three guṇas – sattva, rajas, and tamas – are utterly inseparable though distinguishable. Jīva: The interplay between the Self and the non-Self gives rise to the Jīva and his/her different triads of attributes related to the psychology of cognition, desire, and action. Jīva means a separate self, a living thing, an individual part, so to say, of the Universal Self. Metaphorically speaking, it is like a droplet of the sunshine, passing from the mineral through the vegetable and animal into the human kingdom. Nirvāṇa: Attain nirvāṇa means the experience of the sacred realm hidden in the phenomenological world (saṃsāra). Prakṛti: The non-Self; the natural world. Śakti: The energy that expresses the necessary conjunction of Self and non-Self. It represents the power of life as well as the process of becoming or saṃsāra. Saṃsāra: The phenomenological world of becoming; the natural world. Śraddhā: In the Bhagavad Gītā śraddhā is ground to spirituality, meditation, wisdom, and science. Being itself the truth-force that is common to scientists (atheists or not) and people of different faiths, śraddhā represents a conceptual category that helps us to explain how science evolves shifting from one paradigm to another. The ability to recognize truth and the laws of nature comes from śraddhā, a type of inner strength that shapes one’s views and life.

ENDNOTES

1



2

It means solely the position of one who sees himself or herself and others as one with the Universal Self. It means param-ārtha, or unselfishness, as opposed to svārtha, or selfishness. It is the person with better quality and levels of śraddhā who stays on the path towards truth. In BhG 3.30-1 taken together, Krishna points out that śraddhā comes from the trust and confidence in the universal Ātman. As their śraddhā improves (BhG 7.22), it helps them to liberate themselves from the misunderstanding (moha) about the nature of the Supreme Self (Adhi-Ātman), which comes with birth (sarga) (BhG 7.27-9). Finally, in chapter seventeen, it becomes clear that a truly sattvic śraddhā is the sine qua non condition to find truth. The whole of the chapter is dedicated to the discussion of śraddhā, and the term occurs eight times.

99

 The Yoga of the Bhagavad Gita



3



4



5



6



7



8



9



10



11

100

Emerson’s (1841, 261) essay “Self-Reliance” explores the nature of the aboriginal self [Ātman], the trustee on which a universal reliance [śraddhā] may be grounded. The fact that several interpretations are possible does not mean that the text is not concerned with different and deeper levels of the truth. The Gītā is a constituent of the prasthāna-trayā (threefold canon of the Vedāntins) along with the Upaniṣads and the Vedānta Sūtra. Vedāntic philosophers and theologians have produced many commentaries on the Gītā, according to their school. Among them, we call attention to Śaṅkara, from the Advaita (non-dualist) Vedānta; Rāmānuja, from the Viśiṣṭadvaita (qualified non-dualist); Madhva, from the Dvaita (dualist) Vedānta, and Haṃsa Yogi, from the mysterious and compromising school of Śuddha (pure) Vedānta. Indology provided a method for the application of the positivistic ideals of philology to the texts of non-Western civilizations. As a rule, Ph.D. students working with Indian texts are forbidden to quote the findings of non-scholars. When it comes to the Gītā, they are mostly invited to find a way to disqualify the text as meaningless. To be fair, we must acknowledge also that there are a couple of Indologists, such as Max Müller (who has never traveled to India), that have tried hard to take into account commentators and grammarians from India. His wholehearted approach, however, set him apart from his colleagues Otto von Böhtlingk and William Dwight Whitney. The hermeneutics of suspicion takes for granted that a text is not really about what it claims to be. It appears, initially, more sophisticated than a straightforward reading of the text. However, it lacks a form of self-critique of modernity; for, “the application of a hermeneutics of suspicion to the Bhagavadgītā had the effect of denying it any meaning or value and, above all, any philosophical or ethical dimension” (Adluri and Bagchee 2014, 300). It was mainly used by those scholars who were not interested in the philosophy of the Gītā and opted to ignore the Indian tradition. Turci (2007, pp. v and 26) also managed to avoid the hermeneutics of suspicion on his work due to great help of the thesis committee members Dr. Paul Young and Dr. Graeme MacQueen, who guided the thesis process showing how to stick to the text in its unity, against those willing to apply the “documentary hypothesis” (the idea that the Pentateuch is not the work of one author) to the Gītā. Gandhi’s first encounter with the text of the Gītā happened during the time he lived in London, through a copy of Edwin Arnold’s The Song Celestial. He does not see the Gītā as belonging exclusively to any culture. It belongs to anyone who lives his or her life by it. “Aham”, the “Self”, is the abstract and universal Spirit in all beings, or Jīvas. It is the Iśvara, the One that pervades all. It represents Brahman conjoined with attributes: Saguṇa Brahman. “Aham” is a word with two syllables which correspond to the first and the last letters of the Sanskrit alphabet; “a” and “ha”. Together, they suggest the contents of all possible experiences expressed in Sanskrit. The word “becoming” besides referring to the Jīva, refers also to the totality of the World-Process of Brahman. This is easiest seen if we draw a parallel with the Christian tradition, according to whom Jesus is God (Brahman) and also a full expression of a human being (Jīva). It is worth highlighting how different languages seem to have evolved to account for the process of becoming. In Portuguese, for instance, the verbs “ser” and “estar” are used to express nuances of becoming that the English language does not seem to grasp. So when one says, “Eu sou um filho de Deus” (I am a son of God), he or she is referring to what seems most permanent and belongs to his or her essence (the Self, the Ātman). When one says “Eu estou cansado” (I am tired), one is referring to something that seems to characterize a mode of his or her transient existence (the self, the Jīva). In English, however, we have to use the verb “to be” to refer to both the permanent and

 The Yoga of the Bhagavad Gita



12



13



14



15



16



17



18



19



20



21

the impermanent. Generally speaking, in Sanskrit, verbs correspond to the non-Self, as they express change. Nouns, on the other side, express the Self. Under the verb, we have the first, second, and third persons of the Self. The word Praṇava is a name for the sound AUM. It means, etymologically, “that which renovates, makes new.” The onomatopoeic sound OṂ is full of meaning. It represents the cosmic microwave background noise, the primal sound of Nature. It represents the life-breath of the expanding universe. AUM expresses this sacred truth, assigning to each of its letters a meaning in accordance to the method known as akṣaramudrā, or diagram-seal of letters. “Śakti”, or energy, expresses the necessary conjunction of Self and non-Self. It represents the necessary process of becoming or saṃsāra. From “Śak”; “to be possible”. “Feeling” (bhāva) means the experiencing of being. The continuous recollection of dwelling on a feeling is known as “rasa”. When the feeling is love, the rasa is said to produce peace (śānti). “Feeling” is taken here to be clothed with higher spiritual elements than crude emotions, which are almost irrational and instinctive reactions to some internal or external stimulus. So, joy and love are feelings, while euphoria and passion are emotions. Fear is a feeling; panic, an emotion. As the Gītā explains, when an emotion arises, the result is a loss of spiritual insight. “Dhyāna” comes from “Dhī”, “intelligence,” which is buddhi, the power or faculty of feeling and reflecting on all sides (pari-bhāvana). According to Smith (1979, n. 23, 218) the spiritual dimension of the psyche associated with buddhi is where the experience of śraddhā takes place. The problem is that this subtle distinction between manas and buddhi is hard to express in the Western theological vocabulary. Anyway, śraddhā must, at least, be defined as something related to our inner state. For, as Smith (1979, n. 66, 239) argues the truth (satyam) is dependent upon the heart – and indeed, the heart alone (hṛdaye hy-eva). For a slightly different perspective, see the thorough study on faith presented by Troy A. DuJardin (2019). The Ṛgveda X.113 and in the Śatapatha Brāhmaṇa III.117 depicts śraddhā as the daughter of Sūrya, the sun, a symbol of illumination and revealer of truth. The Ṛgveda X.151 praises śraddhā as the sentiment of conviction and right intention abiding in the heart and responsible for the kindling of Agni (Fire). The Praśna Upaniṣad praises śraddhā as pre-condition to the knowledge of Brahman. The Taittirīya Upaniṣad describes Brahman as having śraddhā as its head, indicating that Brahman is to be attained through śraddhā. The Chāndogya Upaniṣad stresses śraddhā arising out of knowledge. The Bṛhadāraṇyaka Upaniṣad states that śraddhā is placed in the heart. The triplet sat-cit-ānanda is present in every atom. The tree attributes, or guṇas, of people (jñāna or knowledge, icchā or will, and kriyā or activity) and of nature (sattva, rajas, and tamas) spring from this triplet. The guṇas manifest all three at once, but one or other is predominant. The state of sat (being, existence, true, good) is sattva. The oneness of all that is, is sattva. The name of the principle of transformations is rajas. Tamas is that which binds both together. We cannot think of any thesis without also thinking of its antithesis; the two go together. Take the concept of pure consciousness. To begin with, what we understand by pure consciousness is a simple denial of impure consciousness. Otherwise, how could we know at all of the pure, except by at the same time opposing it to the impure? Until the opposition between pure consciousness and impure consciousness is realized the difficulty about consciousness continues. It is all a matter of supposing, opposing, and composing. No wonder why many passages in Hegel’s The Phenomenology of Spirit (1807) read almost like translations from the Vedānta. 101

 The Yoga of the Bhagavad Gita



22



23



24



25 26

102

From “sa” (with) and “akṣa” (sense, eyes, and also the center of a wheel). When a wheel turns, its center is still. “Sākṣin”, then, expresses the capacity to remain steady while events are turning around. It refers to the ultimate Observer, or Witness behind the sense of individuality, or the Self, which resides in the heart. It means that “Pure Awareness” that witnesses the world but does not get affected or involved. It indicates the experience that we are not the body, but rather the witness of the body and of the whole creation. It indicates experience with the Iśvara, the sole Selfconsciousness (cetā) who gives consciousness to every human being and is the witness of all. There are also five of Bindu Upaniṣads that deal with the practice of meditation with AUM to apprehend Ātman (soul, self). Among them, the most known are the Dhyānabindu Upaniṣad, which asserts that OṂ is a means to meditation, to understanding Ātman and the Brahman (ultimate reality); and the Tejobindu Upaniṣad that has its focus completely on meditation in accordance to the nature of sat-cit-ānanda of the Self and the non-Self. “Bindu” means the point (anusvāra, the symbol transliterated as “ṃ”) representing the “M” of the AUM; the power of Brahman. Furthermore, it is well-known among yogis that “ham-sa” is only “sah-aham,” or “so ‘ham,” reversed. Hence the Haṃsa is often identified with the Supreme Spirit and also with those who possess the sacred knowledge of Brahman. The whole manifestation of the Trinity which constitutes the one Brahman is saṃsāra. According to BhG 4.39-40, śraddhā comes from placing Jīva under the jurisdiction of Ātman; it means the awakening of Ātman. For, the ignorant is he whose Ātman is sleeping – a fact that manifests itself as one being dispossessed of śraddhā. Krishna is using in BhG 4.40 the same metaphor he had used in BhG 2.69: the sage is awake to Ātman, which is the night of the other beings.

103

Chapter 7

Molecular Mechanisms Underlying the Effects of Yoga Surabhi Gautam All India Institute of Medical Sciences, New Delhi, India Rima Dada All India Institute of Medical Sciences, New Delhi, India

ABSTRACT Complex chronic lifestyle disorders are the leading causes of death and disability worldwide. Stress and anxiety associated with today’s hectic life schedule and polluted environment have contributed a lot in triggering and causing many chronic diseases and decreased quality of life, even with pharmacologic treatment. Most of the chronic complex diseases, such as chronic obstructive pulmonary disease, depression, autoimmune diseases, cancer, cardiovascular diseases, obesity, and diabetes mellitus share underlying mechanisms like high levels of stress, anxiety, depression, oxidative stress, shorter telomeres, persistent activation of hypothalamo-pituitary adrenal axis, inflammation and dysregulated immune system, and thus need to be managed by an integrated approach that targets both mind and body. The individuals with these conditions have been reported to benefit from yoga, but the underlying mechanism of action of yoga remains unclear. The aim of this chapter is to summarize the mechanism of action underlying the cumulative effect of yoga on multiple pathways at a cellular level.

INTRODUCTION Yoga is a mind body intervention of Indian origin which brings balance to all dimensions of health viz physical, mental, emotional, and spiritual. It aims to bring about the synchrony of physical and mental health by prevention of disease, maintenance of health and attaining peace of mind. The literal meaning of the Sanskrit word “yoga” is “union” (Woodyard, 2011). According to the proponent of Ashtanga yoga “Maharishi Patanjali”, yoga is defined as the discipline to develop one’s inherent power in a balanced manner and subduing modifications of one’s mind. Yoga is an integrative science which regulates various organ systems of the body and offers the means to attain complete self-realization (Gard et al., DOI: 10.4018/978-1-7998-3254-6.ch007

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Molecular Mechanisms Underlying the Effects of Yoga

2014). The most commonly used components of yoga therapy for health benefits are asanas (physical postures), pranayama (regulated breathing) and dhyana (meditation). The features like breath regulation, mindfulness and maintenance of postures are some of the elements which are responsible for the activation of parasympathetic nervous system and hence differentiate yoga practices from physical exercises (Ross & Thomas, 2010). Several studies cited the positive effects of mind body interventions (MBIs) improving patients’ overall well-being and quality of life, reducing their psychological distress, improving sleep patterns and creating a positive mental state and decreasing their pain intensity (Fernros et al., 2008; Gautam et al., 2019; Telles & Singh, 2012; Tolahunase et al., 2018; Tolahunase et al., 2017; Zautra et al., 2008). Yoga is a complete package of mindfulness-based stress reduction and relaxation techniques which acts via a well-defined psycho-neuroendocrine pathway. Various stressors reduce the immune efficiency and fail to protect the health of all vital systems. The cortical region of CNS is the seat of awareness, consciousness and all voluntary psycho-physiological functions including control of breath, which results in modulation of neuro-humoral response as well as monitors the homeostasis of all the physiological systems for optimum health. However, ill-effects of stress causing the cortical dysfunction may jeopardize the overall homeostasis leading to various disorders. The psychophysiological aspects of stress response arising from the brain functions are supposed to reflect on the functions of hypothalamic-pituitary-adrenal (HPA) axis. The HPA axis is hypothalamic, pituitary, adrenal axis, which consist of a central stress response system. Since most of the motor commands are routed through HPA axis and other endocrine functions, the motor commands are propagated via brain stem and reticular activating system to HPA axis. Thus, the stress related effects of HPA working in unconscious mode might modulate neuro-endocrine axis that in turn affects immune system. HPA hyperactivity predisposes to increased cellular oxidative stress (OS). Yoga leads to an inhibition of the posterior or sympathetic area of the hypothalamus, thus, optimizing the body’s sympathetic responses to stressful stimuli, and restores autonomic regulatory reflex mechanisms associated with stress. It is also well known that the hypothalamus and the limbic system are intimately concerned with emotional expressions. Yoga practices inhibit the areas responsible for fear, aggressiveness, and rage, and stimulate the rewarding centers in the median forebrain and other areas, leading to a state of bliss and pleasure (Woodyard, 2011). This results in lower anxiety, heart rate, respiratory rate, blood pressure, and cardiac output in individuals practicing yoga and meditation. Yoga practices probably inhibit the activity of the paraventricular nuclei of the hypothalamus, which in turn affects the anterior pituitary gland to produce less ACTH (Arora & Bhattacharjee, 2008). The decrease in ACTH decreases the synthesis of cortisol from the adrenal glands. Yoga possesses a buffering action on stress mediated immune responses, maintains cardio-vagal tone, and maintains a homeostasis between parasympathetic and sympatho-adrenal nervous system and downregulates HPA axis response (Arora & Bhattacharjee, 2008; Kanojia et al., 2013). The growing popularity of mind body therapies (MBTs) including yoga based lifestyle intervention (YBLI), mindfulness based stress reduction (MBSR), Tai-chi, Qigong etc. have emphasized the need to discover the molecular mechanisms behind their therapeutic actions. A meta-analysis on the effect of mind body therapies on the immune system concluded that yoga has anti-inflammatory and immunomodulatory potential (Morgan et al., 2014). Short term yoga and related practices result in rapid global gene expression profile changes in the peripheral blood mononuclear cells (PBMCs) which may be the basis for their long term cell biological and higher level health effects (Qu et al., 2013). A review of clinical effectiveness of mindfulness training and yoga for the management of chronic non-malignant pain provided the evidence of improved pain acceptance and depression scores (Lachance & McCormack, 2019). Recent studies from our laboratory that focussed on various complex chronic lifestyle disorders 104

 Molecular Mechanisms Underlying the Effects of Yoga

like unexplained male factor infertility, idiopathic recurrent pregnancy losses, depression and glaucoma found that YBLI may have important roles in slowing the rate of cellular aging and improving cellular longevity thus promoting health and lifespan (Dada et al., 2018; Dhawan et al., 2017, 2018; Gagrani et al., 2018; Gautam et al., 2019). Studies from our laboratory on unexplained infertility suggested that simple lifestyle interventions like yoga and meditation can reduce testicular inflammation by optimizing OS markers thereby reducing oxidative DNA damage and therefore reduce the incidence of infertility and future consequences in the next generations. This is due to the impact of yoga on sperm genome and epigenome (Bisht et al., 2017; Bisht & Dada, 2017; Gautam et al., 2018). Studies from our lab also suggested an increase in levels of 𝛽-endorphin, brain derived neurotrophic factor (BDNF), dehydroepiandrosterone (DHEA), melatonin and sirtuin-1 and decrease in levels of cortisol and Interleukin (IL)-6 indicating improvement in stress and inflammatory response after YBLI with regulatory changes in brain through hypothalamic-pituitary-adrenal (HPA) axis in patients with major depressive disorder (Tolahunase et al., 2017; Tolahunase et al., 2018). This was the first study to show that yoga increases levels of sirtuins (a family of proteins that regulate cellular health) independent of intake of resveratrol or caloric restriction. A recent study from our laboratory on active rheumatoid arthritis (RA) patients has shown yoga to reduce pain perception, functional disability and disease severity followed by 8 weeks of YBLI. This study also documented a post yoga upregulation in levels of an immune-modulatory molecule, soluble Human leukocyte antigen (sHLA-G) and a decrease in levels of IL-6 (extra-articular symptoms) and IL-17A and TNF-α (articular symptoms) (Gautam et al., 2019). Thus yoga may aid in achieving immunological tolerance and molecular remission and hence can be beneficial as an adjunct therapy in management of this severe chronic progressive autoimmune arthritis. These studies from our laboratory served as the evidence base which proved the beneficial & therapeutic roles of YBLI in healthy and diseased individuals, further insights may help us providing the exact underlying mechanism of action.

YOGA: A WAY TO HEALTHY LIFE Though aging is inevitable one can definitely age gracefully and prolong youthful life by adopting a yoga based mind body intervention. Cellular longevity and health are dependent on the genomic stability as suggested by findings from our laboratory stating the decreased levels of 8-OHdG (8-oxo-2’-deoxyguanosine) followed by YBLI. Elevated reactive oxygen species (ROS) levels cause oxidative DNA damage. As a consequence of DNA damage, there is accumulation of mutagenic oxidized bases i.e. 8-OHdG due to deficient DNA repairs mechanism. Genomic instability results from a myriad of poor lifestyle factors causing DNA damage to both the mitochondrial and nuclear genome resulting in accumulation of genetic aberrations and genome hyper mutability (Bisht & Dada, 2017; Gautam et al., 2019). This is mainly due to aberrant DNA damage repair mainly responsible for monitoring genomic integrity. Deficient DNA damage response triggers systemic inflammatory conditions which promote pathological aging (Tolahunase et al., 2017). YBLI activates DNA damage repair pathway as suggested by reduction of 8-OHdG levels and upregulation of the expression of DNA repair genes improve genomic stability. The details of activities in a single session of yoga based lifestyle intervention program are tabulated in Table 1. Yogic sukshmavyayama (ancient technique of yogic postures associated with a specific type of breathing used for warming up) and sthoolvyama (ancient technique of yogic dynamic movements associated with a specific type of breathing used for warming up) help to sequentially warm up all joints of body with a special focus on small joints of hand and feet and helps to maintain balance 105

 Molecular Mechanisms Underlying the Effects of Yoga

and co-ordination respectively. This system has a strong purifying effect on energy body of humans (Telles & Singh, 2013). Yogāsanas (postures) were taught to maintain in different positions like sitting, standing, supine and prone. Trikoṇāsana (the triangle pose) maintains flexibility of spine strengthens the calf, thigh and waist muscles and improves lungs capacity (Booklets on Common Yoga Protocol, n.d.). Katichakrasana (spinal twists pose) helps to improve the waist and spinal mobility and stretching of different group of abdominal muscles. Tadasana (palm tree posture) stability in the body helps to clear up congestion of the spinal nerves and corrects faulty posture (Booklets on Common Yoga Protocol, n.d.; Rathore et al., 2017; Telles & Singh, 2013). Veerbhadrasana (warrior pose) helps in stretching and strengthening of the pectoral, shoulder, neck, pelvic, psoas muscles. It also strengthens and stretches the thighs, calves and ankles and leads to the stretching of the entire front of the body while strengthening the thighs, ankles and back. Gomukhasana (cow face pose) stretching of gluteal muscles, relieves chronic knee pain, strengthens spinal and abdominal muscles (Booklets on Common Yoga Protocol, n.d.; Rathore et al., 2017). Paschimottanasana (seated forward bend) stretches the back extensor muscles, stretches the body from head to ankle and contracts the anterior abdominal wall muscles. It also creates pressure on the thorax and abdomen, improving the process of respiration and the functioning of the intra-abdominal glands, especially the secretions and provides a soothing effect on the mind, removing anxiety, anger and irritability (Booklets on Common Yoga Protocol, n.d.; Rathore et al., 2017). Shashankasana or the “hare pose” induces relaxation, relieves depression and provides a good upper body stretch. Vakrasana (simple spinal twist) massages the abdominal organs, helps facilitate digestion, regulates the secretion of digestive juices and reduces constipation and stomach diseases (Booklets on Common Yoga Protocol, n.d.; Telles & Singh, 2012, 2013). Ekpadashalabhasana (single leg locust pose) and poornashalabhasana (complete locust pose) improves flexibility and coordination, exercises the back muscles, and increases strength and stamina (Booklets on Common Yoga Protocol, n.d.; Telles & Singh, 2013). Bhujangasana (cobra pose) relieves stress, reduces abdominal fat and relieves constipation and helps to relieve backache and bronchial problems (Booklets on Common Yoga Protocol, n.d.). Makarasana (crocodile pose) promotes relaxation of the whole body, helps in recovery of back problems and counterbalances stress and anxiety levels (Booklets on Common Yoga Protocol, n.d.; Swanson, 2019). Uttanpadasana (raised leg pose) strengthens the abdominal muscles, tones all the abdominal organs and improves functioning of reproductive organs. Setubandhasana (bridge pose) helps relieves depression, anxiety and strengthens lower back muscles and stretches abdominal organs, improves digestion and helps to relieve constipation (Booklets on Common Yoga Protocol, n.d.; Telles & Singh, 2013). Pawanmukatasana (wind liberating pose) removes constipation, gives relief from flatulence, decreases the bloating sensation in the abdomen, aids digestion, generates deep internal pressure, and improves stretching of muscles, ligaments and tendons in the pelvis and waist region and tones up the back muscles and spinal nerves (Booklets on Common Yoga Protocol, n.d.; Telles & Singh, 2013). Matsyasana (fish pose) stimulates spinal muscles, cervical muscles, thorax, rib cage, and the lungs, thus relieving your fatigue almost instantly. Pranayama is a practice of breath control and helps in developing awareness of one’s mind and helps to establish control over the mind (Booklets on Common Yoga Protocol, n.d.; Mishra, 2012). Cellular longevity and telomere metabolism is maintained by increase in telomerase activity which is responsible to check the attrition of telomere length as shown by our study after YBLI. Telomeres serve as a biological clock and are vital for cellular longevity. Rapid telomere attrition can be attributed to oxidative DNA damage which is associated with senescence and related disease conditions (Shammas, 2011). A recent study from our lab suggests that there is improvement in maintenance of balance in cellular OS (decrease in ROS and increase in TAC) by YBLI. ROS levels 106

 Molecular Mechanisms Underlying the Effects of Yoga

below physiological limits are deleterious to normal cellular function and maintenance of OS at physiological levels is important for cellular longevity (Tolahunase et al., 2017). Increased OS causes damage to DNA and telomeres, affects signal transduction and gene transcription by causing genome wide hypomethylation and thus causes changes in the epigenome. DNA methylation regulates gene expression by recruiting proteins involved in gene repression or by inhibiting the binding of transcription factors to DNA. YBLI holds the potential in protecting cells from OS induced DNA damage and telomere attrition and in reversing epigenetic changes, which are accumulated due to unhealthy lifestyle and adverse environmental conditions (Dada et al., 2018; Gagrani et al., 2018; Gautam et al., 2018). Psychological stress triggers various pathological conditions like neuropsychiatric disorders including depression, psychosomatic disorders contributing to increased prevalence of metabolic syndromes including, obesity, DM (diabetes mellitus), and CVD (cardio vascular disease) (Nunn et al., 2009). The triggering and propagation of chronic diseases can be curbed with the adoption of a simple lifestyle based intervention like yoga which aims to rejuvenate mind by counteracting the stress response, relaxes the body by maintaining flexibility and improving joint mobility and regresses inflammatory state by regulating OS induced cellular damage and maintains an equilibrium of pro-inflammatory & anti-inflammatory cytokines. It also normalizes the levels of dysregulated transcripts especially of genes regulating free radical levels, cell cycle repair, downregulation of apoptosis and maintaining DNA integrity (Dada et al., 2018; Dhawan et al., 2018; Gautam et al., 2018). Yoga helps in the maintenance of cellular longevity and health by optimizing telomere attrition by upregulation of telomerase enzyme activity and genes responsible for DNA repair and cell-cycle control. Most of the pharmacologic treatments fail to cure the psychological component of the disease; hence yoga as an integrative medicine can be used as an adjunct ensuring overall health and wellness. Yoga is one of the commonest mind body intervention practiced worldwide and is beneficial as is inexpensive, safe and has minimal side-effects. Yoga is a scientific technique which aids in maintaining correct posture, increasing joint range of motion, flexibility, strength, coordination and muscular tone (Dada et al., 2018; Gagrani et al., 2018; Geneen et al., 2017; Mazor et al., 2018; Tolahunase et al., 2018). Several studies have shown the positive aspects of yoga in healthy people for improving wellness as well as in patients with chronic diseases by improving their quality of life and coping mechanism (Dada et al., 2018; Jayadevappa et al., 2007; Khalsa et al., 2012; Sharma et al., 2006). Several psycho-social and health-restorative benefits of yoga & meditation have been investigated, but their impact on the cellular health and associated molecular mechanism remains poorly defined. An unhealthy modern lifestyle, environmental factors and stress affect the reproductive system leading to aging of the gonads leading to infertility and recurrent pregnancy loss. Social habits of the male partner, such as nicotine consumption (oral / smoking) and excessive alcohol intake, may lead to systemic and seminal OS, and hence elevated reactive oxygen species (ROS) levels cause oxidative DNA damage to sperm. A high mutagenic load may overburden the oocyte repair mechanism post-fertilization and result in persistence of DNA damage/mutation in the zygote genome. A study from our lab has shown that YBLI can reduce testicular aging and results in significant changes of epigenome and transcriptome to maintain the oxidative eustress (Rima et al., 2016). Improvement in the stress and inflammatory response in our study after YBLI may be mediated by changes in cortisol, 𝛽-endorphin, IL-6, and other factors, with regulation by changes in brain through the HPA axis. Yoga is mind body intervention which regulates ROS rather than simply lowering them by balanced stress-related processes, reducing the rate of cellular aging and appropriate gene expressions and helps in the prevention of chronic diseases associated with stress (Dhawan et al., 2017; Tolahunase et al., 2017; Tolahunase et al., 2018). Though one 107

 Molecular Mechanisms Underlying the Effects of Yoga

cannot change the chronology of their age, but can definitely slow down the pace of aging by adopting YBLI. The biomarkers of cellular aging showed a significant decline after YBLI, and reduced the risk of chronic lifestyle diseases like depression.

YOGA AND MANAGEMENT OF CHRONIC COMPLEX DISEASES Yoga offers many benefits for the management of chronic diseases. Chronic lifestyle disorders have been attributed to endogenous and exogenous risk factors like poor nutrition, physical inactivity, tobacco use, excessive alcohol consumption, smoking etc. Chronic diseases are associated with the activated sympathetic axes and parasympathetic withdrawal leading to an inflammatory overload in the body hence increasing the levels of cortisol, glucagon, prolactin, and various other pro-inflammatory cytokines. There is a dysregulation of HPA axis and sympathetic adrenal medullary (SAM) system in lifestyle related diseases in response to chronic stress exposure. The stressors activate the brain network including amygdala, hippocampus and pre-frontal cortex, which in turn secrete glucocorticoids, catecholamines and inflammatory cytokines. This response to stress is adaptive and short term in nature and causes the wear and tear within HPA axis and SAM system which contributes to the development lifestyle associated stress disorders. In diabetes mellitus, the psychological stress is associated with insulin resistance, hypertension, and cardiovascular event sequelae (Pullen et al., 2018). Yoga practice results in the reductions in stress, depression and anxiety and an overall improvement in quality of life of an individual. Yoga is beneficial in reducing the chronic stress as it turns off the HPA-SAM response leading to reduced levels of cortisol, norepinephrine, epinephrine and pro-inflammatory cytokines. Various postures during a yoga session especially including the abdominal stretching are believed to result in the regeneration of pancreatic cells (Singh et al., 2015). Yoga aids in improving the glycemic control in people with diabetes and helps to develop sensitivity of β-cells to glucose, thereby improving insulin secretion, and increase the blood supply to the muscle and muscle relaxation, thereby improving glucose uptake. Yoga reduces the levels of free fatty acids and improves insulin kinetics by reducing fasting insulin levels, by normalizing the insulin-to-glucose ratio (Krishna et al., 2014). Yoga optimizes OS, as demonstrated by reductions in ROS levels, serum malondialdehyde and leptin levels, and improvements in TAC and adiponectin levels (Sinha et al., 2007). Yoga also results in reduction of fasting and postprandial blood sugar, haemoglobin A1c, and anti-diabetic drug requirements, suggesting an improved glycaemic control. Yoga practice improves the cardiac autonomic functions and reduces the risk of cardiovascular events associated with diabetes (Chu et al., 2016; Pullen et al., 2018). Yoga effectively works via three mechanisms: vagal nerve stimulation, reduction in perceived stress, and musculoskeletal stimulation. Yoga may promote parasympathetic activation by stimulating the vagal nerve thus leading to decreased heart rate, blood pressure, improved heartrate variability, and similar metabolic and psychological benefits resulting in an improved outcome (Cramer et al., 2015). Yoga practice leads to the upregulation of various hormones like β-endorphin, serotonin, and dopamine levels which leads to the feelings of joy and euphoria; improvements in arousal are due to increased argininevasopressin levels, which reduce the gamma-aminobutyric acid (GABA)-ergic inhibition of the supra-optic area of the hypothalamus. Yoga provides a soothing and calming effect due to melatonin whereas the ecstatic and blissful feelings that arise during yoga are due to lateral hypothalamic stimulation, decrease in spatial orientation and anti-anxiety effects are due to increased levels of GABA (Bhasin et al., 2013).

108

 Molecular Mechanisms Underlying the Effects of Yoga

Diseases like chronic obstructive pulmonary disease (COPD) characterized by persistent blockage of airflow and symptoms such as breathlessness, abnormal sputum, and chronic cough have shown a reduced dyspnea-related distress significantly post 12 weeks of Iyenger yoga (Santana et al., 2013). Breast cancer patients showed an improvement in physical functioning based on the physical component scale scores after a 6-week yoga program. The quality of life and stress symptoms showed greater improvements in cancer affected patients when subjected to mindfulness meditation and gentle yoga (Woodyard, 2011). A short-term yoga-based lifestyle intervention may be an important modality to reduce the risk for cardiovascular disease as indicated by a decrease in IL-6 and an increase in adiponectin in overweight and obese men (Stephens, 2017). Yoga therapy also maintains immune-homeostasis by optimizing the inflammatory and anti-inflammatory cytokine release. A study from our laboratory has shown the positive effects of 8 weeks of YBLI in RA patients as it showed a downregulation in pro-inflammatory cytokines which are known to modulate its pathogenesis (Gautam et al., 2019). Various studies reported that yoga reduced stress and inflammation makers in patients with chronic inflammatory diseases who participated in a short-term yoga-based lifestyle program (Gautam et al., 2018; Rima et al., 2016; Tolahunase et al., 2018; Yadav et al., 2012). Yoga intervention maintains the dysregulated T helper 17 (Th17) and T regulatory (Treg) cell populations estimated by flow cytometry analysis, also reverses the aberrant global methylation patterns in RA patients documented in a study from our lab on RA patients (Gautam, Kumar, et al., 2019; Gautam, Tolahunase, et al., 2019). Therefore, regular yoga exercise might reduce the levels of pro-inflammatory cytokines and could be an effective treatment for chronic diseases. Yoga has also been shown to improve levels of BDNF, a marker of neuroplasticity, in major depressive disorder patients. A recent study further confirmed that in depressive patients when subjected to 12 weeks of yoga intervention could significantly increase systemic BDNF levels, in association with improvement in biomarkers of mind-body communication and cellular health (Tolahunase et al., 2018). The immune system is an essential component of the physiological stress-sensing pathways and closely interacts with the body’s primary integrative mind body communicating systems. A previous study has documented the neuro-immunological mechanisms of peripheral immune dysfunction in depression patients, and has shown that these routes converge in the brain to alter neurogenesis and neuroplasticity associated molecular signaling mechanisms (Hodes et al., 2015). A systematic review of gene expression changes indicated that mind body interventions like yoga and meditation are linked to downregulation of nuclear factor kappa B (NFκB) pathway and an upregulation of TGF-β pathway (Buric et al., 2017). Meditation results in a decrease in inflammation as evident from the reduction in levels of inflammatory marker IL-6 and increases neuroplasticity. The findings from one of our studies indicated a decrease in IL-6 after YBLI suggests that MBI practices may lead to a reduced risk of inflammation-related neuropathology. The patients of major depressive disorder receiving YBLI showed increased DHEAS (dehydroepiandrosterone sulfate) levels. Increased DHEAS after YBLI may also decrease depression severity by decreasing heightened brain excitability of major depressive disorder, since DHEAS can inhibit glutamate-induced excitotoxicity and modulate the inhibitory neurotransmitter GABA (Genud et al., 2009). Sirtuin 1 (SIRT1), a member of mammalian class III histone deacetylases, promotes longevity, nutrition and energy sensing pathways. A significant upregulation of SIRT1 after 8 weeks of YBLI indicated preservation of the dysregulated cellular health span and longevity. In major depressive disorder patients, 12 weeks of YBLI could improve mind body connections, contribute to enhanced neuroplasticity and decrease in depression severity scores. Recent evidence has confirmed that yoga and meditation could re-establish oxidative eustress in apparently healthy people by increasing total antioxidant capacity, re109

 Molecular Mechanisms Underlying the Effects of Yoga

ducing inflammation and improvement in mitochondrial DNA integrity (Tolahunase et al., 2018). The pathophysiological relationships between OS and depression, and the potential benefits of both passive and active interventions to achieve optimum oxidative eustress deserve further research. Research evidence suggests that regular practitioners of yoga have increased telomere length compared to controls, and yoga practice by apparently healthy individuals can increase telomere metabolism and promote longevity by slowing the rate of aging (Krishna et al., 2014; Tolahunase et al., 2017; Tolahunase et al., 2018). The main factor that causes rapid attrition of telomeres is OS, and YBLI reduces OS and inflammation and thereby aids in telomere length maintenance and promotes genome stability and maintenance of chromosomal integrity. Previous studies document improvement in mitochondrial and nuclear DNA integrity and improvement in COX activity after YBLI, aids in meeting the energy demands of a tissue with a high metabolic rate (Dai et al., 2014). A study on the impact of YBLI on RA patients indicate suppression of the acute phase responses, reduction of inflammatory markers and suppression of hyperactive immune system by – downregulation of acute phase reactants (ESR and CRP) and pro-inflammatory cytokines (TNF-α, IL-6 and IL-17A), co-stimulation of anti-inflammatory cytokines (TGF-β) and immune-modulatory molecule (sHLA-G). Yoga not only relaxes the body and mind but also maintains homeostasis at internal organ system level and regulation of metabolic processes. The pleiotropic cytokine transforming growth factor (TGF)-β and fork head box P3 (Foxp3), an X-chromosome-linked transcription factor expressing Tregs cells helps in active immunosuppression, a function that is critical to maintain immune homeostasis and prevent autoimmunity. Yoga helps in downregulation of inflammatory processes by reduced pro-inflammatory cytokines and increase anti-inflammatory cytokines (Rajbhoj et al., 2015; Shete et al., 2017). In our laboratory, we have also found an upregulation of sHLAG levels in RA patients followed by 8 weeks of YBLI. HLA-G, a non-classical HLA class I molecule, possesses immune-modulatory and anti-inflammatory properties. It is a good reference parameter for prevention, diagnosis and a potential treatment target in autoimmune and inflammatory diseases since its elevated levels are associated with reduced severity and fewer relapses in RA patients. Our results showed that yoga aids in maintenance of OS as indicated by a significant reduction in ROS levels and elevation of TAC levels in yoga group as compared to control group. 8-hydroxy-2’-deoxyguanosine (8OHdG), major product of DNA oxidation, leads to DNA breaks, base modifications, DNA mutations, genomic instability in nuclear and mitochondrial DNA (Dizdaroglu, 2012). Maintaining optimum oxidative eustress, even under extremes of stress associated with lifestyle and environmental challenges, is a highly challenging task and needs a regulated redressal. Previous studies from our laboratory have documented that yoga based lifestyle intervention improves mitochondrial integrity as evident from increased COX (cyclooxygenase) activity, up-regulating the total anti-oxidant capacity and telomerase enzymes in healthy population, thus reducing and delaying onset of age-related chronic diseases and complex lifestyle disorders and their sequelae (Bisht et al., 2017; Bisht & Dada, 2017; Gautam et al., 2018; Rima et al., 2016). Long-term relaxation response practitioners were found to have upregulated pathways associated with genomic stability such as telomere packaging, telomere maintenance and tight junction interaction whereas downregulated pathways that were associated with the inflammatory responses (Bhasin et al., 2013). Yoga normalizes the levels of transcripts involved in DNA repair, cell cycle checkpoint control and causes down regulation of pro-inflammatory genes. Therefore, adoption of yoga as a lifestyle intervention is beneficial in reducing the severity of psychosomatic symptoms and sequelae of autoimmune disease by its immune-modulatory potential. Thus yoga possesses a curative and rehabilitative potential and may be used as an adjunct therapy.

110

 Molecular Mechanisms Underlying the Effects of Yoga

YOGA OPENS THE DOOR TO SELF-ACTUALIZATION: MOLECULAR MECHANISMS Yoga has the potential to prevent psychosomatic disorders and provides ability to endure stressful situations. Various types of yoga include japa yoga (to concentrate one’s mind on divine name or mantra through repeated recitation or remembrance), karma yoga (to consider yoga as divine action and perform it with whole-hearted dedication and shuns away all desires), gyana yoga (to discriminate between self and non-self and to acquire the knowledge of one’s spiritual entity through the study of scriptures, company of Saints and practices of meditation), bhakti yoga (spiritual practice focused on loving devotion towards a personal god), raja yoga (to achieve the path of self-discipline and practice; also known as ashtanga yoga), swara yoga (realization of cosmic consciousness, through the awareness or observation, then control or manipulate the flow of breath in the nostrils), kundalini (energy that lies dormant at the base of the spine until it is activated and channelized upward through the chakras in the process of spiritual perfection) and nadi (to visualize the flow of energy at a psychic level as having distinct channels, light, colour, sound, and other characteristics). Each yoga school follow its own principles and practices leading to fulfillment of ultimate goal i.e. attainment of self-actualization. Yoga is essentially a spiritual discipline which leads to the union of individual consciousness with that of the universe, bringing a perfect harmony between mind and body (Arora & Bhattacharjee, 2008; Gard et al., 2014; Khalsa et al., 2012; Morgan et al., 2014). The practice of yoga includes a set of asanas (physical postures) associated with pranayama (synchronous deep breathing) followed by dhyana (meditation) which aids to increase blood flow to the muscles, hence improves tissue oxygenation. The maintenance of physical posture keeps the muscles in isometric contraction and improves muscle strength, provides balance and equilibrium and maintains joint mobility by keeping the muscles in isometric contraction (Arora & Bhattacharjee, 2008; Woodyard, 2011). According to yogic scriptures, the practice of yoga leads to the union of individual consciousness with universal consciousness. The aim of yoga practice (sādhana) is to overcome all kinds of sufferings that lead to a sense of freedom in every walk of life with holistic health, happiness and harmony. Yoga possesses a buffering action on stress mediated immune responses, maintains cardio-vagal tone, and maintains a homeostasis between parasympathetic and sympatho-adrenal nervous system and downregulates HPA axis response (Arora & Bhattacharjee, 2008; Innes et al., 2005; McInnes & Schett, 2011; Pilkington et al., 2005; Sengupta, 2012). A short term program of meditation practice has led to elicitation of relaxation response in healthy subjects and showed a down regulation of transcripts associated with regulation of apoptosis, nuclear transport, nuclear factor regulation, metabolic processes, JAK-STAT (Janus kinase-signal transducer and activator of transcription) cascade, T and B cell activation, regulation of cell cycle, insulin sensitivity, glucose transport, DNA replication, chemokine signaling and stress response (Bhasin et al., 2013). The proposed mechanism of action of yoga is shown in Figure 1. Yoga facilitates bidirectional feedback and improve integration and efficiency of high-level (e.g., central executive network, frontal–parietal control network) and low-level brain networks (e.g., autonomic systems, vagal complex, striatopallidal–thalamo cortical network) along with viscerosomatic, musculoskeletal, cardiac, respiratory, and sensory information coming from the periphery. Yoga intervention also upregulates expression of DNA repair genes and anti-inflammatory cytokines such as IL-2 and IL-4 and decreases the levels of pro-inflammatory markers such as IL-6, TNF-α and oxidative DNA damage marker, 8-OHdG (Dada et al., 2018; Tolahunase et al., 2017). Our recent study on active RA patients documented the suppression of acute phase responses by lowering the levels of ESR and CRP, 111

 Molecular Mechanisms Underlying the Effects of Yoga

alteration of inflammatory markers and suppression of hyperactive immune system by downregulation of pro-inflammatory cytokines (IL-6, IL-17A and TNF-α), whereas upregulation of anti-inflammatory cytokines (TGF-β) followed by an 8 weeks YBLI (Gautam, Tolahunase, et al., 2019). Yoga normalizes the levels of transcripts involved in DNA repair, DNA damage, genome stability, cell cycle checkpoint control and causes down regulation of pro-inflammatory genes (Dhawan et al., 2017, 2018; Gautam, Tolahunase, et al., 2019; Tolahunase et al., 2018). The transcriptome analysis by microarray studies on effect of YBLI from our lab showed a downregulation in expression levels of various pro-inflammatory cytokine and chemokine receptor-ligand transcripts an upregulation in expression levels of various antiinflammatory, cell cycle control, DNA repair genes (Bortolotti et al., 2014). Therefore, yoga as a lifestyle intervention possesses an immune-modulatory potential which maintains immunological tolerance and remission at a molecular level. Yoga helps in maintaining cellular health by redox homeostasis regulation, avoiding genomic damage, and telomere length maintenance by upregulation of telomerase. Yoga also increases plasma melatonin levels hence maintains sleep-wake cycles and also aids in upregulation of SIRT1 which promote longevity by deactivation of p53 mediated pathways. RA patients undergo premature immune aging as they have insufficient amounts of telomerase reverse transcriptase enzyme and have shorter telomeres (Costenbader et al., 2011; Fujii et al., 2009). There is an upregulation in telomerase enzyme levels with 8 weeks YBLI which maintains the pace of cellular aging and telomere length in RA patients. Hence, yoga not only delays the premature early aging of immune system in RA but also reduces the disease activity, severity and co-morbid functional disability by maintenance of cellular health and longevity (Gautam, Tolahunase, et al., 2019). Other clinical trials suggested that yoga reduces symptoms of anxiety and depression and reduces inflammation, maintains a balance of autonomic nervous system tone, and increases in vagal activity (Bernardi et al., 2001; Pilkington et al., 2005; Riley, 2004; Shete et al., 2017; Vijayaraghava et al., 2015; Woolery et al., 2004). The deficiency of neurotransmitters and neuroprotective substances like dopamine, serotonin and norepinephrine is associated with anxiety, social phobia and depression (Kim & Gorman, 2005; Schneier et al., 2000; Selden et al., 1990). Studies suggested that yoga is found to significantly elevate the markers associated with neuroplasticity like BDNF, serotonin and β-endorphins post 8 week YBLI along with the reduction of depression measuring scale i.e. BDI-II (Brown & Gerbarg, 2005; Kinser et al., 2012; Lee et al., 2014; Yadav et al., 2012). Yoga focusses on both psychological and physical aspect of the disease and helps to maintain an immune homeostasis between various T cell subset populations especially Th17 and Tregs. Increase in levels of BDNF, DHEA, are both cyto protective and promote synaptogenesis and are the underlying mechanism of increase in hippocampal, prefrontal cortex and cerebellar volume. Yoga acts via psycho-neuro-immunological axis which creates a homeostatic balance between sympathetic and parasympathetic limbs of the autonomic nervous system during aggressive symptomatic phase and helps to normalize the flare and achieve remission (Gard et al., 2014; Kim & Gorman, 2005). Yoga helps to improve overall quality of life by reducing pain perception, disability quotient, and disease activity in active RA patients (Fernros et al., 2008; Gautam, Tolahunase, et al., 2019). Yoga is a cost effective emerging health discipline which unlike drugs have no side-effects and aid in prolonging the periods of remissions with fewer relapses. Yoga not only reduces disease severity, optimize OS levels, increases periods of remission, but also minimize usage of drugs with minimum side effects hence can be beneficial as an adjunct therapy. Over the last two decades, yoga has received increasing awareness and attention from the scientific community seeking to understand the safety and efficacy of these widely used practices (Morgan et al., 2014). A large number of studies suggest mindfulness-based interventions including mindfulness based 112

 Molecular Mechanisms Underlying the Effects of Yoga

stress reduction, mindfulness-based cognitive therapy and integrated mindfulness yoga practices are effective psychological interventions to reduce depression and anxiety in clinical and non-clinical populations (Goldin & Gross, 2010; Spijkerman et al., 2016). Mindfulness can be described as a process of sustained attention and awareness toward present-moment sensations without concentrating on any image, mantra or thought. The meditation approach includes daily mindfulness activities related to awareness of mind and body while focusing on the breath. The mindfulness-based cognitive therapy approach consists of integration of principles of cognitive behavioral therapy with mindfulness based stress reduction, designed to reduce relapse and recurrence of major depressive disorders (Goldin & Gross, 2010; Spijkerman et al., 2016). Yoga creates a balance between the two limbs of autonomic nervous system i.e. induction of parasympathetic nervous system (PNS) activation via release of neurotransmitters responsible for relaxation response, improving cardio-vagal tone, decreasing heart rate and blood pressure; and activation of limbic structures of the brain and suppression of sympathetic nervous system (SNS) activation, which is responsible for the stress response including the release of cortisol (Pramanik et al., 2009). With long term practice of yoga, PNS activation dominates over SNS activation, even during stressful situations (Ross & Thomas, 2010). Yoga can be a useful add-on therapy for the patients with chronic diseases to improve physical and mental health outcomes. A definitive role of yoga in pain management has been documented by various studies as it aids in increasing natural pain relievers, improving physical function, and creating a positive mental state (Telles & Singh, 2012, 2013). Pranayama yoga, Zen, transcendental meditation and other meditation program has diaphragmatic breathing as a mainstay procedure. Meditation proves to reduce anxiety, alleviate chronic pain, improve cognitive functions and lowers OS levels (lower cortisol levels and higher melatonin levels) (Khalsa et al., 2012; Martarelli et al., 2011). Yoga is a comprehensive philosophical system that seeks to bring balance into all aspects of body, mind, and soul and brings about cognitive improvement. Theories of self-regulation are assuming an increasingly central role within various sub-disciplines of cognitive science, psychology, and medicine(Hofmann & Smits, 2008). Self-regulation refers to efforts of monitoring, willpower, and motivation to manage or alter one’s incipient responses and impulses so as to pursue or maintain explicit goals or standards (Gard et al., 2014). As the meditation practice deepens, emphasis on interoception increases, evaluation processes decrease across contexts, and bottom-up strategies may be more strongly present. Top-down strategies involve higher brain processing like attentional deployment, response inhibition, reappraisal etc. Bottom-up regulation strategies have been described as modulation of emotion-generative brain regions (i.e., limbic) without recruitment of “higher” brain regions (i.e., frontal) that are responsible for cognitive forms of regulation (Gard et al., 2014). Bottom-up processes involve the influence of peripheral sensory, visceral, cardiovascular, immune, and autonomic input upon central neural processing and mental activities via ascending pathways (Taylor et al., 2011). Yoga is a complex, adaptive and widely applicable method of physical and mental training with multiple tools for self-development (see Figure 2). According to Patanjali’s yoga sutras, the component of classical yoga include: moral observances (ethics when interacting with others); self-discipline (ethics geared toward the self); physical postures and exercises; breath regulation; sensory withdrawal (minimizing sensory input); concentration (effortful, focused attention); meditation (effortless, unbroken flow of attention), and self-transcendence.

113

 Molecular Mechanisms Underlying the Effects of Yoga

CONCLUSION Yoga affects every cell of the body. The holistic science of yoga works via the regulation of psychoneuro-immune axis to bring about the positive changes to various organ systems of body. It brings about better neuro-effector communication, improves muscle strength and joint mobility, increases the optimum functioning of all organ-systems, increases resistance against stress and diseases and brings bliss and balance with a positive attitude and makes life purposeful. Yoga not only supports the physical and emotional wellbeing, but also opens door to self-actualization to create the perfect union of the mind, body, and soul. Figure 1. A proposed mechanism of action of yoga to enhance quality of life

Abbreviations: QOL: Quality of Life; BDNF: Brain derived neurotropic factor; NS: Nervous system

114

 Molecular Mechanisms Underlying the Effects of Yoga

Figure 2. Yoga based lifestyle intervention: A gateway to self-actualization

Abbreviations: GABA: Gamma aminobutyric acid; BDNF: Brain derived neurotropic factor; ROS: Reactive oxygen species, TAC: Total anti-oxidant capacity, 8OHDG: 8-hydroxy-2’ –deoxyguanosine, IL: Interleukin, TNF-α: Tumor necrosis factor-α, TGF-β: Transforming growth factor-β, CRP: C Reactive protein, ESR: Erythrocyte sedimentation rate, s HLA-G: Soluble human leukocyte antigen, Th17:T helper 17, Treg: T regulatory, DHEAS: dehydroepiandrosterone sulfate

115

 Molecular Mechanisms Underlying the Effects of Yoga

Table 1. Details of activities in a single session of yoga based lifestyle intervention program S. No. 1.

Practice to be done Prayer

2 min

Loosening Practices

3 min Supine

Prone 2.

Yogasana Sitting

Standing 3.

Duration

Relaxation

Shavasana

2 min

Uttanpadasana

2 min

Pawanmuktasana

2 min

Makarasana

2 min

Bhujangasana

2 min

Salabhasana

2 min

Vakrasana

2 min

Ardha-Matsyendrasana

2 min

Vajrasana

2 min

Tadasana

2 min

Vrikshasana

2 min

Ardhachakrasana

2 min

Shavasana

2 min

Nadishodhana Bhramari 4.

Pranayama

Shitkari

20 min

Shitali Brahmamudra 5.

Aumkar recitation

3 min

6.

Rajyoga meditation

10 min

7.

Shanti mantra

5 min

Total

70 min

ACKNOWLEDGMENT The study mentioned here was supported by the Department of Science and Technology (DST), India [SR/SATYAM/55/2016]; and Indian Council of Medical Research [Senior Research Fellowship [45/2/2019-ANA/BMS].

116

 Molecular Mechanisms Underlying the Effects of Yoga

REFERENCES Arora, S., & Bhattacharjee, J. (2008). Modulation of immune responses in stress by Yoga. International Journal of Yoga, 1(2), 45–55. doi:10.4103/0973-6131.43541 PMID:21829284 Bernardi, L., Sleight, P., Bandinelli, G., Cencetti, S., Fattorini, L., Wdowczyc-Szulc, J., & Lagi, A. (2001). Effect of rosary prayer and yoga mantras on autonomic cardiovascular rhythms: Comparative study. BMJ (Clinical Research Ed.), 323(7327), 1446–1449. doi:10.1136/bmj.323.7327.1446 PMID:11751348 Bhasin, M. K., Dusek, J. A., Chang, B. H., Joseph, M. G., Denninger, J. W., Fricchione, G. L., Benson, H., & Libermann, T. A. (2013). Relaxation response induces temporal transcriptome changes in energy metabolism, insulin secretion and inflammatory pathways. PLoS One, 8(5), e62817. doi:10.1371/journal. pone.0062817 PMID:23650531 Bisht, S., & Dada, R. (2017). Oxidative stress: Major executioner in disease pathology, role in sperm DNA damage and preventive strategies. Frontiers in Bioscience, 9(3), 420–447. doi:10.2741495 PMID:28410127 Bisht, S., Faiq, M., Tolahunase, M., & Dada, R. (2017). Oxidative stress and male infertility. Nature Reviews. Urology, 14(8), 470–485. doi:10.1038/nrurol.2017.69 PMID:28508879 Booklets on Common Yoga Protocol | Ministry of AYUSH | GOI. (n.d.). http://ayush.gov.in/bookletscommon-yoga-protocol-0 Bortolotti, D., Gentili, V., Rotola, A., Cassai, E., Rizzo, R., & Di Luca, D. (2014). Impact of HLA-G analysis in prevention, diagnosis and treatment of pathological conditions. World Journal of Methodology, 4(1), 11–25. doi:10.5662/wjm.v4.i1.11 PMID:25237627 Brown, R. P., & Gerbarg, P. L. (2005). Sudarshan Kriya yogic breathing in the treatment of stress, anxiety, and depression: Part I-neurophysiologic model. Journal of Alternative and Complementary Medicine (New York, N.Y.), 11(1), 189–201. doi:10.1089/acm.2005.11.189 PMID:15750381 Buric, I., Farias, M., Jong, J., Mee, C., & Brazil, I. A. (2017). What is the molecular signature of mindbody interventions? a systematic review of gene expression changes induced by meditation and related practices. Frontiers in Immunology, 8, 670. doi:10.3389/fimmu.2017.00670 PMID:28670311 Chu, P., Gotink, R. A., Yeh, G. Y., Goldie, S. J., & Hunink, M. G. M. (2016). The effectiveness of yoga in modifying risk factors for cardiovascular disease and metabolic syndrome: A systematic review and meta-analysis of randomized controlled trials. European Journal of Preventive Cardiology, 23(3), 291–307. doi:10.1177/2047487314562741 PMID:25510863 Costenbader, K. H., Prescott, J., Zee, R. Y., & De Vivo, I. (2011). Immunosenescence and rheumatoid arthritis: Does telomere shortening predict impending disease? Autoimmunity Reviews, 10(9), 569–573. doi:10.1016/j.autrev.2011.04.034 PMID:21575746 Cramer, H., Lauche, R., Haller, H., Dobos, G., & Michalsen, A. (2015). A systematic review of yoga for heart disease. European Journal of Preventive Cardiology, 22(3), 284–295. doi:10.1177/2047487314523132 PMID:24491402

117

 Molecular Mechanisms Underlying the Effects of Yoga

Dada, T., Mittal, D., Mohanty, K., Faiq, M. A., Bhat, M. A., Yadav, R. K., Sihota, R., Sidhu, T., Velpandian, T., Kalaivani, M., Pandey, R. M., Gao, Y., Sabel, B. A., & Dada, R. (2018). Mindfulness meditation reduces intraocular pressure, lowers stress biomarkers and modulates gene expression in glaucoma: A randomized controlled trial. Journal of Glaucoma, 27(12), 1061–1067. doi:10.1097/ IJG.0000000000001088 PMID:30256277 Dai, D.-F., Chiao, Y. A., Marcinek, D. J., Szeto, H. H., & Rabinovitch, P. S. (2014). Mitochondrial oxidative stress in aging and healthspan. Longevity & Healthspan, 3(1), 6. doi:10.1186/2046-2395-3-6 PMID:24860647 Dhawan, V., Kumar, M., & Dada, R. (2017). Effect of sperm molecular factors, oxidative damage and transcripts in childhood disorders. Journal of Childhood & Developmental Disorders, 3(1), 6. doi:10.4172/2472-1786.100044 Dhawan, V., Kumar, M., Deka, D., Malhotra, N., Dadhwal, V., Singh, N., & Dada, R. (2018). Meditation & yoga: Impact on oxidative DNA damage & dysregulated sperm transcripts in male partners of couples with recurrent pregnancy loss. The Indian Journal of Medical Research, 148, S134–S139. PMID:30964091 Dizdaroglu, M. (2012). Oxidatively induced DNA damage: Mechanisms, repair and disease. Cancer Letters, 327(1–2), 26–47. doi:10.1016/j.canlet.2012.01.016 PMID:22293091 Fernros, L., Furhoff, A.-K., & Wändell, P. E. (2008). Improving quality of life using compound mind-body therapies: Evaluation of a course intervention with body movement and breath therapy, guided imagery, chakra experiencing and mindfulness meditation. Quality of Life Research: An International Journal of Quality of Life Aspects of Treatment, Care and Rehabilitation, 17(3), 367–376. doi:10.100711136008-9321-x PMID:18324479 Fujii, H., Shao, L., Colmegna, I., Goronzy, J. J., & Weyand, C. M. (2009). Telomerase insufficiency in rheumatoid arthritis. Proceedings of the National Academy of Sciences of the United States of America, 106(11), 4360–4365. doi:10.1073/pnas.0811332106 PMID:19255426 Gagrani, M., Faiq, M. A., Sidhu, T., Dada, R., Yadav, R. K., Sihota, R., Kochhar, K. P., Verma, R., & Dada, T. (2018). Meditation enhances brain oxygenation, upregulates BDNF and improves quality of life in patients with primary open angle glaucoma: A randomized controlled trial. Restorative Neurology and Neuroscience, 36(6), 741–753. doi:10.3233/RNN-180857 PMID:30400122 Gard, T., Noggle, J. J., Park, C. L., Vago, D. R., & Wilson, A. (2014). Potential self-regulatory mechanisms of yoga for psychological health. Frontiers in Human Neuroscience, 8, 770. doi:10.3389/ fnhum.2014.00770 PMID:25368562 Gautam, S., Chawla, B., Bisht, S., Tolahunase, M., & Dada, R. (2018). Impact of mindfulness based stress reduction on sperm DNA damage. Journal of the Anatomical Society of India, 67(2), 124–129. doi:10.1016/j.jasi.2018.07.003 Gautam, S., Kumar, M., Chaurasia, P., Rana, D., & Dada, R. (2019). Impact of yoga based lifestyle intervention on quality of life, depression and sperm oxidative DNA damage: A randomized controlled trial on infertile men with rheumatoid arthritis. Fertility and Sterility, 112(3), e409. doi:10.1016/j.fertnstert.2019.07.1161

118

 Molecular Mechanisms Underlying the Effects of Yoga

Gautam, S., Tolahunase, M., Kumar, U., & Dada, R. (2019). Impact of yoga based mind-body intervention on systemic inflammatory markers and co-morbid depression in active Rheumatoid arthritis patients: A randomized controlled trial. Restorative Neurology and Neuroscience, 37(1), 41–59. doi:10.3233/ RNN-180875 PMID:30714983 Geneen, L. J., Moore, R. A., Clarke, C., Martin, D., Colvin, L. A., & Smith, B. H. (2017). Physical activity and exercise for chronic pain in adults: An overview of Cochrane Reviews. Cochrane Database of Systematic Reviews, 4, CD011279. PMID:28436583 Genud, R., Merenlender, A., Gispan-Herman, I., Maayan, R., Weizman, A., & Yadid, G. (2009). DHEA lessens depressive-like behavior via GABA-ergic modulation of the mesolimbic system. Neuropsychopharmacology: Official Publication of the American College of Neuropsychopharmacology, 34(3), 577–584. doi:10.1038/npp.2008.46 PMID:18496525 Goldin, P. R., & Gross, J. J. (2010). Effects of mindfulness-based stress reduction (MBSR) on emotion regulation in social anxiety disorder. Emotion (Washington, D.C.), 10(1), 83–91. doi:10.1037/a0018441 PMID:20141305 Hodes, G. E., Kana, V., Menard, C., Merad, M., & Russo, S. J. (2015). Neuroimmune mechanisms of depression. Nature Neuroscience, 18(10), 1386–1393. doi:10.1038/nn.4113 PMID:26404713 Hofmann, S. G., & Smits, J. A. J. (2008). Cognitive-behavioral therapy for adult anxiety disorders: A meta-analysis of randomized placebo-controlled trials. The Journal of Clinical Psychiatry, 69(4), 621–632. doi:10.4088/JCP.v69n0415 PMID:18363421 Innes, K. E., Bourguignon, C., & Taylor, A. G. (2005). Risk indices associated with the insulin resistance syndrome, cardiovascular disease, and possible protection with yoga: A systematic review. The Journal of the American Board of Family Practice, 18(6), 491–519. doi:10.3122/jabfm.18.6.491 PMID:16322413 Jayadevappa, R., Johnson, J. C., Bloom, B. S., Nidich, S., Desai, S., Chhatre, S., Raziano, D. B., & Schneider, R. (2007). Effectiveness of transcendental meditation on functional capacity and quality of life of African Americans with congestive heart failure: A randomized control study. Ethnicity & Disease, 17(1), 72–77. PMID:17274213 Kanojia, S., Sharma, V. K., Gandhi, A., Kapoor, R., Kukreja, A., & Subramanian, S. K. (2013). Effect of yoga on autonomic functions and psychological status during both phases of menstrual cycle in young healthy females. Journal of Clinical and Diagnostic Research: JCDR, 7(10), 2133–2139. PMID:24298457 Khalsa, S. B. S., Hickey-Schultz, L., Cohen, D., Steiner, N., & Cope, S. (2012). Evaluation of the mental health benefits of yoga in a secondary school: A preliminary randomized controlled trial. The Journal of Behavioral Health Services & Research, 39(1), 80–90. doi:10.100711414-011-9249-8 PMID:21647811 Kim, J., & Gorman, J. (2005). The psychobiology of anxiety. Clinical Neuroscience Research, 4(5), 335–347. doi:10.1016/j.cnr.2005.03.008 Kinser, P. A., Goehler, L. E., & Taylor, A. G. (2012). How might yoga help depression? A neurobiological perspective. Explore (New York, N.Y.), 8(2), 118–126. doi:10.1016/j.explore.2011.12.005 PMID:22385566

119

 Molecular Mechanisms Underlying the Effects of Yoga

Krishna, B. H., Pal, P., Pal, G., Balachander, J., Jayasettiaseelon, E., Sreekanth, Y., Sridhar, M. G., & Gaur, G. (2014). A randomized controlled trial to study the effect of yoga therapy on cardiac function and n terminal pro BNP in heart failure. Integrative Medicine Insights, 9, 1–6. doi:10.4137/IMI.S13939 PMID:24737932 Lachance, C. C., & McCormack, S. (2019). Mindfulness training and yoga for the management of chronic non-malignant pain: A review of clinical effectiveness and cost-effectiveness. Canadian Agency for Drugs and Technologies in Health. https://www.ncbi.nlm.nih.gov/books/NBK549581/ Lee, M., Moon, W., & Kim, J. (2014). Effect of yoga on pain, brain-derived neurotrophic factor, and serotonin in premenopausal women with chronic low back pain. Evidence-Based Complementary and Alternative Medicine: ECAM, 2014, 203173. doi:10.1155/2014/203173 PMID:25120574 Martarelli, D., Verdenelli, M. C., Scuri, S., Cocchioni, M., Silvi, S., Cecchini, C., & Pompei, P. (2011). Effect of a probiotic intake on oxidant and antioxidant parameters in plasma of athletes during intense exercise training. Current Microbiology, 62(6), 1689–1696. doi:10.100700284-011-9915-3 PMID:21400082 Mazor, M., Lee, J. Q., Peled, A., Zerzan, S., Irwin, C., Chesney, M. A., Serrurier, K., Sbitany, H., Dhruva, A., Sacks, D., & Smoot, B. (2018). The effect of yoga on arm volume, strength, and range of motion in women at risk for breast cancer-related lymphedema. Journal of Alternative and Complementary Medicine (New York, N.Y.), 24(2), 154–160. doi:10.1089/acm.2017.0145 PMID:29064279 McInnes, I. B., & Schett, G. (2011). The pathogenesis of rheumatoid arthritis. The New England Journal of Medicine, 365(23), 2205–2219. doi:10.1056/NEJMra1004965 PMID:22150039 Mishra, B. (2012). PA01.68. The effect of yoga nidra in the management of rheumatoid arthritis. Ancient Science of Life, 32(5, Suppl 1), S118. doi:10.4103/0257-7941.112125 Morgan, N., Irwin, M. R., Chung, M., & Wang, C. (2014). The effects of mind-body therapies on the immune system: Meta-analysis. PLoS One, 9(7), e100903. doi:10.1371/journal.pone.0100903 PMID:24988414 Nunn, A. V., Bell, J. D., & Guy, G. W. (2009). Lifestyle-induced metabolic inflexibility and accelerated ageing syndrome: Insulin resistance, friend or foe? Nutrition & Metabolism, 6(1), 16. doi:10.1186/17437075-6-16 PMID:19371409 Pilkington, K., Kirkwood, G., Rampes, H., & Richardson, J. (2005). Yoga for depression: The research evidence. Journal of Affective Disorders, 89(1–3), 13–24. doi:10.1016/j.jad.2005.08.013 PMID:16185770 Pramanik, T., Sharma, H. O., Mishra, S., Mishra, A., Prajapati, R., & Singh, S. (2009). Immediate effect of slow pace bhastrika pranayama on blood pressure and heart rate. Journal of Alternative and Complementary Medicine (New York, N.Y.), 15(3), 293–295. doi:10.1089/acm.2008.0440 PMID:19249921 Pullen, P. R., Seffens, W. S., & Thompson, W. R. (2018). Yoga for heart failure: A review and future research. International Journal of Yoga, 11(2), 91–98. doi:10.4103/ijoy.IJOY_24_17 PMID:29755216 Qu, S., Olafsrud, S. M., Meza-Zepeda, L. A., & Saatcioglu, F. (2013). Rapid gene expression changes in peripheral blood lymphocytes upon practice of a comprehensive yoga program. PLoS One, 8(4), e61910. doi:10.1371/journal.pone.0061910 PMID:23613970

120

 Molecular Mechanisms Underlying the Effects of Yoga

Rajbhoj, P. H., Shete, S. U., Verma, A., & Bhogal, R. S. (2015). Effect of yoga module on pro-inflammatory and anti-inflammatory cytokines in industrial workers of lonavla: A randomized controlled trial. Journal of Clinical and Diagnostic Research: JCDR, 9(2), CC01–CC05. PMID:25859450 Rathore, M., Trivedi, S., Abraham, J., & Sinha, M. B. (2017). Anatomical correlation of core muscle activation in different yogic postures. International Journal of Yoga, 10(2), 59–66. doi:10.4103/09736131.205515 PMID:28546675 Riley, D. (2004). Hatha yoga and the treatment of illness. Alternative Therapies in Health and Medicine, 10(2), 20–21. PMID:15055088 Rima, D., Shiv, B. K., Bhavna, C., Shilpa, B., & Saima, K. (2016). Oxidative stress induced damage to paternal genome and impact of meditation and yoga—Can it reduce incidence of childhood cancer? Asian Pacific Journal of Cancer Prevention, 17(9), 4517–4525. PMID:27880996 Ross, A., & Thomas, S. (2010). The health benefits of yoga and exercise: A review of comparison studies. Journal of Alternative and Complementary Medicine (New York, N.Y.), 16(1), 3–12. doi:10.1089/ acm.2009.0044 PMID:20105062 Santana, M.-J., S-Parrilla, J., Mirus, J., Loadman, M. A., Lien, D. C., & Feeny, D. (2013). An assessment of the effects of Iyengar yoga practice on the health-related quality of life of patients with chronic respiratory diseases: A pilot study. Canadian Respiratory Journal, 20(2), e17–e23. doi:10.1155/2013/265406 PMID:23616973 Schneier, F. R., Liebowitz, M. R., Abi-Dargham, A., Zea-Ponce, Y., Lin, S. H., & Laruelle, M. (2000). Low dopamine D(2) receptor binding potential in social phobia. The American Journal of Psychiatry, 157(3), 457–459. doi:10.1176/appi.ajp.157.3.457 PMID:10698826 Selden, N. R., Robbins, T. W., & Everitt, B. J. (1990). Enhanced behavioral conditioning to context and impaired behavioral and neuroendocrine responses to conditioned stimuli following ceruleocortical noradrenergic lesions: Support for an attentional hypothesis of central noradrenergic function. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 10(2), 531–539. doi:10.1523/ JNEUROSCI.10-02-00531.1990 PMID:2303858 Sengupta, P. (2012). Health Impacts of Yoga and Pranayama: A State-of-the-Art Review. International Journal of Preventive Medicine, 3(7), 444–458. PMID:22891145 Shammas, M. A. (2011). Telomeres, lifestyle, cancer, and aging. Current Opinion in Clinical Nutrition and Metabolic Care, 14(1), 28–34. doi:10.1097/MCO.0b013e32834121b1 PMID:21102320 Sharma, V. K., Das, S., Mondal, S., Goswami, U., & Gandhi, A. (2006). Effect of Sahaj Yoga on neurocognitive functions in patients suffering from major depression. Indian Journal of Physiology and Pharmacology, 50(4), 375–383. PMID:17402267 Shete, S. U., Verma, A., Kulkarni, D. D., & Bhogal, R. S. (2017). Effect of yoga training on inflammatory cytokines and C-reactive protein in employees of small-scale industries. Journal of Education and Health Promotion, 6, 76. PMID:28856165

121

 Molecular Mechanisms Underlying the Effects of Yoga

Singh, V. P., Khandelwal, B., & Sherpa, N. T. (2015). Psycho-neuro-endocrine-immune mechanisms of action of yoga in type II diabetes. Ancient Science of Life, 35(1), 12–17. doi:10.4103/0257-7941.165623 PMID:26600662 Sinha, S., Singh, S. N., Monga, Y. P., & Ray, U. S. (2007). Improvement of glutathione and total antioxidant status with yoga. Journal of Alternative and Complementary Medicine (New York, N.Y.), 13(10), 1085–1090. doi:10.1089/acm.2007.0567 PMID:18166119 Spijkerman, M. P. J., Pots, W. T. M., & Bohlmeijer, E. T. (2016). Effectiveness of online mindfulnessbased interventions in improving mental health: A review and meta-analysis of randomised controlled trials. Clinical Psychology Review, 45, 102–114. doi:10.1016/j.cpr.2016.03.009 PMID:27111302 Stephens, I. (2017). Medical yoga therapy. Children (Basel, Switzerland), 4(2), 12. doi:10.3390/children4020012 PMID:28208599 Swanson, A. (2019). Science of yoga. Penguin Random House. Taylor, V. A., Grant, J., Daneault, V., Scavone, G., Breton, E., Roffe-Vidal, S., Courtemanche, J., Lavarenne, A. S., & Beauregard, M. (2011). Impact of mindfulness on the neural responses to emotional pictures in experienced and beginner meditators. NeuroImage, 57(4), 1524–1533. doi:10.1016/j.neuroimage.2011.06.001 PMID:21679770 Telles, S., & Singh, N. (2012). Is yoga a suitable treatment for rheumatoid arthritis: Current opinion. Open Access Journal of Sports Medicine, 3, 81–87. doi:10.2147/OAJSM.S25707 PMID:24198591 Telles, S., & Singh, N. (2013). Science of the mind: Ancient yoga texts and modern studies. The Psychiatric Clinics of North America, 36(1), 93–108. doi:10.1016/j.psc.2013.01.010 PMID:23538080 Tolahunase, M., Sagar, R., & Dada, R. (2017). Impact of yoga and meditation on cellular aging in apparently healthy individuals: A prospective, open-label single-arm exploratory study. Oxidative Medicine and Cellular Longevity, 2017, 7928981. PMID:28191278 Tolahunase, M. R., Sagar, R., Faiq, M., & Dada, R. (2018). Yoga- and meditation-based lifestyle intervention increases neuroplasticity and reduces severity of major depressive disorder: A randomized controlled trial. Restorative Neurology and Neuroscience, 36(3), 423–442. doi:10.3233/RNN-170810 PMID:29614706 Vijayaraghava, A., Doreswamy, V., Narasipur, O. S., Kunnavil, R., & Srinivasamurthy, N. (2015). Effect of yoga practice on levels of inflammatory markers after moderate and strenuous exercise. Journal of Clinical and Diagnostic Research: JCDR, 9(6), CC08–CC12. doi:10.7860/JCDR/2015/12851.6021 PMID:26266115 Woodyard, C. (2011). Exploring the therapeutic effects of yoga and its ability to increase quality of life. International Journal of Yoga, 4(2), 49–54. doi:10.4103/0973-6131.85485 PMID:22022122 Woolery, A., Myers, H., Sternlieb, B., & Zeltzer, L. (2004). A yoga intervention for young adults with elevated symptoms of depression. Alternative Therapies in Health and Medicine, 10(2), 60–63. PMID:15055096

122

 Molecular Mechanisms Underlying the Effects of Yoga

Yadav, R. K., Magan, D., Mehta, N., Sharma, R., & Mahapatra, S. C. (2012). Efficacy of a short-term yoga-based lifestyle intervention in reducing stress and inflammation: Preliminary results. Journal of Alternative and Complementary Medicine (New York, N.Y.), 18(7), 662–667. doi:10.1089/acm.2011.0265 PMID:22830969 Zautra, A. J., Davis, M. C., Reich, J. W., Nicassario, P., Tennen, H., Finan, P., Kratz, A., Parrish, B., & Irwin, M. R. (2008). Comparison of cognitive behavioral and mindfulness meditation interventions on adaptation to rheumatoid arthritis for patients with and without history of recurrent depression. Journal of Consulting and Clinical Psychology, 76(3), 408–421. doi:10.1037/0022-006X.76.3.408 PMID:18540734

123

124

Chapter 8

Neurocognitive Mechanisms of Yoga:

Implications for Yoga Therapy Irina Sheftel Network Yoga Therapy, The Netherlands Anneke Sips Network Yoga Therapy, The Netherlands

ABSTRACT This chapter combines insights from the neuroscientific research on yoga and meditation with the practical application of yoga therapy. The chapter opens with an overview of neuroscientific concepts related to the practice of yoga and meditation, such as attention, emotional regulation, body awareness, and autonomic regulation. The authors summarise the known effects of yoga and meditation, in relation to these concepts and common mental health disorders. The chapter continues with a case study, a personalised yoga therapy intervention in a client experiencing psychotic symptoms. The intervention is grounded in yoga philosophy, and in the neuroscientific concepts introduced previously. This section is written from the perspective of a yoga therapist and includes practical tips. In the final section, the authors suggest possible directions for future research and implementation of yoga therapy as a complementary treatment in mental healthcare.

INTRODUCTION Yoga philosophy is formed around the idea of personal transformation. The ultimate goal of yoga — stilling thoughts and achieving the realisation of self-awareness — can be reached through consistent practice. This idea overlaps with modern neuroscience, which holds that our brain can change in response to experience and training. Yoga and meditation could be seen as a form of mental exercise, that transform the neurocognitive processes underlying our personalities. A large number of empirical studies show the positive effects of yoga and meditation-based interventions on mental health (Goldberg & Tucker, DOI: 10.4018/978-1-7998-3254-6.ch008

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Neurocognitive Mechanisms of Yoga

2020; Wielgosz et al., 2019). Inspired by this wave of research, yoga and meditation gradually become well-validated approaches in Western mental healthcare. Yoga practice has many components, and in the broadest sense can be seen as a holistic approach to wellbeing. Yoga traditions, adopted in the West, vary in their focus on different elements of the practice (postures, breathing exercises, relaxation, and meditation). An individual yoga practice develops and changes over time, following one’s changing needs. The content of yoga practice, and specific instructions can thus hardly be fixed. As a result, yoga interventions do not rely on standardised practices (Sherman, 2012), and clinical research on yoga employ different protocols. Instead of working with standardised protocols and labels, therapeutic yoga is traditionally individualised and tailored to the person’s needs. Every person can have episodes of anxiety, depression or psychotic dissociation to one degree or another. A yoga therapist needs to see the mental health problems on a spectrum, to place them in the context of one’s life events and approach them holistically at many levels of being. It is also essential to understand the neurocognitive mechanisms underlying psychiatric symptoms and map the effects of yoga in terms of cognitive and affective processes. This knowledge may help to choose an optimal approach for a yoga therapy treatment. In this chapter, we combine insights from the neuroscientific research of yoga and meditation with practical implications for yoga therapy. In the first part, we discuss the fundamental neuroscientific concepts related to the practice of yoga and meditation: attention, emotional regulation, body awareness, and autonomic regulation. We summarise evidence on the neurophysiological effects of yoga from existing clinical research, in relation to common mental health problems. We then present a case study of a personalised yoga therapy intervention in a client experiencing psychotic symptoms. We discuss this case from the perspective of a yoga therapist and show that the intervention has its foundation in the yoga philosophy and the modern neuroscientific concepts. Finally, we discuss future perspectives and directions for the implementation of yoga therapy in mental healthcare.

YOGA AND MEDITATION: NEUROCOGNITIVE MECHANISMS AND CLINICAL APPLICATIONS Meditation and Regulation of Attention Modern science considers meditation as a form of mental training that can improve meta-awareness – i.e., the awareness of one’s processes of consciousness, including thinking, feeling and perceiving. Meta-awareness is closely related to mindfulness – an internal stance of attention to present moment characterised with curiosity, openness and acceptance. Even though the term mindfulness is usually connected to Buddhist meditation tradition, it is also a fundamental element of yoga practice, described in the Yoga Sutras of Patanjali as the state of self-observation, concentration and introspection. In recent years, an increasing number of research studies has shown that regular mindfulness training can improve self-regulation, stress-resilience and wellbeing. Meditation-based interventions are promoted these days as a way to cultivate a healthy, balanced and compassionate way of living (Singer & Engert, 2018). Protocol-based interventions such as the 8-week mindfulness-based stress reduction (MBSR) program and mindfulness-based cognitive therapy (MBCT) have built a strong reputation in mainstream clinical and educational settings (Singer & Engert, 2018). At the same time, research in

125

 Neurocognitive Mechanisms of Yoga

psychology and neuroscience is providing a better understanding of the cognitive processes underlying the positive effects of meditation. Dahl and colleagues (2015) proposed a classification system that categorises specific styles of meditation based on their primary cognitive mechanism. Meditation practices described in traditional yoga texts fall under the attentional family. These practices involve systematic training to self-regulate attentional processes, intentionally direct attention to monitor, detect, and disengage from distractors, and to reorient attention toward a chosen object (Tang et al., 2015). The target of meditation may differ: it can be an external object, the breath, an image, a mantra, or bodily sensations, or awareness itself. However, irrespective of the meditation target, instructions for the practice do not differ. The practitioner is invited to deliberately direct attention to the object of meditation and sustain it via continuous monitoring. If distracted, the practitioner is encouraged to notice the distraction and gently bring the attention back to the meditative target. Cultivating the ability to deliberately control attention is thus the central aspect of meditation. This task, seemingly simple, involves several associated cognitive processes. It requires meta-awareness — awareness of mental processes; attentional control — the ability to maintain the focus on the task; and inhibition — the ability to disengage from distractors (Lutz et al., 2008; Wielgosz et al., 2019). These high-level mental processes are supported by large-scale functional networks in the brain. Three brain networks are especially relevant for the attentional meditation practice: 1. The Central Executive Network includes parietal and prefrontal cortices. These brain regions support the control and monitoring of goal-driven behaviour. Prefrontal activations are often enhanced as an effect of meditation in beginners. In contrast, compared to novice practitioners, experienced meditators show reduced activation in these regions during meditation (Tang et al., 2015). These findings indicate that voluntary attention regulation, while challenging in the beginning, may become automated with practice (Tang et al., 2015). 2. The Salience Network is involved in detecting and filtering sensory stimuli and orienting attention. The network incorporates the anterior insula, dorsal anterior cingulate cortex (ACC), amygdala and dopaminergic midbrain regions. Studies on the neurocognitive effects of meditation consistently show increased activation in the anterior insula, dorsal ACC and midcingulate cortex (Fox et al., 2016; Tang et al., 2015). Functional changes in the Salience network resulting from meditation are attributed to changes in attention monitoring and interoceptive awareness (Cotier et al., 2017; Farb et al., 2013). 3. Default Mode Network (i.e., medial prefrontal cortex, posterior cingulate cortex, and medial temporal areas) is consistently more active when no particular task must be done, and a person is resting. It is commonly assumed that this brain network is a source of spontaneous thoughts, e.g. the mind wandering experience (Christoff et al., 2016). Research has linked attention impairments such as attention-deficit/hyperactivity disorder (ADHD) with a distorted regulation of DMN (Bozhilova et al., 2018; Christoff et al., 2016). Neuroimaging studies converge on the finding that meditation reduces the activation of DMN and alters connectivity within sub networks of DMN. The reduction in activity in the Default Mode Network in the brain is consistent with the observation that meditation reduces spontaneous mind wandering. Meditation thus could especially benefit people who have ADHD. Studies have reported positive effects of meditation, which are potentially comparable to those of medication, in this clinical group (Hepark et al., 2019).

126

 Neurocognitive Mechanisms of Yoga

Box 1. Yoga therapy in attention problems: a yoga therapist’s perspective People diagnosed with ADHD/ADD often receive medication. These medications do not provide a permanent cure for ADHD and have a long list of side effects. Ideally, a non-pharmacological approach should be sought to improve attention deficit in such patients. In yoga therapy, attention capacity can be trained in many ways. Attention is the core component of all yoga practices: meditation, asanas, mantra, pranayama. The practice requires time and discipline, which can be difficult for those with ADHD to master. In many cases, it may take 6-12 months for the yoga practice to make any difference, while medication works within minutes. On the other hand, drug effects wear off while the benefits of yoga practice, e.g. ability to focus, be less impulsive, feel calmer, in addition to learning and practising new skills as such, last much longer. The yoga therapy approach empowers the client because she learns to apply yoga tools when necessary. From the holistic yoga perspective, learning and practice transform the entire person. There is no learning or transformation involved in taking a pill. Supporting the client in finding ways to integrate the practices in her/his life are the skills of a yoga therapist that should not be underestimated. The therapist should know how to customise the practice for the person in the most effective way. There is not a “one size fits all” solution for all clients.

In summary, meditation training results in long-term and short-term changes in the attentional and executive networks in the brain. Given the observed changes at the neural level, one might expect substantial positive effects of meditation on executive function. Indeed, meditation produces measurable effects on meta-awareness and cognitive flexibility, e.g. the capacity to be aware of the processes of thinking, to detect distractors, and switch attention from a distractor to the task (Brandmeyer et al., 2019; Wielgosz et al., 2019). At the same time, little support has been found for the improvement of working memory and sustained attention (Lao et al., 2016). This suggests that cognitive control functions are selectively affected by meditation.

Meditation and Emotional Regulation The Link between Attention and Emotions Imagine seeing a snake. The threatful image produces a sense of fear and instantly captures attention. A memory or thought can also cause strong emotions, and grab our attention just like an image or a scene. Studies show that the emotional intensity determines how well the thought can be recalled later, and whether it will be present in subsequent dreams (Tyng et al., 2017; Vallatet et al., 2018). Because of their effects on attention, thoughts associated with negative emotions may become reinforced in a closed feedback loop (Delorme & Brandmeyer, 2019). Through reinforcement, negative thinking may become repetitive and excessive and may contribute substantially to psychological disorders such as depression, anxiety, and attention deficit. Meditation and mindfulness-based practices are strongly linked with emotional regulation. Research studies in healthy participants as well as in participants suffering from depression show that meditation improves the ability to identify negative thoughts and disengage from them (Lutz et al., 2008; Wielgosz et al., 2019). By being aware of the mental processes, one does not directly attribute emotional value to the thought. When a thought is not reinforced with strong negative emotions, the mental process exits the reinforcement feedback loop (Delorme & Brandmeyer, 2019). Thus, with practice, negative thoughts may lose their grabbing power. “Meditation is to be aware of every thought and every feeling, never to say it is right or wrong, but just to watch it and move with it. In that watching, you begin to understand the whole movement of thought and feeling. And out of this awareness comes silence.” (Krishnamurti, 2002). Meditation also helps to cultivate adaptive emotional regulation strategies, such as acceptance, perspective taking and positive reappraisal (Lutz et al., 2008; Wielgosz et al., 2019). Acceptance is a mental

127

 Neurocognitive Mechanisms of Yoga

attitude of non-judgment and receptivity toward internal and external experiences (Engert et al., 2017; Lindsay & Creswell, 2017). Positive reappraisal is a cognitive skill that permits seeing a negative or stressful event as meaningful or beneficial. These techniques may help to improve emotional regulation further and lessen the impact of negative experiences and thoughts (Dahl et al., 2015). For example, a recent study shows that long-term meditation practitioners have faster cortisol recovery from stress, which is mediated by these emotional regulation strategies (Gamaiunova et al., 2019). Exercising these cognitive strategies within meditation-based intervention results in significant physiological stress reduction (Engert et al., 2017) and supports mental wellbeing.

Cultivating Positive Emotions: Loving-Kindness Meditation The effects of meditation on emotions that we discussed so far are gained through increased meta-awareness and cognitive flexibility. However, the practitioner may also directly cultivate positive emotional qualities, such as loving-kindness, compassion, and gratitude. These practices fall under the constructive family of meditations (Dahl et al., 2015), and aim at strengthening positive psychological patterns that foster well-being. Training of loving-kindness involves the active generation of the feeling of warmth, love, kindness towards oneself and others. Extensive training in such practices leads to a sustained change in emotional regulation. For example, people who practised loving-kindness meditation in an extensive 3-month training made more altruistic choices at the end of the training, compared to an active control group (Engen et al., 2018). Loving-kindness and compassion meditation induces plastic changes in parietal and frontoinsular regions, two important components of the brain network that mediates social and affective skills (Singer & Engert, 2018). Long-term practitioners of loving-kindness meditation show increased cortical thickness in the left ventrolateral prefrontal cortex and anterior insula, brain regions that support attention, empathy, and communication (Engen et al., 2018). Importantly, studies show that these results are specific for socio-affective training. Other types of meditation do not necessarily produce such effects (Singer & Engert, 2018).

Self-Compassion Socio-affective training, such as loving-kindness meditation, increases the quality of compassion, e.g. the attitude of kindness and care. Compassion can be understood as an integrated flow, involving compassion to others, receiving compassion from others, and also self-compassion (Gilbert, 2019). There is evidence that compassion and self-compassion have an effect on a range of physiological and psychological conditions, which included depression, anxiety, and stress (Neff & Germer, 2018; Germer & Neff, 2019). The practice of self-compassion includes three main components: i) self-kindness ii) a sense of common humanity iii) mindfulness (Germer & Neff, 2019). Self-kindness entails being warm and caring toward ourselves in challenging times in our lives. Common humanity recognises the shared nature of suffering when difficult situations arise, rather than feeling desperately alone. Mindfulness refers here to the ability to open up to a painful experience (“Ouch! It hurts!”) with a non-reactive and balanced awareness. These three components offer a stable and calmer response to trauma and distress. Self-kindness can have a calming effect; common humanity can be the answer to the feeling of isolation and shame; and balanced, mindful awareness allows us to take a different perspective on our painful memories and feelings.

128

 Neurocognitive Mechanisms of Yoga

Box 2. Compassion and self-compassion in yoga therapy “By cultivating attitudes of friendliness toward the happy, compassion for the unhappy, delight in the virtuous, and disregard toward the wicked, the mind-stuff retains its undisturbed calmness”. Yoga Sutra 1.33 The practice of yoga offers space for developing compassion and self-compassion, for example, by learning how to allow and to stay with painful feelings and self-critical thoughts that may come up during the practice. When, for example, holding a pose and one’s inner critic sneaks in, yoga teaches one to observe feelings and thoughts and choose a compassionate response. By practising self-compassion, a client becomes equipped with tools to make more sensible, self-supportive decisions on the mat and in daily life. At the same time, clients may feel isolated and lonely, and be very blocked on receiving compassion. Some even have a fear of selfcompassion as they believe it might reduce their productivity. In such cases compassion experience may begin with experiencing compassion from a teacher/therapist before they can become self-compassionate. In this approach, the teacher prioritises two elements: providing a sense of safety and normalising a person’s vulnerability. The yoga therapist does not try to deny and undermine problems; rather, together with the client, we learn how to cultivate more compassion toward themselves and their experiences.

Movement and Body Awareness Modern yoga combines meditative practices and postural exercise. Asana practice can be considered as a form of meditation: slow movement creates a flow of sensory information that modulates overall arousal and attention, and the practitioner has to keep an active intention to stay focused on the movement and monitor mental and physical experiences. Just as any other form of physical exercise, regular asana practice improves balance, strength, stability and motor control. On top of that, the combination of movement and meditation uniquely improves body awareness. Body awareness constitutes the recognition and awareness of senses related to the body position in the space (proprioception) and internal bodily states (interoception). Proprioception is the process of perceiving the position, location, orientation, and movement of the body and its parts, and to sense stimuli arising within the body regarding position, motion, and equilibrium. The integration of proprioceptive signals allows movement and provides a sense of agency (i.e. the subjective awareness of controlling one’s own motor actions). Interoception refers to constant monitoring of visceral sensations such as temperature, stretch, and pain from the gut, touch, itch, tickle, hunger, nausea, thirst, sleepiness, and sexual desire (Craig, 2002). These sensations are used to maintain homeostasis, to infer emotional states, to elicit behaviours (Quattrocki & Friston, 2014). Ultimately, proprioceptive and interoceptive signals provide a sense of the embodied “self”, which encompasses bodily self-awareness, emotional self-awareness, continuity of ‘me’ across time and the sense of agency (Quattrocki & Friston, 2014). Body awareness is thus uniquely central in emotional regulation and the sense of self (Azzalini et al., 2019; Clark et al., 2015). Deficits in proprioceptive and interoceptive awareness are common for mental health conditions and are often associated with poor emotional recognition in self and others. Movement and body awareness are critical elements of yoga therapy. Research in clinical populations (Russell & Arcuri, 2015) suggests that mindful movement could particularly benefit for several clinical groups, such as individuals who may have: i) compromised attentional capabilities and meta-awareness (e.g. ADHD); ii) over-activity in the motor system (e.g. Tics, Tourette syndrome); iii) and those suffering from disorganised, and distressing mental experiences: visual or auditory hallucinations, paranoid delusions, racing thoughts and suicidal ideation, depersonalisation (e.g., psychosis and schizophrenia). The tools that yoga provides, applied in daily life, may assist a person in connecting to their body and finding the inner resources that help them to be more confident in the challenging journey toward recovery.

129

 Neurocognitive Mechanisms of Yoga

Box 3. Body-awareness in yoga therapy Tuning in with bodily sensations gives us a key for monitoring and regulating emotions. We know that emotions have physical correlates — for example, mental stress may produce muscular tension, and fear and anxiety result in a shallow breath. Yoga practice in general, and particularly the body-focused meditation (e.g., body scan) instructs participants to discover physical tension in different parts of the body and to become aware of the connection between the physical discomfort and emotional state. Participants are encouraged to take a stance of compassionate acceptance toward every experience, including thoughts, feelings, memories, and emotions. These practices help to be ‘at peace’ with both physical and mental discomforts.

Bottom-up Effects: Yoga and Autonomic Regulation As laid out in the sections above, current neuroscientific research sees meditation as mental training, attributing its beneficial effects to such top-down processes as attention and meta-cognition. This section will bring into the picture the bottom-up effects of meditation and yoga: the visceral changes and their subsequent effect on the mental state. Many contemplative traditions, including Zen, vipassana, and yoga, emphasise slow deep diaphragmatic breathing. This particular pattern of slow breathing stimulates the vagus nerve activity, lowering heart rate and blood pressure, inhibiting the sympathetic nervous system and stress response, and modulating the limbic system. Moreover, in response to slower breathing, afferent vagal nerve pathways signal a state of relaxation and low-threat, resulting in a further increase of vagal tone. This positive reinforcement loop may be responsible for long-term changes in vagal tone, supporting the emotional and cognitive effects of meditation (Gerritsen & Band, 2018). Understanding of the connection between the visceral and cognitive effects of yoga has its foundation in the polyvagal theory (Lucas et al., 2018; Porges, 2007). The key components of yoga, slow movement and slow breathing, calm the physiological and behavioural state by increasing the vagal influence through a branch of the vagus nerve that originates in the nucleus ambiguus. This functional subsystem of the autonomic neural system can be collectively described as the ventral vagal complex. These neural circuits are involved in the regulation of the heart and the muscles of the face and head and control the visceral and cognitive responses in a state of safety and positive engagement with the environment (Porges, 2011). The polyvagal theory posits that when the ventral vagal complex is activated, the body and mind shift into the optimal state that supports health, growth, and restoration (Lucas et al., 2018). Stephen Porges, the founder of the polyvagal theory, has coined the term “neuroception” – the unconscious ability to quickly adapt bodily responses to danger or safety. When we detect danger, we shrink and tense, our breathing becomes shallow, our heart beats faster, and our voice pitch becomes higher. On the other hand, when we feel safe, we release muscle tension, soften the breath and the voice, and tune in for connecting with others. All these changes reflect a shift in the state of the autonomic neural system. While these reactions are outside of our awareness, we can consciously use posture, breath, and voice to get to this optimal state. Yoga practice gives us a unique way to reset our physiology to experience safety. The experience of safety is closely linked to the connectedness to another human being. We often see connecting to others as a psychological or mental act. But connection happens at the bodily level too. To connect to another person, we tune our breath, voice, and posture. This tuning is controlled by the endocrine system and the autonomic nervous system and happens outside of our conscious awareness. The polyvagal perspective on yoga effects is particularly useful in trauma therapy (Price et al., 2017; Rhodes et al., 2016; van der Kolk et al., 2014). Traumatic experiences undermine the basic sense of safety. At the physiological level, trauma may result in long-term dysregulation of the autonomic neural

130

 Neurocognitive Mechanisms of Yoga

system (Sahar et al., 2001), characterised by increased arousal and reduced vagal regulation. By engaging the ventral vagal pathway, one can bring their mind and body into a state of connectedness, safety, and calm. Yoga can be seen as an active “neural exercise” that affects the psychophysiological states, fostering adaptability, self-regulation, and resilience (Sullivan et al., 2018). Box 4. Safety and connection in yoga therapy As yoga teachers and yoga therapists, we can cultivate the ‘state of connection’ in the class/ therapy session.          1. Safety. Create a safe environment. Think about the space where you teach (light, colours, scent, curtains open or close, placing the yoga mats in a circle, etc.), your appearance, language and tone of voice etc.          2. Invite. The class/ session is based on compassion and is an invitation to learn how to notice what feels good for a client.          3. Intention. Start with the cultivation of an intention. Invite the qualities of connectedness and acceptance.          4. Mantra. Integrate mantra in the session, practised by itself or integrated into asana. Chanting together may become a ritual and develop a shared intention. Keep it simple when choosing a mantra (together) and use a mantra that connects to the desired feeling. Mantra can lengthen the exhalation.          5. Breath. Teach breath awareness. Slow breathing with emphasis on exhalation is the most powerful and accessible tool for shifting physiological states.          6. Ritual. A ritual not only creates the repetition needed for personal change but also creates a bonding with you as a therapist with the client or clients as a group when practised together.          7. Touch. Integrate elements of gentle self-touch or self-massage.          8. Body awareness. Encourage students to listen to their bodies and learn how to notice physical and mental signals in their practices. Body awareness is the key to regulating physiological states.          9. Attitude. Normalise any experience of a client and listen with curiosity, patience, open attention and without judgment. Showing a person that you believe in recovery and healing, even when a client does not believe him/ herself can be a very connecting act.          10. Boundaries. Set healthy boundaries. Without boundaries, compassionate, healthy (therapeutic-) relationships are not possible.          11. Social. Foster social bonding between you and the client and the clients with each other, for example, by creating space for tea and chats after practice.          12. Self-practice. Practice yoga and self-compassion yourself as a yoga therapist teacher/yoga therapist.

Summary: The Neurocognitive Foundations of Yoga and Meditation As we show in the sections above, much is known about the effects of meditation on the brain and cognitive functions. At the same time, the impact of yoga goes beyond the top-down cognitive effects of meditation. In yoga, we approach the human being holistically at many levels of being. “When we hear the word ‘heart’ in ancient yoga texts and the Upanishads, it does not refer to the physical organ pumping away inside our chests. Instead, it refers to space within the centre of the chest, space where emotions seem to resonate, and the centre of our identities seems to reside in our bodies. That’s where we point when we say “I.” (Mohan & Mohan, 2018b). The latest research shows a paradigm shift in Western science — we see now that the mind and body form a single system and are inseparable (Azzalini et al., 2019). Our cognition is formed by the interplay between the brain and the bodily senses, and our heart can really, and not just metaphorically, carry emotions and the sense of self. More fundamental research in neuroscience is needed to extend our understanding of the mind-body connection, and its role in the generation of the conscious perception, cognition, emotions, and self. This knowledge will pave the way for holistic approaches in healthcare — such as yoga and yoga therapy.

131

 Neurocognitive Mechanisms of Yoga

CASE-STUDY: YOGA THERAPY IN PSYCHOSIS Psychosis and Schizophrenia: A General Overview Schizophrenia is a chronic multi factorial mental condition that affects about 1 percent of people (Freedman, 2003; Picchioni & Murray, 2007). Symptoms of schizophrenia typically emerge in adolescence and early adulthood. There are what we call “negative” and “positive” symptoms. Negative symptoms refer to those that consist of a decrease or absence of function, including flat expressions or little emotion, poverty of speech, inability to experience pleasure, lack of desire to form relationships, and lack of motivation. Positive symptoms are those that consist of increased or distorted function, including symptoms associated with psychosis such as hallucinations. The term “positive symptoms” can be confusing because the word “positive” gives the impression that it concerns symptoms that are pleasant. This is not the case. In the case of psychosis, one can suffer from positive symptoms. The word “positive” refers to the fact that these symptoms are phenomena that, in addition to normal functioning, have been added extra. These are experiences that were not there at first but were added to it. Psychosis is one of the main manifestations of schizophrenia. Psychosis is a mix (syndrome) of different types of symptoms. The mix looks different for everyone. For example, one person has feelings of distrust; the other hears hostile voices. Another person suffers from mood swings and is deeply gloomy and slow (depression) one time and abnormally excited and hyperactive (manic) at other periods. Still, other people experience insufficient motivation to perform daily tasks (motivation problems), lack of attention and concentration and are less able to plan and learn (cognition problems). Psychotic experiences may also include delusions or disorganised behaviour or thinking, when thoughts move quickly from idea to idea, making links between things that other people do not make. Sometimes individuals see, hear, smell, taste or feel things that others do not. Hallucinations often take the form of auditory hallucinations, or “hearing voices”. These experiences differ for individuals; some may hear a running commentary, voices from a particular memory, or non-verbal auditory hallucinations. The content of these auditory hallucinations also varies, ranging from hearing positive and pleasant comments to the very negative and abusive. They can often seem all-knowing, powerful and persuasive and may be understood as belonging to people that are known, e.g. religious figures, friends, relatives). When psychotic experiences become more intense, persist longer and together with depression, mania, or with changes in motivation or cognition, problems can arise. We speak of a full psychosis when psychotic experiences start to dominate a person’s life so much that they no longer function well in daily life. They are completely absorbed in their reality and can no longer distinguish between delusion and reality. For them, delusions, voices or hallucinations in psychosis can be as real as life. Thinking follows the logic of psychosis. The environment often thinks that they are acting strangely and want to get help. The person experiences this very differently in their psychosis: they feel that the environment has changed and that other people act strangely. This feels threatening, and a ‘clash’ with the world around them is then inevitable. Psychosis has no single cause, and the factors that lead to psychosis differ from person to person. These may be a combination of genetic, biological, psychological, social and environmental factors. Psychotic experiences may be associated with trauma and adverse childhood experiences (ACE’s). Trauma and psychosis can be interrelated in different ways (van den Berg et al., 2018): i) psychosis and PTSD can

132

 Neurocognitive Mechanisms of Yoga

occur together ii) psychosis can have a traumatic meaning for the person iii) traumas contribute directly to psychosis iv) trauma contribute to psychosis indirectly. The symptoms of psychosis and schizophrenia can severely affect the quality of life quite a lot, as can the social stigma attached to the condition. Treatment for schizophrenia seeks to give a person back their ability to function. Although current treatments provide control of, rather than a cure for, schizophrenia, long-term hospitalisation is generally not required, and one’s prognosis is much better than traditionally assumed. The general treatment for schizophrenia is medication, which often must be taken for a very long time. Unfortunately, the medication often comes with numerous side effects, and it’s hard to find the optimal balance of medication. Common side effects include stiffness and shakiness in the body, restlessness, weight gain, and movements of the jaw, lips, and tongue. Fatigue and constipation are common too.

Yoga and Mindfulness in Treatment of Schizophrenia The use of yoga in the complementary treatment of schizophrenia still remains in its early phase, but some recent reports indicate that it might be beneficial (Li et al., 2018; Vogel et al., 2019; Vancampfort et al., 2012). For example, mind-body practices that integrate physical movement with mindful attention show a more significant effect in negative symptoms than only aerobic exercise (Li et al., 2018; Vogel et al., 2019). At the same time, the positive effects of mindfulness-based interventions on psychosis, relative to no-treatment, have been reported (Goldberg et al., 2018). Meditation-based interventions for psychosis emphasise the use of mindful awareness of psychotic experiences (Chadwick et al., 2016; Wang et al., 2016; Chadwick, 2019). At the heart of mindfulness for psychosis is a humanising therapeutic process: the intention is to reduce the distress associated with psychotic experiences and increase acceptance both of psychotic experience and the self (Chadwick, 2019). In the sections below, we will discuss the application of yoga therapy in a client experiencing psychotic symptoms. We will look at psychosis from different angles and with awareness to normalising the experience as a destigmatising approach in treatment. Based on the previous sections, we choose the yoga techniques that may help the client on the way to recovery and introduce a positive change in the client’s life through the mechanisms discussed in the first part of this chapter.

Case Andrew, 24, is a young man from Amsterdam, The Netherlands who grew up in Aruba (Dutch Antilles). His parents divorced when he was seven years old. High school on the island was difficult for Andrew because he was bullied and had shown aggression at school, which resulted in dropping out eventually. He is still very angry about this. After he dropped out, he started to experiment with marijuana. He did manage to finish his degree at another school and came to The Netherlands to follow college education (programming / ICT). In the Netherlands, he moved in with his grandmother. It was hard for Andrew to connect with others; he stayed home a lot playing video games and smoking weed. He loved to write raps and draw. He did get into college, but six months in, his father died unexpectedly due to heart failure. This loss hit Andrew very hard. Shortly after this, he experienced his first psychotic episode, for which he was admitted to a psychiatric ward for three weeks. He was diagnosed with psychotic disorder NAO and set on medications. He was already heavy, but the drug made him gain weight, and obesity became a real issue after a couple of months: he would have weighed 100 kg at the age of 17.

133

 Neurocognitive Mechanisms of Yoga

Now - seven years later - he still uses antipsychotic medication. He weighs 144 kg, and the psychiatrist changed his diagnosis into schizophrenia. He is still feeling upset and angry about the things that happened to him in life. He does not agree with his diagnoses; he does not think he is sick; he does not want help. He does take his medication because his grandmother tells him he has to. He still lives with her. He has not started his study yet and has different jobs all the time. He gets fired because he struggles with showing up on time and doing the task. He seems to be in a loop of negativity, and there is not much he enjoys. He still draws, alone in his room and listens to rap music. Twice a year, he speaks to his psychiatrist, and once a week he visits the day-activity centre where they offer art activities, music, and yoga classes. He does not want to join a group class, but when offered private sessions first, he is willing to give it a try. He now has a weekly yoga therapy session. Step 1: Assessment We assess the client on many levels to work on the optimal balance. In the first one to three meetings, we talk, and we practice yoga. During these sessions, we collect information with the most important question we need to answer: what does the client wish? What does balance mean for him? Very often (especially when people felt disappointed and hurt by others in the past) it takes a while before they feel safe enough to figure out what they want in life and this therapy. It could be that we start with something, and after a couple of sessions, when we feel a safer relationship is established, we reflect on the first ideas before we go further with the goals. The above list (Table 1) lists the qualities and functions that one can influence in the yoga practice. This list is not complete; each subject could be discussed with the client. We may add more specific or indirect assessments based on disorders, i.e., particular functions that are impaired such as blurred vision, bowel and bladder urgency etc. You may be able to change these qualities directly, or indirectly – these topics can be discussed with the client, creating a personalised roadmap towards optimal balance. Step 2: Setting therapy goals When choosing goals, we may use the yogic concept of optimal balance (Sattva) in a way that fits the person. We seek goals that are sattvic for a person, not for the therapist. For example, from the therapist’s perspective, potential client goals may include: to lower stress and anxiety, learn to regulate stress with yoga tools; improve sleep, improve memory or focus. In reality, we see that clients often have different goals (for example ‘get back to work’ or ‘find a girlfriend’. The potential of yoga therapy is directly or indirectly related to the client’s goal. This connection should be discussed with the client. Step 3: Choosing practices As yoga therapists, we choose practices from a broad range of options. We consider all yogic practices that are accessible to the person. Whichever technique we choose, we teach with the attitude that is optimistic, hopeful, compassionate, respectful and empowering. The therapist’s goal is to empower a person to find their practices and solutions for sustainable self-care. Another essential aspect is normalising the disorder, which means contextualising the physical symptoms on a broad spectrum, seeing them in the context of the client’s life events and acknowledging them as coping mechanisms.

134

 Neurocognitive Mechanisms of Yoga

Table 1. Example of holistic assessment criteria Physical body           Physical fitness                       Strength                       Stamina                       Flexibility                       Range of Motion                       Endurance                       Balance           Energy level           Pain           Weakness           Breathing Mind           Feel calm           Feel content           Feel relaxed           Feel ready to engage with the world Thoughts           Think clearly           Set intentions and goals           Concentrate on work/tasks           Have an awareness of thoughts and feelings Rhythms           Sleep           Hunger           Digestion           Elimination Relationships           Able to engage in supportive relationships Identity           Feel compassion for self and others           Feel positive about the future           Have a clear vision/mission in life

The key components in Andrew’s therapy would be the experience of safety in body/self and positive experience of yoga intervention. The basic state of safety (in the environment (e.g. the room), and the body), allows us to work with the client on being aware of and connecting to his body—and all their experiences—as a first step in making positive changes such as cultivating compassion, patience, and acceptance toward himself. Then, when awareness becomes a little easier, and trust is gained, we

135

 Neurocognitive Mechanisms of Yoga

might move on to observing the body and all experiences, thoughts, and behaviour while we practice if this feels safe enough. It is essential to constantly evaluate the client’s sense of safety and go back to a “safer” place if needed. Box 5. The Sattva, optimal balance           Lasting positive change in body and mind in yoga arises from shifting the base model of the mind and body toward sattva — lightness in the body and clarity in mind. The sattvic state entails the natural state of being satisfied, with clarity and wisdom. The work is to bring the focus on the sattvic qualities and bring those sattvic qualities slowly in life and practice.           “To ensure that our practice is relevant to us, we must ensure that we stay mindful of the goal. Not only should the practice feel good, but it must also lead us toward a more balanced and positive state of body and mind, one of lightness and clarity. Moving the body reduces heaviness and dullness, comfortably long and smooth breathing reduces the restlessness in the body and mind.           As this state of mind emerges, we endeavor to deepen and extend it as the foundation for deeper practices of pranayama and meditation, and as the support for a healthy body and a calm mind. This is the basis of asana practice for positive transformation rather than only feeling good at the moment.” (Mohan & Mohan, 2018a).

Results After six months of weekly sessions, Andrew can observe himself, his actions and struggles with increased awareness. Yoga helps Andrew to manage his experiences. It seems that accepting and managing the experiences, instead of trying to push them away, decreases the intensity of them. Andrew reports feeling a bit lighter after a session as if the experiences have “dissolved” a little bit. He gradually begins to practice small exercises from the session outside of class. These techniques help him to make small changes in the patterns of daily life. He’s been motivated to change his eating habits, which has resulted in losing some weight, which was something he wanted to do, and it has made him feel more confident. Andrew’s new, healthier habitual patterns include hope, trust, and self-acceptance.

CONCLUSIONS The story of Andrew shows how the elements of yoga therapy can be applied in practice. Importantly, the described therapeutic approach combines insights from the neuroscientific research and the traditional vision of yoga as the personal transformation path. As we see, yoga practice entrains attentional capacities, such as meta-awareness, cognitive flexibility, and non-reactivity, which helps the client to reduce stress and accept their experiences. Yoga uniquely trains body awareness, which, as we showed, may affect the emotional state and the persons’ attitude towards oneself and others. The subtle shift in the physiological responses through yoga practice creates a state of safety, connectedness, and content. The most important, however, is the positive change in the patterns of thinking and behaviour. Andrew’s case shows the value of yoga therapy, as a unique holistic approach with a focus on optimal balance and empowerment. In the final part of this chapter, we would like to suggest possible directions for future research and implementation of yoga therapy, as a complementary modality in mental healthcare.

136

 Neurocognitive Mechanisms of Yoga

PERSPECTIVE: IMPLEMENTATION OF YOGA THERAPY IN MENTAL HEALTHCARE Substantial evidence from randomised controlled trials indicates that mindfulness and meditation-based interventions (MMBI) can be beneficial for patients experiencing symptoms of depression, anxiety, pain, and post-traumatic stress and psychosis. The effects size reported in randomised clinical studies of meditation-based interventions is comparable to those of established treatments (Wielgosz et al., 2019). The positive effects of MMBI’s are attributed to the developing of the core attentional capacities (meta-awareness, cognitive flexibility, non-reactivity), adaptive cognitive strategies (acceptance, positive reappraisal), and pro-social emotional state (non-judgment, compassion, self-compassion). Yoga practice is similar to MMBI’s in that it emphasises the skills of meta-awareness, kindness, and compassion. At the same time, the physical exercise component adds an extra dimension to the yoga intervention, compared to MMBI’s. The asana practice not only improves physical health and fitness but, as we have seen above, may augment the positive psychological effects of meditation. Body awareness is central in psychopathological conditions. By improving body awareness through yoga practice, the client builds a more accepting attitude towards her physical and emotional experiences. She learns to use the tools of yoga to achieve a subtle positive change in the physiological and psychological states. Physical practice helps the client to establish positive and trusting relationships with the body and the self. Yoga is a practice of personal transformation: a positive change happens when one is ready to shift their attitude and behaviour. Yoga therapy empowers and supports individuals to manage their health using accessible practices and principles of yoga (AAYT, 2019). Instead of curing a medical diagnosis, yoga therapy encompasses the wellbeing of the whole person, and the client becomes an active participant in the recovery process. Yoga therapy assists the client to obtain optimal balance through healthy behaviour and sustainable self-care. In the future, more high-quality methodologically rigorous research addressing the effects of yoga on different aspects of physical and mental health is warranted. We need more longitudinal and crosssectional studies that look into the effects of individual components of yoga and their combinations. Yoga practice is still most prevalent and most well researched among highly educated, generally healthy, middle-aged women (Cramer et al., 2019). Thus, applicability in different sociodemographic groups also requires more research. Finally, the effective implementation of yoga therapy in clinical settings requires resolving significant practical challenges and questions related to intervention format, protocols, educational and quality standards. Yoga therapy for mental health conditions is uniquely personalised, with an emphasis on empowerment and making the client an active participant in the recovery process. Yoga therapists do not put diagnostic labels but rather see the client as a whole and work with their needs and complaints, looking for the state of optimal health and balance. Still, clear guidelines for each step of the intervention have to be formulated. Another relevant question is the applicability of ethical and spiritual elements of yoga, which are essential components that motivate regular practitioners to maintain yoga practice (Park et al., 2019), in the therapy of mental health conditions. These questions need to be addressed for yoga therapy to become a profession, recognised by institutions and professionals in other fields, such as medicine, mental health, physical therapy, social welfare, and education.

137

 Neurocognitive Mechanisms of Yoga

REFERENCES Australian Association of Yoga Therapists. (2019). http://yogatherapy.org.au/about-yoga-therapy/ Azzalini, D., Rebollo, I., & Tallon-Baudry, C. (2019). Visceral signals shape brain dynamics and cognition. Trends in Cognitive Sciences, 23(6), 488–509. doi:10.1016/j.tics.2019.03.007 PMID:31047813 Bozhilova, N. S., Michelini, G., Kuntsi, J., & Asherson, P. (2018). Mind wandering perspective on attention-deficit/hyperactivity disorder. Neuroscience and Biobehavioral Reviews, 92, 464–476. doi:10.1016/j. neubiorev.2018.07.010 PMID:30036553 Brandmeyer, T., Delorme, A., & Wahbeh, H. (2019). The neuroscience of meditation: Classification, phenomenology, correlates, and mechanisms. Progress in Brain Research, 244, 1–29. doi:10.1016/ bs.pbr.2018.10.020 PMID:30732832 Chadwick, P. (2019). Mindfulness for psychosis: A humanising therapeutic process. Current Opinion in Psychology, 28, 317–320. doi:10.1016/j.copsyc.2019.07.022 PMID:31382104 Chadwick, P., Strauss, C., Jones, A.-M., Kingdon, D., Ellett, L., Dannahy, L., & Hayward, M. (2016). Group mindfulness-based intervention for distressing voices: A pragmatic randomised controlled trial. Schizophrenia Research, 175(1-3), 168–173. doi:10.1016/j.schres.2016.04.001 PMID:27146475 Christoff, K., Irving, Z. C., Fox, K. C., Spreng, R. N., & Andrews-Hanna, J. R. (2016). Mind-wandering as spontaneous thought: A dynamic framework. Nature Reviews. Neuroscience, 17(11), 718–731. doi:10.1038/nrn.2016.113 PMID:27654862 Clark, D., Schumann, F., & Mostofsky, S. H. (2015). Mindful movement and skilled attention. Frontiers in Human Neuroscience, 9, 297. doi:10.3389/fnhum.2015.00297 PMID:26190986 Cotier, F. A., Zhang, R., & Lee, T. M. C. (2017). A longitudinal study of the effect of short-term meditation training on functional network organization of the aging brain. Scientific Reports, 7(1), 598. doi:10.103841598-017-00678-8 PMID:28377606 Craig, A. D. (2002). How do you feel? Interoception: The sense of the physiological condition of the body. Nature Reviews. Neuroscience, 3(8), 655–666. doi:10.1038/nrn894 PMID:12154366 Cramer, H., Quinker, D., Pilkington, K., Mason, H., Adams, J., & Dobos, G. (2019). Associations of yoga practice, health status, and health behavior among yoga practitioners in Germany—Results of a national cross-sectional survey. Complementary Therapies in Medicine, 42, 19–26. doi:10.1016/j. ctim.2018.10.026 PMID:30670242 Dahl, C. J., Lutz, A., & Davidson, R. J. (2015). Reconstructing and deconstructing the self: Cognitive mechanisms in meditation practice. Trends in Cognitive Sciences, 19(9), 515–523. doi:10.1016/j. tics.2015.07.001 PMID:26231761 Delorme, A., & Brandmeyer, T. (2019). When the meditating mind wanders. Current Opinion in Psychology, 28, 133–137. doi:10.1016/j.copsyc.2018.12.006 PMID:30654311

138

 Neurocognitive Mechanisms of Yoga

Engen, H. G., Bernhardt, B. C., Skottnik, L., Ricar, M., & Singer, T. (2018). Structural changes in socioaffective networks: Multi-modal MRI findings in long-term meditation practitioners. Neuropsychologia, 116, 26–33. doi:10.1016/j.neuropsychologia.2017.08.024 PMID:28842274 Engert, V., Kok, B. E., Papassotiriou, I., Chrousos, G. P., & Singer, T. (2017). Specific reduction in cortisol stress reactivity after social but not attention-based mental training. Science Advances, 3(10), e1700495. doi:10.1126ciadv.1700495 PMID:28983508 Farb, N. A. S., Segal, Z. V., & Anderson, A. K. (2013). Mindfulness meditation training alters cortical representations of interoceptive attention. Social Cognitive and Affective Neuroscience, 8(1), 15–26. doi:10.1093can/nss066 PMID:22689216 Fox, K. C. R., Dixon, M. L., Nijeboer, S., Girn, M., Floman, J. L., Lifshitz, M., Ellamil, M., Sedlmeier, P., & Christoff, K. (2016). Functional neuroanatomy of meditation: A review and meta-analysis of 78 functional neuroimaging investigations. Neuroscience and Biobehavioral Reviews, 65, 208–228. doi:10.1016/j.neubiorev.2016.03.021 PMID:27032724 Freedman, R. (2003). Schizophrenia. The New England Journal of Medicine, 349(18), 1738–1749. doi:10.1056/NEJMra035458 PMID:14585943 Gamaiunova, L., Brandt, P.-Y., Bondolfi, G., & Kliegel, M. (2019). Exploration of psychological mechanisms of the reduced stress response in long-term meditation practitioners. Psychoneuroendocrinology, 104, 143–151. doi:10.1016/j.psyneuen.2019.02.026 PMID:30849720 Germer, C., & Neff, K. (2019). Teaching the mindful self-compassion program: A Guide for Professionals. Guilford Publications. Gerritsen, R. J. S., & Band, G. P. H. (2018). Breath of life: The respiratory vagal stimulation model of contemplative activity. Frontiers in Human Neuroscience, 12, 397. doi:10.3389/fnhum.2018.00397 PMID:30356789 Gilbert, P. (2019). Explorations into the nature and function of compassion. Current Opinion in Psychology, 28, 108–114. doi:10.1016/j.copsyc.2018.12.002 PMID:30639833 Goldberg, S. B., & Tucker, R. P. (2020). Allegiance effects in mindfulness-based interventions for psychiatric disorders: A meta-re-analysis. Psychotherapy Research, 30(6), 753–762. doi:10.1080/1050330 7.2019.1664783 PMID:31506031 Goldberg, S. B., Tucker, R. P., Greene, P. A., Davidson, R. J., Wampold, B. E., Kearney, D. J., & Simpson, T. L. (2018). Mindfulness-based interventions for psychiatric disorders: A systematic review and meta-analysis. Clinical Psychology Review, 59, 52–60. doi:10.1016/j.cpr.2017.10.011 PMID:29126747 Hepark, S., Janssen, L., de Vries, A., Schoenberg, P. L. A., Donders, R., Kan, C. C., & Speckens, A. E. M. (2019). The efficacy of adapted mbct on core symptoms and executive functioning in adults with ADHD: A preliminary randomized controlled trial. Journal of Attention Disorders, 23(4), 351–362. doi:10.1177/1087054715613587 PMID:26588940 Krishnamurti, J. (2002). Meditations. Shambhala Publications.

139

 Neurocognitive Mechanisms of Yoga

Lao, S. A., Kissane, D., & Meadows, G. (2016). Cognitive effects of MBSR/MBCT: A systematic review of neuropsychological outcomes. Consciousness and Cognition, 45, 109–123. doi:10.1016/j. concog.2016.08.017 PMID:27580462 Li, J., Shen, J., Wu, G., Tan, Y., Sun, Y., Keller, E., Jiang, Y., & Wu, J. (2018). Mindful exercise versus non-mindful exercise for schizophrenia: A systematic review and meta-analysis of randomized controlled trials. Complementary Therapies in Clinical Practice, 32, 17–24. doi:10.1016/j.ctcp.2018.04.003 PMID:30057047 Lindsay, E. K., & Creswell, J. D. (2017). Mechanisms of mindfulness training: Monitor and acceptance theory (MAT). Clinical Psychology Review, 51, 48–59. doi:10.1016/j.cpr.2016.10.011 PMID:27835764 Lucas, A. R., Klepin, H. D., Porges, S. W., & Rejeski, W. J. (2018). Mindfulness-based movement: A polyvagal perspective. Integrative Cancer Therapies, 17(1), 5–15. doi:10.1177/1534735416682087 PMID:28345362 Lutz, A., Slagter, H. A., Dunne, J. D., & Davidson, R. J. (2008). Attention regulation and monitoring in meditation. Trends in Cognitive Sciences, 12(4), 163–169. doi:10.1016/j.tics.2008.01.005 PMID:18329323 Mohan, A. G., & Mohan, G. (2018a). Yoga should lead to positive changes. https://www.thehindu.com/ society/history-and-culture/yoga-should-lead-to-positive-changes/article22825726.ece Mohan, A. G., & Mohan, G. (2018b). Put some heart into it. https://www.thehindu.com/society/historyand-culture/put-some-heart-into-it/article23602401.ece Neff, K., & Germer, C. (2018). The mindful self-compassion workbook: a proven way to accept yourself, build inner strength, and thrive. Guilford Publications. Park, C. L., Quinker, D., Dobos, G., & Cramer, H. (2019). Motivations for adopting and maintaining a yoga practice: A national cross-sectional survey. Journal of Alternative and Complementary Medicine (New York, N.Y.), 25(10), 1009–1014. doi:10.1089/acm.2019.0232 PMID:31460773 Picchioni, M. M., & Murray, R. M. (2007). Schizophrenia. British Medical Journal, 335(7610), 91–95. doi:10.1136/bmj.39227.616447.BE PMID:17626963 Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74(2), 116–143. doi:10.1016/j. biopsycho.2006.06.009 PMID:17049418 Porges, S. W. (Ed.). (2011). The Polyvagal Theory: Neurophysiological foundations of emotions, attachment, communication, and self-regulation (norton series on interpersonal neurobiology). W. W. Norton & Company. Price, M., Spinazzola, J., Musicaro, R., Turner, J., Suvak, M., Emerson, D., & van der Kolk, B. (2017). Effectiveness of an extended yoga treatment for women with chronic posttraumatic stress disorder. Journal of Alternative and Complementary Medicine (New York, N.Y.), 23(4), 300–309. doi:10.1089/ acm.2015.0266 PMID:28121466 Quattrocki, E., & Friston, K. (2014). Autism, oxytocin and interoception. Neuroscience and Biobehavioral Reviews, 47, 410–430. doi:10.1016/j.neubiorev.2014.09.012 PMID:25277283

140

 Neurocognitive Mechanisms of Yoga

Rhodes, A., Spinazzola, J., & van der Kolk, B. (2016). Yoga for adult women with chronic ptsd: A longterm follow-up study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 22(3), 189–196. doi:10.1089/acm.2014.0407 PMID:26863321 Russell, T. A., & Arcuri, S. M. (2015). A Neurophysiological and neuropsychological consideration of mindful movement: Clinical and research implications. Frontiers in Human Neuroscience, 9, 282. doi:10.3389/fnhum.2015.00282 PMID:26074800 Sahar, T., Shalev, A. Y., & Porges, S. W. (2001). Vagal modulation of responses to mental challenge in posttraumatic stress disorder. Biological Psychiatry, 49(7), 637–643. doi:10.1016/S0006-3223(00)010453 PMID:11297721 Sherman, K. J. (2012). Guidelines for developing yoga interventions for randomized trials. Evidence-Based Complementary and Alternative Medicine, 2012, 143271. doi:10.1155/2012/143271 PMID:23082079 Singer, T., & Engert, V. (2018). It matters what you practice: Differential training effects on subjective experience, behavior, brain and body in the ReSource Project. Current Opinion in Psychology, 28, 151–158. doi:10.1016/j.copsyc.2018.12.005 PMID:30684917 Sullivan, M. B., Erb, M., Schmalzl, L., Moonaz, S., Noggle Taylor, J., & Porges, S. W. (2018). Yoga therapy and polyvagal theory: The convergence of traditional wisdom and contemporary neuroscience for self-regulation and resilience. Frontiers in Human Neuroscience, 12, 67. doi:10.3389/fnhum.2018.00067 PMID:29535617 Tang, Y.-Y., Hölzel, B. K., & Posner, M. I. (2015). The neuroscience of mindfulness meditation. Nature Reviews. Neuroscience, 16(4), 213–225. doi:10.1038/nrn3916 PMID:25783612 Tyng, C. M., Amin, H. U., Saad, M. N. M., & Malik, A. S. (2017). The influences of emotion on learning and memory. Frontiers in Psychology, 8, 1454. doi:10.3389/fpsyg.2017.01454 PMID:28883804 Vallat, R., Chatard, B., Blagrove, M., & Ruby, P. (2018). Correction: Characteristics of the memory sources of dreams: A new version of the content-matching paradigm to take mundane and remote memories into account. PLoS One, 13(2), e0193440. doi:10.1371/journal.pone.0193440 PMID:29466438 van den Berg, D., de Bont, P. A. J. M., van der Vleugel, B. M., de Roos, C., de Jongh, A., van Minnen, A., & van der Gaag, M. (2018). Long-term outcomes of trauma-focused treatment in psychosis. The British Journal of Psychiatry, 212(3), 180–182. doi:10.1192/bjp.2017.30 PMID:29436320 van der Kolk, B. A., Stone, L., West, J., Rhodes, A., Emerson, D., Suvak, M., & Spinazzola, J. (2014). Yoga as an adjunctive treatment for posttraumatic stress disorder: A randomized controlled trial. The Journal of Clinical Psychiatry, 75(6), e559–e565. doi:10.4088/JCP.13m08561 PMID:25004196 Vancampfort, D., Vansteelandt, K., Scheewe, T., Probst, M., Knapen, J., De Herdt, A., & De Hert, M. (2012). Yoga in schizophrenia: A systematic review of randomised controlled trials. Acta Psychiatrica Scandinavica, 126(1), 12–20. doi:10.1111/j.1600-0447.2012.01865.x PMID:22486714 Vogel, J. S., van der Gaag, M., Slofstra, C., Knegtering, H., Bruins, J., & Castelein, S. (2019). The effect of mind-body and aerobic exercise on negative symptoms in schizophrenia: A meta-analysis. Psychiatry Research, 279, 295–305. doi:10.1016/j.psychres.2019.03.012 PMID:30879703

141

 Neurocognitive Mechanisms of Yoga

Wang, L.-Q., Chien, W. T., Yip, L. K., & Karatzias, T. (2016). A randomized controlled trial of a mindfulness-based intervention program for people with schizophrenia: 6-month follow-up. Neuropsychiatric Disease and Treatment, 12, 3097–3110. doi:10.2147/NDT.S123239 PMID:27994466 Wielgosz, J., Goldberg, S. B., Kral, T. R. A., Dunne, J. D., & Davidson, R. J. (2019). Mindfulness meditation and psychopathology. Annual Review of Clinical Psychology, 15(1), 285–316. doi:10.1146/ annurev-clinpsy-021815-093423 PMID:30525995

KEY TERMS AND DEFINITIONS Attentional Meditation: Systematic training to keep attention focused on a chosen object. Autonomic Neural System: Part of the peripheral nervous system that influences internal organs and regulates bodily functions. Body Awareness: The recognition and awareness of senses related to the body position in the space and internal bodily states. Constructive Meditation: Systematic training to cultivate positive emotional qualities, such as loving-kindness, compassion, and gratitude, aiming at strengthening positive psychological patterns. Interoception: Monitoring of visceral sensations and internal bodily states. Meta-Awareness: Awareness of the processes of consciousness, including the processes of thinking, feeling and perceiving. Mindfulness: An intentional state (or skill) of alert, relaxed, non-judgmental attention onto the present moment. Mindfulness Training: A training, developed in the USA in the 1980s and adopted worldwide. The training integrates meditation, mindful movement, relaxation, and cognitive/emotional regulation strategies. Yoga Therapy: The professional application of the principles and practices of yoga to promote health and well-being within a therapeutic relationship that includes personalized assessment, goal setting, lifestyle management, and yoga practices for individuals or small groups.

142

Section 2

Yoga as a Therapeutic Intervention

144

Chapter 9

Standards in Yoga Research and Reporting Steffany Moonaz Maryland University of Integrative Health, USA Daryl Nault Maryland University of Integrative Health, USA Atiera Abatemarco Maryland University of Integrative Health, USA

ABSTRACT Yoga research is growing rapidly in volume and rigor but varies in sample size, study design, and reporting transparency. Yoga professionals may not be versed in the current research due to a lack of research literacy and may not be well positioned to discern research relevance and quality. Research literacy is necessary to apply research in yoga teaching and therapeutics. Research is part of evidence-informed practice, along with clinical experience and client preferences, and is aligned with concepts from yoga philosophy. Several strategies are available to improve research literacy and evidence-informed practice for yoga professionals, which can help to expand inter-professional collaboration and inform the trajectory of yoga research toward better alignment and application to clinical practice. Yoga research reporting guidelines will improve transparency in research dissemination for application to practice, policy, replication, comparison, and summarization.

INTRODUCTION Yoga practice is increasing worldwide. In some countries, practitioners have more than doubled in the last decade (Clarke et al., 2015) and worldwide prevalence is estimated at over 300 million individuals (Montigny, 2018). As the use of yoga expands worldwide, there is a corresponding expansion in the diversity of those who practice yoga. For example, older persons and clinical populations are two groups in which yoga practice is increasing (Clarke et al., 2015). This trend is in a cyclical relationship with yoga DOI: 10.4018/978-1-7998-3254-6.ch009

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Standards in Yoga Research and Reporting

research. As more individuals begin to practice, the need for research that verifies its safety in diverse populations also grows. Reciprocally, as research evidence grows regarding the potential effectiveness of yoga for prevention or management of various health conditions, the prevalence of participation increases. In response, published yoga research has grown exponentially in recent decades (Jeter et al., 2015). And while yoga appears to be incredibly safe in clinical trials (Cramer et al., 2015), the rate of yoga injuries is growing (Swain & Mcgwin, 2016). This disconnect may be in part due to the increase in yoga use by older and clinically complicated populations, while the training of yoga teachers worldwide remains focused on teaching relatively young and healthy students. Those delivering yoga in clinical trials, in contrast, are utilizing protocols developed specifically for a target population. The apparent disconnect may also be due to the high risk of bias in yoga research (Cramer et al., 2015). The need for yoga that is safe and appropriate for diverse populations has been met by the rise in both yoga therapy and accessible yoga. Yoga therapy, while ancient in its origins, has become more codified and professionalized in recent years, including the establishment and growth of the International Association of Yoga Therapists (n.d.), which has developed training competencies, credentialing of schools and professionals, codes of conduct and a scope of practice for the field. Accessible Yoga (n.d.), conversely, is a grassroots movement to foster awareness and access to yoga spaces, teachers, and practices by anyone, regardless of physical ability, mental health status, or demographics. Unfortunately, little research has been done to examine whether the practices of yoga therapy and accessible yoga in real world settings parallels the practices that have suggested safety and acceptability in the growing body of yoga research. The objectives of this chapter are to first identify the known challenges of bias and research literacy surrounding the current body of yoga literature. Second, these challenges are addressed more specifically within the context of study design, research reporting, and Evidence Informed Practice (EIP). Finally, solutions and recommendations are made to advise on the role the yoga research community might play in improving future yoga research and evidence informed practice.

BACKGROUND While the yoga literature has expanded rapidly in recent years, it continues to suffer from important weaknesses. Yoga research is subject to a few inherent methodological biases that are worthy of consideration when conducting studies, reading the literature, and applying it to clinical or policy decisions. One chief concern in designing a yoga study is avoiding selection bias. When researchers recruit participants for a research study, the very fact that it’s a yoga study will likely influence who might choose to participate. Yoga studies may attract people who would be likely to try yoga anyway, those who have prior experience with yoga, those who perceive yoga as within their capabilities, and those who believe they would feel welcome in a yoga space. This could lead to study samples that are younger, more mobile, and less diverse than the clinical population of interest. Observational studies that gather information about people who are already practicing yoga will also suffer from such biases. Efforts should be made by research teams to recruit diverse study samples, perhaps even with oversampling of underrepresented demographic groups, and comparisons should be made between study samples and the larger population to better understand the potential generalizability of study findings. Another form of bias that impacts the yoga research literature is reporting bias. This happens when the information provided in the published manuscript is incomplete or misleading and it can occur by intentional omission but more often it likely occurs by oversight. An example is that many yoga studies 145

 Standards in Yoga Research and Reporting

fail to report on adverse events (AEs). There is no way of knowing whether there were no AEs to report or whether the authors failed to report adverse events that occurred. Especially in large or long trials, it is likely that some minor events occurred, even if they were not necessarily related to the yoga intervention itself. If there were no related AEs or no serious AEs, the authors should make that clear. If it is assumed that no AE reporting means that no AEs occurred, there is a risk of concluding that a particular yoga intervention is safer than it may actually be. Authors may also fail to report data from all outcomes or may make statements in their manuscript without providing the data to support those statements. An example of this would be an article that reports which outcomes had a statistically significant improvement without noting those that didn’t reach statistical significance or stating that statistical significance was reached without providing enough information about how the data was analyzed. Another common reporting bias occurs when attrition rates are excluded. Those who fail to complete a study may have done so because of a lack of improvement, so including only the information about study completers would be misleading to the true results. While yoga research has gained legitimacy throughout the world and more researchers are building careers focused on yoga research, that group remains small compared to the global cadre of career researchers in other areas. Additionally, substantial gaps remain in some areas of the literature. Low back pain, for example, has received a lot of attention in the yoga literature, while cerebral palsy has very little evidence. This does not mean that yoga is effective for low back pain and not for cerebral palsy, but that there has not been enough research conducted at this point to demonstrate its effects in that population. As the saying goes, “absence of evidence is not evidence of absence.” A paucity of literature in a given area, however, can appear to various stakeholder groups as a signal of yoga’s irrelevance for that particular concern, which then impacts funding for future research. This is a prime example of the way that a lack of research literacy limits the initiation, execution, and application of yoga therapy research.

STUDY DESIGN CHALLENGES IN YOGA RESEARCH Bias may be mitigated thorough careful study design. In yoga studies that do include a control group, the selection of what the control group should do or receive is not as clear as it might be for pharmacological research. There is no yoga placebo that would include everything about yoga except its active ingredients. In many yoga studies, the control group does nothing and receives nothing beyond whatever measurement and testing the study requires. The issue here is that the yoga group is getting attention that the control group is not. Just receiving attention is enough to shift many types of outcomes (Mohammad, 2019). In such studies, it is not possible to determine whether it is the yoga that is responsible for the measured changes, or whether the same effects could have been achieved with attention alone. Ideally, the control group in a yoga study would receive some other form of attention, referred to as an equal attention control. This often takes the form of an educational group or a support group. Another strategy that is sometimes used in yoga research is the waitlist control. In this design, some participants are randomly assigned to receive yoga, while the others are told that they will get yoga too, but they must wait a while first. During the waiting period, they serve as an inactive control group. This is done to reduce attrition, because people are more likely to remain in a study with the expectation of receiving something later. It also could be considered more ethical because everyone in the study has the potential to benefit from participation. And lastly, it means that both groups expect to receive something that might make them

146

 Standards in Yoga Research and Reporting

feel better. Such expectations can be part of why outcomes improve. A waitlist control can help to control for some of those expectation benefits. Study design in yoga research faces some specific challenges. The chief among these challenges is likely sample size. In a recent review of yoga intervention studies, Bridges and colleagues (2017) found that although studies showed a positive correlation between yoga practice and improved depressive symptoms, among the studies’ limitations were small sample sizes ranging from 14-136 study subjects. Thus, future yoga therapy studies could gain more traction with larger sample sizes to better determine the true effectiveness of an intervention. Small sample sizes reduce statistical power, which is the ability to discover a difference when it truly exists, whether it is within a single group over time or between two groups in a comparison. Small sample sizes may lead to type II statistical errors where studies conclude that yoga had no effect when an effect exists. Large sample sizes require more funding and more time, both of which are rare in yoga research, which is often conducted on a shoestring budget as a researcher’s passion project. Another issue in yoga research, and other areas of complementary and integrative health, is the challenge of adequate and appropriate comparison groups. The scientific community operates largely under the paradigm of pharmacological research, in which a double-blind, randomized, placebo-controlled trial is both possible and ideal. This study design is often considered to be the gold standard in rigorous health research. This is for good reason since placebo effects, also called non-specific effects or context effects, can be as strong as the effect of the intervention itself. When someone knows they are receiving something that could help them feel better, they usually do feel better. This goes beyond subjective measures such as pain and is exemplified with objective measures such as blood pressure, which can be easily influenced by expectations and beliefs (Schedlowski et al., 2015). However, it is not clear how someone might practice yoga but not know they are practicing yoga. There have been some attempts at ‘sham yoga’ but those interventions necessarily contain some of the ‘active ingredients’ of yoga itself (Park et al., 2014). In fact, many yoga studies are designed without any comparison group, but simply follow a single group over time to see if providing yoga results in any changes. This is not only problematic due to expectation effects, but because we can’t be certain that the changes were due to yoga. For example, if a yoga study for people with arthritis begins in a cold weather season or a rainy season and ends in a warm or dry season, many study participants will feel a reduction in their symptoms due to the change in weather, regardless of the yoga. In such a scenario, it is difficult to determine how much of the effect was weather and how much was yoga. Even an inactive comparison group, which is a group who does not receive anything, would help control for these effects. Evidence suggests that the therapeutic potential of placebo effects contribute to psychobiological changes that are clinically relevant for both healthy individuals and those with various medical conditions (Finniss et al., 2010). Placebo effects may be better understood as the effect of expectations and beliefs on clinical outcomes. Because yoga has become popularized around the globe for its role in health and wellness, participants enrolled in a yoga trial (who can’t be adequately blinded to the fact that they are receiving yoga), may have expectations of improved physical health. For those practicing yoga and clinicians who work with them, the additive or synergistic effects of yoga and placebo can be beneficial and may be fostered for better outcomes, but in research studies, the placebo is intended to ensure focused measurement of the “active ingredients” in the treatment under investigation.

147

 Standards in Yoga Research and Reporting

YOGA RESEARCH REPORTING QUALITY In addition to challenges with study design, yoga research has been plagued by inadequate reporting in the medical literature (Cramer et al., 2015). This means that when yoga research is published in peerreviewed journals, it is often lacking important detail about the yoga intervention itself. These details include: the dose of yoga (intensity, duration, and frequency of practices), what specific practices were used, and how they may have been modified. Other aspects of yoga research are often not clearly reported, including outcome measurement, analysis procedures, participation rates, or adverse events, as previously mentioned. Inadequate reporting of yoga research may provide an inaccurate assessment of yoga’s safety, but also reduces transparency about what yoga practices were associated with the reported findings. This impedes the ability of yoga professionals (yoga teachers and yoga therapists) to consider yoga research and engage in evidence-informed practice (see below) when teaching and providing clinical care. Inadequate reporting also limits scientists from conducting effective systematic reviews that combine the findings of multiple clinical trials and present the current state of the research for yoga in any given population. This is especially important when, as mentioned previously, many yoga studies are small and may have inadequate statistical power to show significant findings, however, these small studies may be combined to suggest stronger overall evidence. Studies that are not well reported also cannot be replicated to confirm findings, nor can scientists engage in comparative effectiveness research to establish best practices in yoga teaching and yoga therapy. One of the primary purposes of biomedical research is to better inform health practice and application. Even methodologically sound studies will fall short of this task if they are not reported clearly enough for others to follow. “High-quality research reports contribute to more efficient translations of research findings into clinical practice and help advance scientific knowledge and patient care” (Moher et al., 2010). Prior research suggests that the use of reporting guidelines may also improve reporting quality so adherence to research reporting guidelines could stand to improve the quality and usefulness of yoga research literature (Cramer et al., 2015). Additionally, improved research reporting is associated with greater application of research findings in the form of evidence-informed practice (Simera et al., 2010), which results in better clinical care and improved clinical outcomes. To put it simply, yoga professionals can only apply the research literature to the extent that the literature is clearly reported. Transparent reporting, however, requires commonly accepted and widely used guidelines to ensure that all authors can follow the same standards.

YOGA RESEARCH REPORTING GUIDELINES Reporting biases may be reduced through clear and consistent reporting. The EQUATOR (Enhancing the Quality and Transparency Of health Research) Network was created in response to the need for complete and transparent reporting standards in health research. An international initiative, the EQUATOR Network is focused on improving the quality of health research reporting by promoting clear reporting practices and broader uptake of reporting guidelines (Turner et al., 2012). The EQUATOR website (n.d.) contains a collection of reporting guidelines and online resources at no cost (Simera et al., 2010). It is important to note that these reporting guidelines were not created to tell researchers how to conduct research. They may however, be useful in the study design process. Reporting guidelines typically state the minimum information required for a clear account of the research rationale, methods, analysis, and 148

 Standards in Yoga Research and Reporting

results. Clear reporting enables others to replicate the study or apply it to practice when appropriate. Some journals require that authors adhere to reporting guidelines like the EQUATOR Network’s, and peer-reviewers may use them to help appraise whether a manuscript is appropriate for publication (Altman & Simera, 2016). There are multiple reporting guidelines that fall under the EQUATOR Network, each one is suited to a specific study type. Four of the EQUATOR Network’s primary guidelines cover major study designations commonly seen in biomedical research. Each reporting guideline generally includes three important pieces: (1) statement on the development and updates to the guideline; (2) a corresponding checklist of items that need to be reported; (3) an explanation and elaboration (E&E) document that further clarifies the use of the checklist. When a study that randomly allocates participants to a clinical intervention or a control group (known as a Randomized Controlled Trial or RCT) has been completed, the CONSORT (Consolidated Standards of Reporting Trials) guidelines are recommended (Turner et al., 2012). These guidelines were first published in 1996 (Begg, 1996; Hopewell et al., 2008), and have since seen updates in 2010 (Schulz et al., 2010). Along with the previously mentioned three CONSORT documents (Moher et al., 2010), authors are guided to the use of a CONSORT study flow chart, which has seen significant popularity in publication since it was introduced (Altman & Simera, 2016). Studies presenting non-experimental or observational research, similarly, have the STROBE (Strengthening the Reporting of Observational Studies in Epidemiology) guidelines for reporting (vonElm et al., 2007). Case studies, which are indepth investigations of a clinical phenomenon, have their own CARE (Case Report) statement, E&E, and checklist as well (Gagnier et al., 2013; Riley et al., 2017). Finally, Systematic Reviews and MetaAnalysis studies, which systematically gather and report on the qualitative and/or quantitative summary of primary research, should be reported using the PRISMA (Preferred Reporting Items for Systematic reviews and Meta-Analyses) documents (Moher et al., 2009; Liberati et al., 2009). Extensions to reporting guidelines are especially useful for integrative health modalities, which have unique considerations for both study and practice. However, yoga researchers cannot be expected to adequately report yoga interventions in the absence of widely accepted standards for reporting of yoga research. Reporting guideline extensions exist for other complementary and integrative therapies with unique intervention features, such as acupuncture (Niemtzow, 2010), herbal interventions (Gagnier et al., 2013), and Chinese Herbal Medicine (Cheng et al., 2017). Yoga research currently lacks clear and consistent reporting guidelines, leaving the structure and delivery of yoga interventions in research unclear for readers. As noted earlier, a lack in clear guidelines limits dissemination and implementation of evidence-based yoga to practice. This absence of reporting standards in yoga research presents a need for a yoga research guideline extension, which at the time of writing this chapter, are currently being derived via Delphi panel consensus of international yoga researchers (Moonaz et al., 2019). The Delphi process uses an anonymous method for forming a consensus and can be used to determine guidelines such as, intervention protocols and the direction of future yoga research (Manincor et al., 2015). Prior to the first round of the Delphi survey to establish the Check List standardising the Reporting of Interventions For Yoga (CLARIFY), 58 items were developed by an international steering committee. In Round 1 of the survey, panelists were asked to rate those 58 items for importance of inclusion in reporting guidelines for yoga research. Pre-defined consensus levels determined item inclusion or exclusion at each round, which decided the content of subsequent survey rounds. In Round 2 of the survey, with the addition of 15 new recommended items from the panelists, 52 items were rated or re-rated by the Delphi panel. Fifteen items had not yet achieved consensus and

149

 Standards in Yoga Research and Reporting

were re-rated in Round 3. Based on the same pre-defined consensus levels, a fourth round is currently being rated by the expert panelists, which will ultimately form the final CLARIFY reporting guideline. In total, 53 panelists agreed to participate, representing 11 different countries. Panelists were primarily researchers, had at least five years research experience, and identified yoga as their primary research field. Over the course of the first three rounds, several major yoga-specific themes have emerged as important for the CLARIFY reporting extension. The three themes with the most guideline items for reporting included: (1) the dose of the intervention (frequency, intensity, timing, type, etc.), (2) instructor fidelity to the study protocol, and (3) materials provided for participant’s home practice, if any. Remaining themes include the reporting of: (4) specifics on the activities included in the intervention, (5) alterations to and rationale for protocol changes, (6) participant adherence to the intervention itself, (7) specifications for presentation of the title, (8) inclusion of theory behind the choice of yoga intervention, and (9) expertise of those creating and delivering the intervention. Creation of this yoga-specific reporting guideline extension may serve to enhance yoga research reporting across study designs. Quality reporting will in turn improve clinical application, policy recommendations, public understanding, and future research. The completion of CLARIFY is expected in December 2019 with subsequent publication anticipated in 2020.

YOGA RESEARCH IN EVIDENCE-INFORMED PRACTICE Sufficient reporting alone will not, in itself, improve evidence-informed practice and uptake of yoga research by yoga professionals. It also requires a willingness to utilize research literature in clinical decision-making. Evidence-Based Medicine, also now commonly referred to as Evidence-Based Practice or Evidence-Informed Practice (EIP), has three essential components: clinical experience, patient perspective, and best available evidence. While there is some misconception that EIP relies only on the last of these three, the first two are also fundamental to EIP in any field. These three components are also broader than they might appear at first glance. Clinical experience, for example, includes clinical expertise, training, mentorship, and texts written by experts in the field. And there is no specific formula for how each component should be applied. As with a three-legged stool, a clinician might lean more heavily into the strongest of its legs. There is some evidence, therefore, that experienced clinicians rely more heavily on their own experience, whereas newly trained clinicians might rely more heavily on the best available evidence (Karthikeyan & Pais, 2010). If the evidence is not very strong, it may play a lesser role. If the patient has strong opinions or concerns, those might take precedence. Failing to consider the best available evidence, however, is a failure to use all tools available in the care of the individual. There has been some concern among yoga providers that EIP is not in alignment with the philosophy and practice of yoga. In fact, there is alignment between EIP and the Samkhya Karika (“Verses on Samkhya”) concepts of direct perception, inference, and valid testimony (Sullivan et al., 2017). The skill of direct perception is necessary to develop clinical expertise that informs decision-making. Inference uses information from direct perception and valid testimony to better understand what is directly observed and observed by trusted others. Valid testimony includes the work of career researchers, esteemed teachers, and the clients themselves. These three components combine to assist the practitioner in developing discriminative wisdom that is the foundation of EIP. In order to utilize EIP in yoga teaching and yoga therapy, yoga providers must be equipped with basic skills and knowledge. In order to apply the best available evidence, providers must possess the research literacy to find, read, comprehend, assess and apply that evidence in real-world settings. 150

 Standards in Yoga Research and Reporting

Research literacy, like any other form of literacy, is developed through a combination of training and exposure. Because research literacy is not a part of the competencies developed by IAYT and because many yoga professionals around the world are trained outside of university settings where research is prioritized, they may have little to no formal training in research literacy. Given that fact, the research can be intimidating at best and disregarded at worst. Yoga professionals may feel uncomfortable in their exposure and understanding of the literature, and in their ability to discern what evidence is high quality and relevant to their client or student population. EIP does not require that the research be strong or robust, but that the provider has an ability to find the best of that literature and apply it whenever clinically appropriate. Training for yoga professionals, therefore, should emphasize the role of research evidence in a comprehensive toolbox for best practices in yoga. Next, there is also the problem of access. Since many yoga professionals are operating outside of formal institutions and much of the existing research literature exists behind a pay wall (meaning journal subscription or other payment is required to access the information), yoga providers may not have the affiliations or means to access and therefore discern the best available evidence. Because EIP contains three essential components, its application also requires an understanding of both expert opinion and client preference in addition to clinical knowledge. While it is outside the scope of this chapter and won’t be addressed extensively, it could be debated that yoga providers worldwide may not have breadth in their understanding of expert opinion in yoga, including both ancient and modern diverse perspectives. And while yoga therapists are generally taught the skills to determine the concerns and priorities of their clients, yoga teachers may not be trained to do the same.

RESEARCH LITERACY IN YOGA PROFESSIONALS In order to benefit the quality of client care, clinical collaboration, and the field, competencies in research literacy and evidence-informed practice have been proposed for yoga therapists. These skills are taught and evaluated in most health-related fields, including integrative health approaches such as massage therapy, acupuncture, chiropractic and naturopathic medicines, which may be considered peer professions to yoga therapy. And while yoga teachers may not utilize a plan of care in the way yoga therapists do, an understanding of current yoga research can inform their teaching and individualized recommendations to ensure student safety and incorporate accurate science into instructional approaches. The recommended competencies are listed below: General Competency Statement: Explain, evaluate, and apply scientific evidence in the context of evidence-informed practice, which includes: 1) research evidence, 2) expert opinion (i.e.: ancient text, yoga training, clinical experience), and 3) patient/client preferences; and apply evidence-informed decision-making in yoga therapy practice (Moonaz et al., 2017). 1. Explain the role of scientific evidence in yoga therapy practice. 2. Describe common research methodologies within the context of both clinical and mechanistic research, focusing on an assessment of yoga therapy. 3. Discuss contemporary issues in yoga research, including those relative to evaluating whole practices, diverse approaches, client-centered approaches and health outcomes (objective and client-reported). 4. Analyze the yoga research base including the indications and contraindications for specific practices. 5. Apply fundamental skills in research evaluation, including quality assessment tools. 151

 Standards in Yoga Research and Reporting

6. Demonstrate evidence informed decision-making in clinical care. Because research literacy and evidence-informed practice are not taught in many yoga programs, yoga professionals should seek other strategies for enhancing these skills. Pursuing formal continuing education is one such strategy. Courses in research literacy and evidence-informed practice may be found in university settings, online, and as continuing-education geared toward clinical professionals. Literacy is gained not simply through formal instruction, as any student of a second language can attest, but through ongoing exposure. Routinely reading research in the field is another strategy for enhancing literacy. Yoga professionals may choose to subscribe to peer-reviewed journals that are focused on yoga research or to seek out literature focused on their specific student or client population. To gain literacy, however, it is important to also address the language that is challenging or unfamiliar. As with learning any new language, one must dig deeper into the unfamiliar vocabulary by looking up new terms to broaden understanding. Many universities also host journal clubs, some of which may be open to the public. Journal clubs may also be found online, or self-started. A journal club operates somewhat like a book club in that everyone reads the article beforehand and then comes together to discuss it in depth, usually with a designated facilitator with related expertise. As the field of yoga therapy becomes more integrated into standard healthcare settings, it is expected that evidence-informed practice will be a consistent theme throughout training programs, interwoven with ancient philosophy, anatomy/physiology, and other core content areas. It is up to yoga professionals to dismantle misconceptions that research is somehow not ‘yogic’ and that systematic inquiry, observation, and evidence are very much aligned with yoga’s tenets. It is likely that a culture of research literacy and evidence-informed practice will take root in yoga teaching and yoga therapy as the evidence-base grows and the benefits become increasingly apparent.

INTER-PROFESSIONAL RESEARCH COLLABORATION Another important reason for yoga professionals to become research literate is to become involved in the trajectory of yoga research. If yoga research is conducted exclusively by career researchers with an interest in yoga but no formal training, studies are designed without input regarding real-world practice. For yoga research to be clinically relevant, it should be informed by practicing clinicians and educators. It may also be useful to consider the generalizability across different socio-economic classes, different cultural groups, and in different parts of the world. A yoga intervention that can easily be applied in one population may not work in another. Yoga professionals do not need to become career researchers in order to inform the trajectory of yoga research. Most yoga studies would benefit from a subject-matter expert with clinical experience who can inform decisions from the initial research proposal through the dissemination of findings. Subjectmatter expertise doesn’t require a PhD in clinical research, but it does require a basic understanding of research processes such as study designs, ethical oversights, scientific conventions, and the existing literature. Another common role for yoga professionals in the research process is that of interventionist. The interventionist delivers the yoga to study participants and, in addition to having extensive training and experience in yoga, should be familiar with concepts such as protocol fidelity, adherence measurement, adverse events reporting, and interventionist effects.

152

 Standards in Yoga Research and Reporting

For yoga professionals to effectively collaborate with career researchers, they should be familiar with various types of study designs and their variations, as well as the challenges and opportunities within each. They should understand research ethics and the ethical oversight process, as well as their own risk of bias and the main risks of bias for a particular project. They should be generally familiar with the existing literature relevant to the project underway. They should be sufficiently literate in the language of research in order to follow and contribute to discussions of the project and its role in the trajectory of yoga research. Finally, they should also have depth of knowledge and skill in yoga and its application for the particular population or concern that the project aims to address. Just as one is never finished studying yoga, because it is a lifelong endeavor, one is also never finished learning about the research process as a means to acquire generalizable knowledge and a deeper understanding of phenomena that may ultimately be applied for the greater good.

CONCLUSION Yoga research has grown exponentially in volume and rigor in recent years, though challenges remain including sufficient sample sizes, appropriate comparison groups, thorough reporting transparency, and risk of bias. Evidence-informed practice (EIP) calls for use of the best available evidence, which requires discernment by yoga providers. Emerging reporting guidelines for yoga research will be of use for discerning high quality and relevant research. The principles of EIP are very much aligned with foundational concepts in yoga philosophy and should be embraced by yoga professionals. EIP requires basic research literacy, which is often excluded from yoga education, but can be fostered through continuing education, exposure to research literature, journal clubs and other opportunities. Yoga research will be more relevant to clinical practice when yoga professionals serve as members of the research team, which requires an understanding of the research process and its challenges. Improved research literacy will also allow for improved inter professional communication and collaboration with healthcare professionals for improved patient outcomes.

ACKNOWLEDGMENT This research received no specific grant from any funding agency in the public, commercial, or not-forprofit sectors.

REFERENCES Altman, D. G., & Simera, I. (2016). A history of the evolution of guidelines for reporting medical research: The long road to the EQUATOR Network. Journal of the Royal Society of Medicine, 109(2), 67–77. doi:10.1177/0141076815625599 PMID:26880653 Balasubramaniam, M. M., Telles, S. M., & Doraiswamy, P. M. (2013). Yoga on our minds: Asystematic review of yoga for neuropsychiatric disorders. Frontiers in Psychiatry, 3, 117. doi:10.3389/ fpsyt.2012.00117 PMID:23355825

153

 Standards in Yoga Research and Reporting

Begg, C. (1996). Improving the quality of reporting of randomized controlled trials. Journal of the American Medical Association, 276(8), 637. doi:10.1001/jama.1996.03540080059030 PMID:8773637 Bridges, L., & Sharma, M. (2017). The efficacy of yoga as a form of treatment for depression. Journal of Evidence-Based Complementary & Alternative Medicine, 22(4), 1017–1028. doi:10.1177/2156587217715927 PMID:28664775 Cheng, C.-W., Wu, T.-X., Shang, H.-C., Li, Y.-P., Altman, D. G., Moher, D., & Bian, Z.-X. (2017). CONSORT extension for Chinese herbal medicine formulas 2017: Recommendations, explanation, and elaboration (Simplified Chinese Version). Annals of Internal Medicine, 167(2), W21–W34. doi:10.7326/ IsTranslatedFrom_M17-2977_2 PMID:28654957 Clarke, T. C., Black, L. I., Stussman, B. J., Barnes, P. M., & Nahin, R. L. (2015). Trends in the use of complementary health approaches among adults: United States, 2002–2012. National Health Statistics Reports, 79, 1–16. PMID:25671660 Cramer, H., Langhorst, J., Dobos, G., & Lauche, R. (2015). Associated factors and consequences of risk of bias in randomized controlled trials of yoga: A systematic review. PLoS One, 10(12), e0144125. doi:10.1371/journal.pone.0144125 PMID:26629905 Cramer, H., Ward, L., Saper, R., Fishbein, D., Dobos, G., & Lauche, R. (2015). A systematic review and meta-analysis on the safety of yoga. Integrative Medicine Research, 4(1), 81. doi:10.1016/j.imr.2015.04.119 PMID:26116216 Finniss, D. G., Kaptchuk, T. J., Miller, F., & Benedetti, F. (2010). Biological, clinical, and ethical advances of placebo effects. Lancet, 375(9715), 686–695. doi:10.1016/S0140-6736(09)61706-2 PMID:20171404 Gagnier, J. J., Boon, H., Rochon, P., Moher, D., Barnes, J., & Bombardier, C. (2006). Reporting randomized, controlled trials of herbal interventions: An elaborated CONSORT statement. Annals of Internal Medicine, 144(5), 364. doi:10.7326/0003-4819-144-5-200603070-00013 PMID:16520478 Gagnier, J. J., Kienle, G. G., Altman, D., Moher, D., Sox, H., & Riley, D. (2013). The CARE guidelines: Consensus-based clinical case reporting guideline development. Global Advances in Health and Medicine: Improving Healthcare Outcomes Worldwide, 2(5), 38–43. doi:10.7453/gahmj.2013.008 PMID:24416692 Hopewell, S., Clarke, M., Moher, D., Wager, E., Middleton, P., Altman, D. G., & Schulz, K. F. (2008). CONSORT for reporting randomized controlled trials in journal and conference abstracts: Explanation and elaboration. PLoS Medicine, 5(1), e20. doi:10.1371/journal.pmed.0050020 PMID:18215107 International Association of Yoga Therapists. (n.d.). Retrieved from https://www.iayt.org Jeter, P. E., Slutsky, J., Singh, N., & Khalsa, S. B. S. (2015). Yoga as a therapeutic intervention: Abibliometric analysis of published research studies from 1967 to 2013. Journal of Alternative and Complementary Medicine (New York, N.Y.), 21(10), 586–592. doi:10.1089/acm.2015.0057 PMID:26196166 Karthikeyan, G., & Pais, P. (2010). Clinical judgement & evidence-based medicine: Time for reconciliation. The Indian Journal of Medical Research, 132, 623–626. PMID:21150015

154

 Standards in Yoga Research and Reporting

Liberati, A., Altman, D. G., Tetzlaff, J., Mulrow, C., Gøtzsche, P. C., Ioannidis, J. P. A., Clarke, M., Devereaux, P. J., Kleijnen, J., & Moher, D. (2009). The PRISMA statement for reporting systematic reviews and meta-Analyses of studies that evaluate health care interventions: Explanation and elaboration. BMJ (Clinical Research Ed.), 339(jul21 1), b2700. doi:10.1136/bmj.b2700 PMID:19622552 Manincor, M. D., Bensoussan, A., Smith, C., Fahey, P., & Bourchier, S. (2015). Establishing key components of yoga interventions for reducing depression and anxiety, and improving well-being: A Delphi method study. BMC Complementary and Alternative Medicine, 15(1), 85. doi:10.118612906-015-0614-7 PMID:25888411 Mohammad, A., Thakur, P., Kumar, R., Kaur, S., Saini, R. V., & Saini, A. K. (2019). Biological markers for the effects of yoga as a complementary and alternative medicine. Journal of Complementary & Integrative Medicine, 16(1), 16. doi:10.1515/jcim-2018-0094 PMID:30735481 Moher, D., Hopewell, S., Schulz, K. F., Montori, V., Gøtzsche, P. C., Devereaux, P., … Altman, D. G. (2010). CONSORT 2010 Explanation and Elaboration: updated guidelines for reporting parallel group randomised trials. BMJ (Clinical research ed.), 340, c869. Moher, D., Liberati, A., Tetzlaff, J., & Altman, D. G. (2009). Preferred reporting items for systematic reviews and meta-Analyses: The PRISMA Statement. BMJ (Clinical Research Ed.), 339(jul21 1), b2535. doi:10.1136/bmj.b2535 PMID:19622551 Moher, D., Schulz, K. F., Simera, I., & Altman, D. G. (2010). Guidance for developers of health research reporting guidelines. PLoS Medicine, 7(2), e1000217. doi:10.1371/journal.pmed.1000217 PMID:20169112 Montigny, D. (2018). The Unstoppable Trend of Yoga. Retrieved July 9, 2018, from https://www.yogitimes.com/article/unstoppable-trend-yoga-infographic-business Moonaz, S., Jeter, P., & Schmalzl, L. (2017). The importance of research literacy for yoga therapists. International Journal of Yoga Therapy, 27(1), 131–133. doi:10.17761/1531-2054-27.1.131 PMID:29131733 Moonaz, S., Nault, D., Cramer, H., & Ward, L. (2019, October). Development of international yoga reporting guidelines- a Delphi survey [Paper Presentation]. Symposium on Yoga at Kripalu Center for Yoga & Health, Stockbridge, MA, USA. Niemtzow, R. C. (2010). Comments regarding the new revised standards for reporting interventions in clinical trials of acupuncture (STRICTA): Extending the CONSORT statement. Medical Acupuncture, 22(3), 155–156. doi:10.1089/acu.2010.2024 Park, C. L., Groessl, E., Maiya, M., Sarkin, A., Eisen, S. V., Riley, K., & Elwy, A. R. (2014). Comparison groups in yoga research: A systematic review and critical evaluation of the literature. Complementary Therapies in Medicine, 22(5), 920–929. doi:10.1016/j.ctim.2014.08.008 PMID:25440384 Riley, D. S., Barber, M. S., Kienle, G. S., Aronson, J. K., Schoen-Angerer, T. V., Tugwell, P., ... Gagnier, J. J. (2017). CARE guidelines for case reports: Explanation and elaboration document. Journal of Clinical Epidemiology, 89, 218–235. doi:10.1016/j.jclinepi.2017.04.026 PMID:28529185

155

 Standards in Yoga Research and Reporting

Schedlowski, M., Enck, P., Rief, W., & Bingel, U. (2015). Neuro-bio-behavioral mechanisms of placebo and nocebo responses: Implications for clinical trials and clinical practice. Pharmacological Reviews, 67(3), 697–730. doi:10.1124/pr.114.009423 PMID:26126649 Schulz, K. F., Altman, D. G., & Moher, D.CONSORT Group. (2010). CONSORT 2010 Statement: Updated guidelines for reporting parallel group randomised trials. BMJ (Clinical Research Ed.), 340(mar23 1), c332. doi:10.1136/bmj.c332 PMID:20332509 Simera, I., Moher, D., Hirst, A., Hoey, J., Schulz, K. F., & Altman, D. G. (2010). Transparent and accurate reporting increases reliability, utility, and impact of your research: Reporting guidelines and the EQUATOR Network. BMC Medicine, 8(1), 24. doi:10.1186/1741-7015-8-24 PMID:20420659 Sullivan, M., Finlayson, D., & Moonaz, S. (2017, Summer). Understanding yoga’s roots in evidence informed practice. Yoga Therapy Today, 40–42. Swain, T. A., & Mcgwin, G. (2016). Yoga-related injuries in the United States from 2001 to 2014. Orthopaedic Journal of Sports Medicine, 4(11), 2325967116671703. doi:10.1177/2325967116671703 PMID:27896293 The EQUATOR Network | Enhancing the Quality and Transparency of Health Research. (n.d.). Retrieved October 19, 2019, from https://www.equator-network.org/ Turner, L., Shamseer, L., Altman, D. G., Schulz, K. F., & Moher, D. (2012). Does use of the CONSORT Statement impact the completeness of reporting of randomised controlled trials published in medical journals? Systematic Reviews, 1(1), 60. doi:10.1186/2046-4053-1-60 PMID:23194585 von Elm, E., Altman, D. G., Egger, M., Pocock, S. J., Gøtzsche, P. C., & Vandenbroucke, J. P. (2007). Strengthening the reporting of observational studies in epidemiology (STROBE) statement: Guidelines for reporting observational studies. British Medical Journal, 335(7624), 806–808. doi:10.1136/ bmj.39335.541782.AD PMID:17947786

ADDITIONAL READING Green, B. N., & Johnson, C. D. (n.d.). Inter professional collaboration in research, education, and clinical practice: working together for a better future. doi:10.7899/JCE-14-36 How to Make Sense of Yoga-Related Research. (n.d.). Retrieved October 19, 2019, from https://yogainternational.com/article/view/how-to-make-sense-of-yoga-related-research International Association of Yoga Therapists. Yoga Therapy Competency Areas For use in IAYT Continuing Education. (n.d.). https://cdn.ymaws.com/www.iayt.org/resource/resmgr/docs_certification_all/ docs_certification/recertification/ce_competency_extract_06_201.pdf International Journal of Yoga Therapy. Current Issue. International Association of Yoga Therapists (IAYT). (n.d.). Retrieved October 19, 2019, from https://www.iayt.org/page/IJYTCurrent

156

 Standards in Yoga Research and Reporting

Maryland University of Integrative Health. Professional Certificate in Evidence-Based Research and Informed Practice - Continuing Education. (n.d.). Retrieved October 19, 2019, from https://ce.muih.edu/ browse/ce/programs/ebrcertificate Project to Enhance Research Literacy (PERL) from the Academic Collaborative for Integrative Health. (n.d.). Retrieved October 19, 2019, from https://integrativehealth.org/welcomeperl Scottish Intercollegiate Guidelines Network. Checklists and notes. (n.d.). Retrieved October 19, 2019, from https://www.sign.ac.uk/checklists-and-notes.html Society for Vascular Nursing. (2008). Practical Tips in Starting a Journal Club. Retrieved October 19, 2019, from https://svnnet.org/wp-content/uploads/2015/09/Journal-Club-Outline.pdf The Journal of Alternative and Complementary Medicine. Mary Ann Liebert, Inc., publishers. (n.d.). Retrieved October 19, 2019, from https://home.liebertpub.com/publications/the-journal-of-alternativeand-complementary-medicine/26 World Health Organization. Ethical standards and procedures for research with human beings. (n.d.). Retrieved October 19, 2019, from https://www.who.int/ethics/research/en/

KEY TERMS AND DEFINITIONS Attrition: A reduction in the number of participants in a study. Clinical Trials (Including Double-Blind, Randomized, Placebo-Controlled Trial, Randomized Controlled Trial): A type of study design that assigns participants to an intervention (treatment) or a control/comparison group in order to test the effects or efficacy of the intervention. Clinical trials may involve randomization (the random allocation of participants to groups) and/or blinding, which prevents participants, researchers, or other involved personnel from knowing which participants were assigned to which group. Comparative Effectiveness Research: Research that directly compares currently used healthcare interventions to one another to establish which is most useful for a given population. Comparison Groups (Equal Attention Control, Waitlist Control): In a study design that utilizes a comparison, these are the groups to which the intervention or exposure groups are compared. Generalizability: The ability to infer study results taken from a sample to the larger population. Intervention/Interventionist: The intervention component of a study is the treatment of interest. The interventionist therefore would be the individual tasked with administering the intervention to participants. In the case of a yoga study, the interventionist could be a yoga instructor. Methodological Biases: Errors that are systematically introduced during the design, execution, analysis, or reporting of a research study that encourages one set of results over another. Observational Research: A type of study wherein participants are surveilled or surveyed in an uncontrolled setting. No interventions are applied in this type of study, instead exposures are recorded to determine their association to an outcome. The most common observational study designs seen in biomedical research are cohort, case-control, and cross-sectional designs.

157

 Standards in Yoga Research and Reporting

Placebo Effects: Placebo controls indicate a non-treatment, which ideally has no effect on the participants, but still gives them the impression that they are engaging with a treatment. It follows then that a placebo effect is a positive effect that is still seen, even though the participant is not engaging with an active treatment. Sample Size: Sample size describes the number of participants within a study. Since we cannot feasibly test everyone in most populations, we rely on a subset or sample of the population who is willing to participate in the study. Researchers will calculate power to determine that they have a large enough sample size in order to achieve significance with the expected effect size.

158

159

Chapter 10

Yoga Therapy:

An Overview of Key Research and the Underlying Mechanisms Eileen M. Lafer https://orcid.org/0000-0002-0493-6522 University of Texas Health Science Center at San Antonio, USA & Beyond Wellness Aljezur, Portugal Margaret McCuiston Southern California University of Health Sciences, USA & Seashore Healing, USA Ann Swanson Maryland University of Integrative Health, USA & Kripalu School of Integrative Yoga Therapy, USA

ABSTRACT This chapter reviews the evidence supporting the efficacy of yoga therapy for wellness, and as an adjunct to standard care for a number of chronic conditions. The underlying mechanisms are explored, including the physiological and biochemical changes that have been observed in yoga practitioners. Yoga has been found to activate the relaxation response—a physiological state which reduces stress on bodily systems. Yoga leads to changes in gene expression, including decreases in the expression of genes involved in stress and inflammation. The positive effects of yoga therapy are interpreted through the lens of the biopsychosocial-spiritual model, which cultivates eudaimonic well-being and salutogenesis. Researchers attribute a wide range of yoga’s therapeutic benefits largely to its whole-person approach to well-being.

INTRODUCTION This chapter provides an overview of the research supporting the therapeutic benefits of yoga for health promotion, and as an adjunct to standard care for the treatment of a number of common physical and mental health conditions (MHCs). It also explores the underlying physiological mechanisms. A key mechanism behind yoga’s benefits is that yoga reduces extraneous load from stress on the body’s systems. The extra load is due to the fact that the body has to work harder to maintain homeostasis, taxing DOI: 10.4018/978-1-7998-3254-6.ch010

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga Therapy

the organs and systems and contributing to chronic disease. Allostasis, a term often used in integrative healthcare, refers to the physiological changes which a living organism undergoes in order to maintain homeostasis during stress (Sterling & Eyer, 1988). The stress response is mediated by the regulated release of hormones and other chemical messengers by the neuroendocrine system. During real or perceived threats an individual undergoes several adaptations prompted by the activation of the sympathetic nervous system (SNS) in which a cascade of neurophysiological and metabolic shifts occurs to maintain homeostasis. This response is adaptive to the organism following an acute stress, and maladaptive under conditions of chronic stress, referred to as allostatic overload, which may lead to various comorbid physical or mental health diseases. The whole-person approach of yoga therapy, which looks at individuals through a biopsychosocialspiritual lens known in the yoga tradition as the pañcamaya (five layers of being) model acts through neurophysiological mechanisms that lead to psychological and eudaimonic well-being. The accessibility and adaptability of the top-down and bottom-up approaches of yoga allow for self-regulation of the nervous system, thereby reducing allostatic load. Top-down strategies derive from or are influenced by frontal brain regions and cognitive function and bottom-up approaches are derived from the emotion generative regions of the brain and are influenced by peripheral sensory afferents (Gard et al., 2014). In the practices of yoga, top-down may include meditation, attention-controlled practices, ethics, and monitoring while bottom-up may include sustained asana (movement) and pranayama (breath) practices. The research highlighted in this chapter discusses the utility of yoga as a therapeutic intervention and model for health.

MUSCULOSKELETAL AND NEUROLOGICAL HEALTH It is well-accepted that physical exercise greatly benefits health, including improved range of motion, flexibility, strength, and balance - all of which improve outcomes in therapeutic populations. Although yoga asana offers a form of physical exercise, the whole practice of yoga differentiates from other commonly recommended exercises due to additional tools including breath regulation, moment-to-moment awareness, and the integration of philosophical teachings. This whole-person approach to well-being makes yoga particularly well-suited for complex musculoskeletal and neurological-based chronic diseases. Yoga asana seems to have positive effects on joint and skeletal health. A meta-analysis of 10 randomized controlled trials (RCTs) with 967 participants concluded that yoga may be effective for improving pain, function, and stiffness for those with osteoarthritis of the knee (Lauche et al., 2019). Yoga shows promise for other musculoskeletal and neurological concerns and conditions. A systematic review found that yoga improved balance in healthy individuals (Jeter et al., 2014), and smaller studies demonstrated improvements for those in stroke recovery (Schmid et al., 2012), older adults with a history of falling (Ni et al., 2014), and individuals with Parkinson’s disease (Van Puymbroeck et al., 2018). Other studies showed outcome improvements with yoga for specific conditions such as rheumatoid arthritis and osteoarthritis (Moonaz et al., 2015), osteoporosis (Lu et al., 2016), rotator cuff dysfunction (Fishman et al., 2011), sciatica and disc herniation (Monro et al., 2015), scoliosis (Fishman et al., 2014), hyperkyphosis (Greendale et al., 2009), fibromyalgia (Carson et al., 2010) and headaches (Bhatia et al., 2007; Kim, 2015; Vasudha et al., 2018). However, these studies are small and more research and systematic reviews are needed in all of these areas.

160

 Yoga Therapy

Chronic back pain, particularly nonspecific low back pain is a larger area of study within the field of yoga research. This may be due to both the need for a safe, low-cost intervention and the prevalence – 80 percent of people will experience lower back pain at some point in their life (Balague et al., 2012). Yoga provides several components that seem to help back pain, including physical movement. Exercise has been shown to be a generally effective treatment for general back pain as it can contribute to combating a sedentary lifestyle and obesity, while improving posture and functional status (Hayden et al., 2005). A growing body of research supports yoga for general back pain (Chuang et al., 2012; Cramer et al., 2013; Groessl et al., 2017). A systematic review and meta-analysis on yoga for lower back pain reported strong evidence that yoga reduced pain in the short term and moderate evidence that yoga reduced pain at longer-term (Cramer et al., 2013). Yoga’s whole-person approach contributes to the promising evidence supporting benefits for multifaceted conditions with a strong neurological basis including chronic pain (Govindaraj et al., 2016). Research indicates that many of yoga’s beneficial effects are through actions on the master control system - the nervous system. Research supports yoga for many pain-related conditions as well as general pain management (Ward et al., 2013; Zeidan et al., 2015). These effects seem to have several contributing mechanisms. Yoga increased levels of the neurotransmitter gamma-aminobutyric-acid (GABA) and low levels are associated with chronic pain (Streeter et al., 2007; Streeter et al., 2010). Additionally, brain imagery showed that yoga practices, including meditation, increased cortical gray matter thickness in key brain areas associated with improved tolerance to pain (Villemure et al., 2014). The evidence supporting yoga and chronic pain is promising and represents an area where therapeutic yoga can make a notable impact on healthcare (Moonaz, 2019). By addressing chronic pain, yoga has the potential to improve outcomes and quality of life for people with disabilities due to a variety of musculoskeletal and neurological diseases.

MENTAL HEALTH While genetic and environmental factors play a role in the pathology of MHCs, dysregulation of the stress response can lead to chronic disease and various comorbid conditions including post-traumatic stress disorder (PTSD), depression, generalized anxiety disorder, sleep disorders, and other psychiatric disorders (Miller & O’Callaghan, 2002; Muehsam et al., 2017). Research indicates mind-body therapies like yoga which integrate top-down and bottom-up approaches may benefit individuals suffering from MHCs; yoga affects various neurophysiological, neurocognitive, neuroendocrine, and cardiorespiratory mechanisms that are induced during the stress response (Muehsam et al., 2017; Streeter et al., 2012). The changes a yoga practice offers may include increased vagal tone, increased heart rate variability (HRV), increased brain-derived neurotrophic factor, and increased serum prolactin. Other adaptations include down regulation of hypothalamus-pituitary-adrenal (HPA) axis and serum cortisol, increased electroencephalogram alpha wave activity, increased GABA levels, increased serotonin, increased acetylcholine, increased glutamate, and regulated dopamine and norepinephrine (Balasubramaniam et al., 2013; Streeter et al., 2012). A meta-analysis of 23 RCTs with 977 participants found the physical aspects of yoga were an effective intervention for depression (Kvam et al., 2016). Systematic reviews found that yoga interventions increased remission rates, reduced depressive symptoms, and improved mood with and without adjunct antidepressants or pharmacotherapy in depressed individuals (Balasubramaniam et al., 2013; Macy et 161

 Yoga Therapy

al., 2018). The American Academy of Family Physicians suggested that yoga be prescribed as both a monotherapy and adjunct therapy for the treatment of depression (Saeed et al., 2019). Yoga may be beneficial for decreasing anxiety, especially when performed in group settings and over long-term periods (Balasubramaniam et al., 2013). One study suggested that the neurophysiological changes that occur post-yoga such as increased thalamic GABA levels are part of the mechanism whereby yoga decreases anxiety, noting the effects breathwork and meditation have on SNS and HPA axis (Cramer et al., 2018). Other studies have suggested that long-term reduction of anxiety is increased when yoga is adjunct to other supportive interventions (Macy et al., 2018; Saeed et al., 2019). Research also supports the somatic exploratory aspects of yoga in the treatment of PTSD (Farb et al., 2015; Sullivan et al., 2018). RCTs and systematic reviews indicated a decrease in PTSD symptomology dependent upon style of yoga and duration, and support its use as an ancillary approach, having noted the benefit a group format had with its patient-choice focus and encouraged participation (Gallegos et al., 2017; Uebelacker & Broughton, 2016). Reduced symptoms and a remission period for women with chronic treatment-resistant PTSD when compared to other non-conventional therapies suggested that the somatic and interoceptive aspects of yoga were crucial variables (van der Kolk et al., 2014). The limitations in this research included short duration and no formal follow-up. As sleep issues are often comorbid with MHCs systematic reviews have indicated yoga as an ancillary treatment for sleep problems with limited adverse effects including a significant reduction in necessity of sleep aids and high outcomes in daytime dysfunction (Balasubramaniam et al., 2013). Where yoga is designed to promote vagal tone and reduce allostatic load for MHCs and comorbidities, its components including self-reported outcomes and heightened body awareness may promote eudaimonic well-being and resilience. Recent research looked at polyvagal theory and the yoga gunas as a framework to understand and promote resilience and eudaimonic well-being (Sullivan et al., 2018). More research which informs and supports healthcare as it relates to MHCs and trauma-related illnesses is encouraged (Schmalzl et al., 2015).

ENDOCRINE AND DIGESTIVE HEALTH At the root of most endocrine disorders is allostatic overload (McEwen, 2005). It was hypothesized that the mechanism underlying the effectiveness of yoga for the treatment of a number of medical conditions is through a reduction in allostatic load (Streeter et al., 2012). There is emerging support for this hypothesis (Pascoe et al., 2017). In a recent meta-analysis of 42 RCTs in which 2944 subjects were given either a yoga intervention or standard exercise, improvements were seen in a number of markers indicating improved regulation of the HPA axis, as well as of the SNS in the yoga subjects relative to the controls. This included decreases in waking and evening cortisol, resting blood pressure (BP), resting heart rate (RHR), fasting blood glucose (FBG), cholesterol and low-density lipoprotein (LDL). In addition, the yoga subjects showed increases in HRV supporting models in which yoga promotes resilience through regulation of the ventral vagal pathway within the parasympathetic nervous system (Sullivan et al., 2018). A strength of this analysis is that it only included studies in which yoga was compared to a standard exercise group. Some statistical heterogeneity was observed which may be due to variations in the health of the subjects under study, variations in the yoga and exercise protocols utilized, and relatively small sample sizes. Despite these limitations, the aggregate analysis is extremely encouraging.

162

 Yoga Therapy

Support for yoga for the treatment of a number of specific endocrine disorders is growing. The most well studied is type 2 diabetes which has a major lifestyle component including poor diet and lack of physical activity. In a recent meta-analysis from 8 controlled trials in which 842 human subjects with type 2 diabetes were either given a yoga intervention or standard exercise (Jayawardena et al., 2018) significant reductions were found in FBG, post-prandial blood glucose (PPG), HbA1C, and body-mass index (BMI) in the yoga subjects relative to the active controls, although some statistical heterogeneity was observed. A particular strength of this study is that it is the first meta-analysis in which yoga was compared to standard exercise, as it is already well established that exercise alone is beneficial for glycemic control in type 2 diabetics (Snowling & Hopkins, 2006). These results are consistent with a meta-analysis of 2473 individuals from 23 controlled trials in which a yoga intervention was compared to some control group, although the nature of the control group varied and was most frequently standard care (Thind et al., 2017). Improvements were reported in FBG, PPG, HbA1C, BMI, BP, waist-hip ratio, and cortisol in the yoga subjects relative to the controls. There is preliminary support for the benefits of yoga for patients with metabolic syndrome and obesity, although a common theme in systematic reviews and meta-analyses in this area is that we need more large RCTs with robust design and reporting methodologies (Cramer et al., 2016). Metabolic syndrome is the name given to a cluster of risk factors which predispose an individual to type 2 diabetes and cardiovascular disease. Individuals with at least three of the following risk factors are diagnosed with metabolic syndrome: abdominal obesity, high triglycerides, low HDL cholesterol, high BP and high FBG. However, the heterogeneous nature of the syndrome complicates its study. A meta-analysis was performed on seven controlled trials in which 794 participants diagnosed with metabolic syndrome were assigned to either a yoga intervention or some type of control group (Cramer et al., 2016). While the nature of the control group varied it was most frequently standard care. No adverse effects were reported due to the yoga intervention, and improvements were noted in waist circumference and systolic BP in the yoga subjects relative to the controls. However, these effects were not considered robust with respect to selection bias so only preliminary conclusions were reached. Likewise, a meta-analysis of studies looking at the effects of yoga on weight related outcomes on a variety of subject types found a significant reduction in BMI for overweight and obese subjects that had been exposed to a yoga intervention versus standard care, but the results were also not robust with respect to selection bias so again, only preliminary conclusions were drawn (Lauche et al., 2016). Inflammatory bowel disease (IBD) is a chronic condition with unpredictable flare-ups that cause considerable stress to patients. There is increasing interest in assessing whether psychosocial interventions can improve quality of life for these patients. A systematic review of mindfulness-based approaches concluded that a small number of studies show fairly strong support for mindfulness-based interventions improving quality of life, and reducing anxiety and depression in IBD patients (Hood & Jedel, 2017). This is another area in which more research is needed.

CARDIORESPIRATORY HEALTH Yoga shows great promise for the prevention and treatment of cardiovascular disease. A meta-analysis of the effect of yoga based interventions on cardiovascular disease risk factors was conducted on 3168 subjects enrolled in 44 RCTs (Cramer et al., 2014). Yoga was found to be superior to usual care for a number of clinical risk factors including BP, heart rate (HR), abdominal obesity, blood lipid levels and 163

 Yoga Therapy

insulin resistance. Interestingly, it was noted that yoga was found to be comparable or superior to national guideline endorsed interventions in the US and UK, and multinational guideline endorsed interventions in the EU. These results are consistent with the findings of a systematic review of nine RCTs which found that yoga promoted similar cardiovascular benefits in older adults (Barrows & Fleury, 2016), as well as with another meta-analysis of 45 RCTs with active controls which found that mindfulness-based meditation practices positively affected physiological measures of stress including HR, HRV, FBG, and cholesterol (Pascoe et al., 2017). A recent meta-analysis looking at the effects of different exercise interventions including endurance training, strength training, combined endurance and strength training, school sports programs, yoga, tai chi and qigong on RHR in 121 controlled trials found that the only forms of exercise which significantly decreased RHR in both males and females was yoga and endurance training (Reimers et al., 2018). A review of evidence-based lifestyle modification strategies found that yoga was associated with increased exercise capacity, reduced angina episodes, and improvements in atherosclerosis in patients with coronary heart disease (Freeman et al., 2019). The Ornish Lifestyle Medicine Program- which includes stress management (meditation, gentle yoga asana, and breathing techniques), physician supervised exercise, a plant-based diet, and social support—has shown improvements in cardiovascular outcomes, including a clinically significant reversal of coronary heart disease (Ornish et al., 1990). Yoga and mindfulness practices also have positive effects on respiratory health. A systematic review of 68 studies found that pranayama practices had both physiological and clinical benefits (Saoji et al., 2019). Improvements were found in neurocognitive, psychophysiological, biochemical, and metabolic systems in healthy individuals. A number of improvements were also seen in clinical populations, including improvements in cardiovascular function in patients with hypertension and cardiac arrhythmias, and the enhancement of pulmonary function in patients with bronchial asthma and pulmonary tuberculosis. There is also emerging evidence that yoga benefits patients with asthma and COPD. A systematic review of 15 RCTs of 1048 participants found moderate quality evidence that yoga probably leads to small improvements in symptoms and quality of life in patients with asthma (Yang et al., 2016). A meta-analysis of studies looking at the effects of meditative movement practices including yoga on COPD found that yoga may enhance exercise capacity, quality of life and lung function in COPD patients although the conclusions are limited by the heterogeneity of the study populations and interventions, as well as by the quality of the available studies (Wu et al., 2018).

LYMPHATIC AND IMMUNE HEALTH Cancer patients experience stress-related changes in immunity, which may include decreases in cellmediated immunity as well as increases in inflammatory cytokines (Lutgendorf & Sood, 2011). Indeed, chronic inflammation is a major risk factor for cancer, and also associated with poor outcomes in cancer patients (Wu et al., 2014). A systematic review of 15 RCTs concluded that yoga may improve immune system function by the down regulation of pro-inflammatory markers (including IL-1beta, IL-6, and TNFalpha), as well as by the enhancement of cell-mediated and mucosal immunity, although small sample sizes and heterogeneity in study design precluded performing a meta-analysis (Falkenberg et al., 2018). Cancer patients have to deal with the direct consequences of unregulated cell growth in the organ or tissue affected, the toxic consequences of their treatments, as well as the psychosocial effects of their diagnosis. Cancer related fatigue (CRF) is a serious problem which interferes with the quality of life of 164

 Yoga Therapy

cancer patients (Abrahams et al., 2016). A meta-analysis of 2183 subjects diagnosed with breast cancer enrolled in 17 RCTs assigned to either a yoga intervention or a varied control group found that yoga had a large impact on the mitigation of physical fatigue, a medium impact on cognitive fatigue, and a small impact on mental fatigue in the breast cancer patients (Dong et al., 2019). Another meta-analysis of 245 studies found that yoga as well as a number of other exercise and non-pharmacological interventions had moderate to large effects on the mitigation of CRF in cancer patients with a variety of different cancer types while they were undergoing treatment, and that yoga had the largest effect of all the modalities evaluated post treatment (Hilfiker et al., 2018). A meta-analysis of 41 RCTs evaluated the efficacy of a number of non-pharmacological interventions including standard exercise to mitigate depressive symptoms in breast cancer patients relative to standard care. Yoga, mindfulness and psychotherapy emerged as the interventions that significantly reduced depressive symptoms. Systematic reviews have reported that yoga may improve depression, distress, anxiety, sleep quality, fatigue, musculoskeletal symptoms, as well as biomarkers of stress, inflammation, and immune function in cancer patients (Danhauer et al., 2019; Danhauer et al., 2017; Lin et al., 2018). A comprehensive review of 138 clinical trials in which 10,660 cancer patients from 20 countries were studied found strong support for the integration of yoga into standard care for cancer patients and noted that while studies on the role of yoga in cancer prevention are lacking, this is an area that is ripe for future investigation (Agarwal & Maroko-Afek, 2018).

REPRODUCTIVE AND UROLOGICAL HEALTH Preliminary research on yoga and reproductive health, fertility, sexual health and urological health is encouraging. Depression can affect both the physical and mental health of mother and baby during prenatal, intranatal, and postnatal states. Systematic reviews and RCTs pointed at yoga’s effect to reduce anxiety and depression throughout the stages of pregnancy with little to no contraindications. A review of RCTs involving 298 women found that a yoga intervention during pregnancy decreased the duration of labor, improved maternal comfort during childbirth, and decreased stress antepartum when compared to routine or standard prenatal care. This same review evaluated eight studies that collectively demonstrated improved quality of life, decreased stress, decreased anxiety, and decreased sleep disturbances likely due to yoga’s ability to act on the SNS (Babbar et al., 2012). In more than half of these cases the mothers were educated, relatively healthy, and married, so more studies surveying other socioeconomic populations are desired to compare results. A review of RCTs representing 375 participants showed depressive symptoms were significantly lower in prenatally depressed women in yoga groups when compared to control groups. Sub-group analysis revealed that the levels of depression in integrated yoga groups which included some form of pranayama, meditation, or deep relaxation were significantly reduced, while the levels of depression in exercise-based yoga alone were not significantly reduced (Gong et al., 2015). Limitations included high participant dropout rate and varying or undisclosed yoga method. A review of 13 studies also supported yoga’s ability to reduce anxiety and depression in mothers yet found mixed results with yoga’s effect on sleep, relationships, and fatigue likely due to small sample sizes and time of intervention (first trimester versus third semester) (Sheffield & Woods-Giscombe, 2016). Further limitations on this topic include poor study design, incongruence in method of yoga, and variable levels of mental symptomatology. Meditation and controlled breathing practices may help to support mental health during pregnancy, and the movement-based components of yoga also prove to be beneficial. A review of 10 studies in165

 Yoga Therapy

volving 766 participants found yoga intervention reduced common physical ailments during pregnancy including lumbopelvic pain and leg pain. One study found fewer women developed preeclampsia and pregnancy-induced hypertension than those in the control (walking) group and suggested yoga as an ideal treatment (Jiang et al., 2015). Secondary outcomes of yoga interventions from prenatal to postnatal included yoga’s ability to act on anger levels, relationships, birth weight, gestational age, maternal-fetal attachment, HRV, mindfulness and well-being, however with so little research on these topics more research is needed to make definitive conclusions. Yoga research on fertility and sexual health in women and men is limited, however, there are promising RCTs supporting yoga’s efficacy. A study of 143 women under the age of 40 found increased frequency of pregnancy using mind-body therapies including relaxation training and social support components during in vitro fertilization cycles (Domar et al., 2011). A small study of 40 women found statistically significant improvements after yoga in all areas of sexual health, including desire, arousal, lubrication, orgasm, satisfaction, and decreased pain (Dhikav et al., 2010). Another study found self-reported increases in desire, intercourse satisfaction, performance, confidence, partner synchronization, erection, ejaculatory control, and orgasm in men after a 12-week yoga camp (Dhikav et al., 2010). Although little research has been directly conducted on yoga for sexual health to date, the well-known improvements yoga offers for the cardiovascular, respiratory, and nervous system would likely positively affect sexual health (Brotto et al., 2009). Research as it relates to the impact of yoga on urological health and incontinence in men and women is extremely limited. A small study of 34 women in a community center in Korea found improved continence after pelvic muscle exercises and suggested that these exercises, along with a yoga program, were an effective treatment for incontinence (Kim, 2015). Limitations included sample size and length of intervention.

HEALTH THROUGH THE LIFE STAGES There is a growing body of research on the therapeutic benefits of yoga for children and adolescents. Systematic reviews indicated that yoga can improve physical and mental well-being for children, and illustrated how yoga in schools may help students improve resilience, mood, and self-regulation skills pertaining to emotions and stress (Hagen & Nayar, 2014). Another review found physical activity including yoga improved physical self-perceptions and enhanced self-esteem (Lubans et al., 2016). A systematic review of 11 articles totaling 251 participants surveying attention-deficit/hyperactivity disorder (ADHD) and youth found statistically significant positive effects of yoga and mindfulness-based interventions. Results included improvements on ADHD symptomatology, hyperactivity, and inattention in the children. Positive effects were also found on the parents with regard to the parent-child relationship, parental stress, and parental trait-mindfulness (Chimiklis et al., 2018). Limitations included possibly inflated effects due to small sample sizes, poor methodology, and general bias. Research on yoga and autism spectrum disorder remains inconclusive, however, preliminary studies showed positive effects on social, emotional, and behavioral markers (Semple, 2019). Yoga is being effectively used in geriatric populations as it relates to physical mobility, executive functioning, cognition, and memory; all of which are affected as an individual ages. A comprehensive review of 47 studies which included clinical trials, RCTs, systematic reviews, and meta-analyses on yoga and the elderly found a wide range of benefits in physical function, emotional and mental states, social 166

 Yoga Therapy

connections, and vitality for life, noting the preventive and management-based functions yoga practices offer (Mooventhan & Nivethitha, 2017). A RCT surveying 87 subjects aged 60-years or older found those in the yoga group versus the control group receiving no yoga intervention showed greater performance scores in areas such as immediate verbal memory, delayed recall of visual and verbal memory, attention, and processing speed (Hariprasad et al., 2013). Limitations included an over 25% drop-out rate and lack of randomization as all participants were living in residential care facilities. A smaller study of 42 women aged 60-years and above found significantly increased cortical thickness in a left prefrontal lobe cluster in the participants who had been practicing yoga at least eight years when compared to the non-yoga practitioners indicating a likely increase in cognitive preservation (Afonso et al., 2017). Research on yoga, dementia, and mild cognitive impairment is promising. A review of eight studies found practicing yoga may have beneficial effects on cognitive function as it relates to attention, processing speed, and verbal memory. Yoga may also support sleep, neuropsychiatric symptoms and mood, and brain regions and functions affected in dementia patients (Brenes et al., 2019). Limitations included inconsistency of practice, lack of regulation on frequency and duration, and weak control groups. High satisfaction reports by elderly yoga participants and the variations and modifications offered by yoga suggested it as a supplementary treatment for adults with dementia or mild cognitive impairment. While there is very limited research on yoga and Alzheimer’s disease (AD), a small RCT found significant increases in bilateral hippocampus volume following six months of yoga which supports growing research on yoga’s ability to support neuroplasticity (Hariprasad et al., 2013). Several other small studies discuss the therapeutic benefits of yoga on AD patients including increased physical functioning and increased self-esteem, however, much more research on this topic is needed. There is emerging research on the caregivers of AD patients as it relates to compassion fatigue and high levels of stress, anxiety, and depression. One RCT surveying 46 volunteers found yoga to be an effective intervention in reducing these symptoms in caregivers (Danucalov et al., 2013). Research on both AD patients and caregivers is steadily growing.

CELLULAR HEALTH Telomeres are short sequences of DNA that protect the ends of chromosomes. Telomeres shorten with each division of a somatic cell, and as such are correlated with cellular aging. Cells with shortened telomeres are at risk of apoptosis (cell death), and shortened telomeres have been associated with atherosclerosis, diabetes, chronic stress and depression (Zhang et al., 2014). However, telomeres can also lengthen in a reaction catalyzed by the enzyme telomerase (Epel et al., 2009). A systematic review was conducted of 12 RCTs in which the effects of yoga practices on telomere length and telomerase activity were studied (Rathore & Abraham, 2018). Three studies looked at radiation induced telomere shortening in breast cancer patients; two of the three studies found less telomere shortening in the yoga group versus the control. Five additional studies measured telomere length in varied populations; four of the five studies found telomeres were longer in the mindfulness practitioners relative to controls. Six studies measured telomerase; five of the six studies found it to be higher at all time points measured in the mindfulness practitioners relative to controls. While these studies are small, they constitute emerging evidence that yoga practices including asana, pranayama, and meditation maintain telomere length. The mechanisms are suggested to include increased delivery of oxygen to cells through yoga practices such as asana and

167

 Yoga Therapy

pranayama, as well as stress reduction through regulation of the HPA axis by a wide range of mindfulness practices (Epel et al., 2009; Rathore & Abraham, 2018). A biochemical indicator of cellular aging is the epigenetic clock, which is characterized by changes in DNA methylation patterns. The epigenetic clock is accelerated by stress (Zannas et al., 2015). Faster running clocks have been associated with chronic disease (Horvath et al., 2014; Perna et al., 2016). Analysis of individual methylomes allows an estimate of their Intrinsic Epigenetic Age Acceleration (IEAA). In a small scale study the IEAA profiles of long-term meditators versus meditation naive controls were calculated (Chaix et al., 2017). The analysis revealed that IEAA was reduced for long-term meditators >52 years of age relative to age matched naive controls, and that the epigenetic aging rate decreased with increasing years of meditation practice. This indicated that sustained mindfulness practices may slow the epigenetic clock in aging adults and reduce age-related chronic disease. While it is well established that yoga practices are important for overall health and wellness, we are just beginning to understand the mechanisms for these effects at the molecular level. Gene expression profiling studies indicated that meditative practices increased the expression of genes associated with energy metabolism, mitochondrial function, insulin secretion and telomere maintenance, and at the same time reduced the expression of genes associated with stress and inflammation, with the nuclear factor kappa B pathway (NF-kappaB) pathway genes as the top down-regulated hub (Bhasin et al., 2013). This is consistent with a recent review of 18 studies of a variety of mind-body interventions which consistently found downregulation of the NF-kappaB pathway which may lead to a reduced risk of inflammatory diseases (Buric et al., 2017).

BIOPSYCHOSOCIAL-SPIRITUAL MODEL AND SALUTOGENESIS The biopsychosocial-spiritual model is a holistic and humanist view of life-sciences that fits well with the study of yoga (Evans et al., 2009; Saad et al., 2017). This approach directly reflects the yoga-based pañcamaya model, which is an ancient, whole-person approach to health and well-being that acknowledges both the physical aspects, as well as the more subtle spiritual aspects of being. Biologically, yoga leads to improved overall physical fitness and health, affecting every system of the body. The psychological effects can be seen through improved mood, self-regulation, and emotional-regulation, and reduced allostatic load through both top-down and bottom-up processes. Group yoga classes offer social support that is often observed as a valuable component of efficacy in research (Buffart et al., 2012; Ross et al., 2013). In the spiritual domain, yoga therapy works directly on inner resourcing, connecting to one’s values, and helping a person develop their own meaning and purpose in life. The spiritual effects may be harder to quantify and researchers have tried (Newberg, 2014). A deeper understanding of the spiritual effects of yoga therapy is desired as the bridge between evidence-based research and spirituality is being built (Sullivan et al., 2018). Allopathic medicine tends to focus on pathogenesis, a disease-based model with the primary aim being symptom management. Although yoga is successful at symptom management including pain relief, yoga therapy works mostly on a level of salutogenesis, which is a health and well-being model. Lifestyle changes and mindset shifts from yoga can help people move past a disease focus to cultivate human flourishing (Jonas et al., 2014). Yoga therapy moves individuals and populations beyond diagnosis, disease, and even death to emphasize meaning and purpose in life (Sullivan et al., 2018). It is important to note that much of the research to date has been conducted on standardized yoga and/or meditation protocols 168

 Yoga Therapy

rather than on individualized yoga therapy. As the therapeutic relationship and specifically tailored practices are key to this modality, we might expect enhanced outcomes of yoga therapy compared to a more limited, less individual approach (McCall et al., 2016). Yoga therapy has great potential for the prevention and management of major lifestyle diseases, as well as helping individuals and populations find a greater sense of meaning and purpose.

CONCLUSION Yoga therapy is a burgeoning field with both ancient roots and a growing body of modern scientific evidence to support its efficacy. There has been an exponential growth of yoga research, largely due to therapeutic populations with the top areas of study including mental health, cardiovascular disease, respiratory disease, and musculoskeletal disorders (Jeter et al., 2015). This growth of research, particularly in these key areas, is due to the promising findings which show that yoga has benefits for health promotion, and as an ancillary treatment for many conditions that affect the modern world. Yoga therapy will play a vital role as healthcare shifts towards a more sustainable model including integrative and preventative care. As a consequence, it is extremely important that research on yoga therapy continues to increase. There is a need for larger, well-designed RCTs. Many of the reported studies are small, lack estimates of statistical power, and lack information needed to evaluate the risk of bias. While meta-analyses have been used to manage some of the limitations, heterogeneity concerning the specific yoga interventions (style, intensity, frequency, duration), as well as the specific outcomes measured, has limited their utility. However, despite the limitations of the present studies, there is now a very solid research base that supports the efficacy of yoga therapy for health promotion and disease management. Research supports yoga as a cost-effective and affordable adjunct treatment for various conditions highlighting its adaptability and accessibility to meet the needs of the individual (Chuang et al., 2012; Macy et al., 2018; Stahl et al., 2015). The self-regulatory top-down and bottom-up approaches of yoga and the whole-person lens of the pañcamaya model encourage and promote eudaimonic well-being and resilience (Gard et al., 2014). In summary, research suggests that yoga therapy offers the potential to improve health outcomes, lower healthcare costs, prevent inflammation and stress-related lifestyle diseases, maintain and foster well-being, and promote a life of more purpose and meaning.

REFERENCES Abrahams, H. J., Gielissen, M. F., Schmits, I. C., Verhagen, C. A., Rovers, M. M., & Knoop, H. (2016). Risk factors, prevalence, and course of severe fatigue after breast cancer treatment: A meta-analysis involving 12 327 breast cancer survivors. Annals of Oncology: Official Journal of the European Society for Medical Oncology, 27(6), 965–974. doi:10.1093/annonc/mdw099 PMID:26940687 Afonso, R. F., Balardin, J. B., Lazar, S., Sato, J. R., Igarashi, N., Santaella, D. F., Lacerda, S. S., Amaro, E. Jr, & Kozasa, E. H. (2017). Greater cortical thickness in elderly female yoga practitioners-a cross-sectional study. Frontiers in Aging Neuroscience, 9, 201–201. doi:10.3389/fnagi.2017.00201 PMID:28676757 Agarwal, R. P., & Maroko-Afek, A. (2018). Yoga into cancer care: A review of the evidence-based research. International Journal of Yoga, 11, 3–29. PMID:29343927

169

 Yoga Therapy

Babbar, S., Parks-Savage, A. C., & Chauhan, S. P. (2012). Yoga during pregnancy: A review. American Journal of Perinatology, 29(06), 459–464. doi:10.1055-0032-1304828 PMID:22399208 Balague, F., Mannion, A. F., Pellise, F., & Cedraschi, C. (2012). Non-specific low back pain. Lancet, 379(9814), 482–491. doi:10.1016/S0140-6736(11)60610-7 PMID:21982256 Balasubramaniam, M., Telles, S., & Doraiswamy, P. M. (2012). Yoga on our minds: A systematic review of yoga for neuropsychiatric disorders. Frontiers in Psychiatry, 3, 117–117. PMID:23355825 Barrows, J. L., & Fleury, J. (2016). Systematic review of yoga interventions to promote cardiovascular health in older adults. Western Journal of Nursing Research, 38(6), 753–781. doi:10.1177/0193945915618610 PMID:26689218 Bhasin, M. K., Dusek, J. A., Chang, B. H., Joseph, M. G., Denninger, J. W., Fricchione, G. L., Benson, H., & Libermann, T. A. (2013). Relaxation response induces temporal transcriptome changes in energy metabolism, insulin secretion and inflammatory pathways. PLoS One, 8(5), e62817–e62817. doi:10.1371/ journal.pone.0062817 PMID:23650531 Bhatia, R., Dureja, G. P., Tripathi, M., Bhattacharjee, M., Bijlani, R. L., & Mathur, R. (2007). Role of temporalis muscle over activity in chronic tension type headache: Effect of yoga based management. Indian Journal of Physiology and Pharmacology, 51, 333–344. PMID:18476387 Bravo, C., Skjaerven, L. H., Guitard Sein-Echaluce, L., & Catalan-Matamoros, D. (2019). Effectiveness of movement and body awareness therapies in patients with fibromyalgia: A systematic review and metaanalysis. European Journal of Physical and Rehabilitation Medicine, 55(5), 646–657. doi:10.23736/ S1973-9087.19.05291-2 PMID:31106558 Brenes, G. A., Sohl, S., Wells, R. E., Befus, D., Campos, C. L., & Danhauer, S. C. (2019). The effects of yoga on patients with mild cognitive impairment and dementia: A scoping review. The American Journal of Geriatric Psychiatry, 27(2), 188–197. doi:10.1016/j.jagp.2018.10.013 PMID:30413292 Brotto, L. A., Mehak, L., & Kit, C. (2009). Yoga and sexual functioning: A review. Journal of Sex & Marital Therapy, 35(5), 378–390. doi:10.1080/00926230903065955 PMID:20183005 Buffart, L. M., van Uffelen, J. G., Riphagen, I. I., Brug, J., van Mechelen, W., Brown, W. J., & Chinapaw, M. J. (2012). Physical and psychosocial benefits of yoga in cancer patients and survivors, a systematic review and meta-analysis of randomized controlled trials. BMC Cancer, 12(1), 559. doi:10.1186/14712407-12-559 PMID:23181734 Buric, I., Farias, M., Jong, J., Mee, C., & Brazil, I. A. (2017). What is the molecular signature of mindbody interventions? A systematic review of gene expression changes induced by meditation and related practices. Frontiers in Immunology, 8, 670. doi:10.3389/fimmu.2017.00670 PMID:28670311 Carson, J. W., Carson, K. M., Jones, K. D., Bennett, R. M., Wright, C. L., & Mist, S. D. (2010). A pilot randomized controlled trial of the yoga of awareness program in the management of fibromyalgia. Pain, 151(2), 530–539. doi:10.1016/j.pain.2010.08.020 PMID:20946990

170

 Yoga Therapy

Chimiklis, A. L., Dahl, V., Spears, A. P., Goss, K., Fogarty, K., & Chacko, A. (2018). Yoga, mindfulness, and meditation interventions for youth with ADHD: Systematic review and meta-analysis. Journal of Child and Family Studies, 27(10), 3155–3168. doi:10.100710826-018-1148-7 Chuang, L. H., Soares, M. O., Tilbrook, H., Cox, H., Hewitt, C. E., Aplin, J., Semlyen, A., Trewhela, A., Watt, I., & Torgerson, D. J. (2012). A pragmatic multicentered randomized controlled trial of yoga for chronic low back pain: Economic evaluation. Spine, 37(18), 1593–1601. doi:10.1097/BRS.0b013e3182545937 PMID:22433499 Cramer, H., Langhorst, J., Dobos, G., & Lauche, R. (2016). Yoga for metabolic syndrome: A systematic review and meta-analysis. European Journal of Preventive Cardiology, 23(18), 1982–1993. doi:10.1177/2047487316665729 PMID:27550905 Cramer, H., Lauche, R., Anheyer, D., Pilkington, K., de Manincor, M., Dobos, G., & Ward, L. (2018). Yoga for anxiety: A systematic review and meta-analysis of randomized controlled trials. Depression and Anxiety, 35(9), 830–843. doi:10.1002/da.22762 PMID:29697885 Cramer, H., Lauche, R., Haller, H., & Dobos, G. (2013). A systematic review and meta-analysis of yoga for low back pain. The Clinical Journal of Pain, 29(5), 450–460. doi:10.1097/AJP.0b013e31825e1492 PMID:23246998 Cramer, H., Lauche, R., Haller, H., Steckhan, N., Michalsen, A., & Dobos, G. (2014). Effects of yoga on cardiovascular disease risk factors: A systematic review and meta-analysis. International Journal of Cardiology, 173(2), 170–183. doi:10.1016/j.ijcard.2014.02.017 PMID:24636547 Danhauer, S. C., Addington, E. L., Cohen, L., Sohl, S. J., Van Puymbroeck, M., Albinati, N. K., & CulosReed, S. N. (2019). Yoga for symptom management in oncology: A review of the evidence base and future directions for research. Cancer, 125(12), 1979–1989. doi:10.1002/cncr.31979 PMID:30933317 Danhauer, S. C., Addington, E. L., Sohl, S. J., Chaoul, A., & Cohen, L. (2017). Review of yoga therapy during cancer treatment. Supportive Care in Cancer, 25(4), 1357–1372. doi:10.100700520-016-3556-9 PMID:28064385 Danucalov, M. A., Kozasa, E. H., Ribas, K. T., Galduroz, J. C., Garcia, M. C., & Verreschi, I. T. (2013). … Leite, J. R. (2013). A yoga and compassion meditation program reduces stress in familial caregivers of Alzheimer’s disease patients. Evidence-Based Complementary and Alternative Medicine, •••, 513149–513149. PMID:23690846 Dhikav, V., Karmarkar, G., Gupta, R., Verma, M., Gupta, R., Gupta, S., & Anand, K. S. (2010). Yoga in female sexual functions. Journal of Sexual Medicine, 7(2), 964–970. doi:10.1111/j.1743-6109.2009.01580.x PMID:19912493 Domar, A. D., Rooney, K. L., Wiegand, B., Orav, E. J., Alper, M. M., Berger, B. M., & Nikolovski, J. (2011). Impact of a group mind/body intervention on pregnancy rates in IVF patients. Fertility and Sterility, 95(7), 2269–2273. doi:10.1016/j.fertnstert.2011.03.046 PMID:21496800 Dong, B., Xie, C., Jing, X., Lin, L., & Tian, L. (2019). Yoga has a solid effect on cancer-related fatigue in patients with breast cancer: A meta-analysis. Breast Cancer Research and Treatment, 177(1), 5–16. doi:10.100710549-019-05278-w PMID:31127466

171

 Yoga Therapy

Epel, E., Daubenmier, J., Moskowitz, J. T., Folkman, S., & Blackburn, E. (2009). Can meditation slow rate of cellular aging? Cognitive stress, mindfulness, and telomeres. Annals of the New York Academy of Sciences, 1172(1), 34–53. doi:10.1111/j.1749-6632.2009.04414.x PMID:19735238 Evans, S., Tsao, J. C., Sternlieb, B., & Zeltzer, L. K. (2009). Using the biopsychosocial model to understand the health benefits of yoga. Journal of Complementary & Integrative Medicine, 6(1), 1–22. doi:10.2202/1553-3840.1183 Falkenberg, R. I., Eising, C., & Peters, M. L. (2018). Yoga and immune system functioning: A systematic review of randomized controlled trials. Journal of Behavioral Medicine, 41(4), 467–482. doi:10.100710865-018-9914-y PMID:29429046 Farb, N., Daubenmier, J., Price, C. J., Gard, T., Kerr, C., Dunn, B. D., Klein, A. C., Paulus, M. P., & Mehling, W. E. (2015). Interoception, contemplative practice, and health. Frontiers in Psychology, 6, 763. doi:10.3389/fpsyg.2015.00763 PMID:26106345 Fishman, L. M., Groessl, E. J., & Sherman, K. J. (2014). Serial case reporting yoga for idiopathic and degenerative scoliosis. Global Advances in Health and Medicine: Improving Healthcare Outcomes Worldwide, 3(5), 16–21. doi:10.7453/gahmj.2013.064 PMID:25568820 Fishman, L. M., Wilkins, A. N., O’vadia, T., Konnoth, C., Rosner, B., & Schidhofer, S. (2011). Yoga-based maneuver effectively treats rotator cuff syndrome. Topics in Geriatric Rehabilitation, 27(2), 151–161. doi:10.1097/TGR.0b013e31821bfe68 Freeman, A. M., Taub, P. R., Lo, H. C., & Ornish, D. (2019). Intensive cardiac rehabilitation: An underutilized resource. Current Cardiology Reports, 21(4), 19. doi:10.100711886-019-1104-1 PMID:30828747 Gallegos, A. M., Crean, H. F., Pigeon, W. R., & Heffner, K. L. (2017). Meditation and yoga for posttraumatic stress disorder: A meta-analytic review of randomized controlled trials. Clinical Psychology Review, 58, 115–124. doi:10.1016/j.cpr.2017.10.004 PMID:29100863 Gard, T., Noggle, J. J., Park, C. L., Vago, D. R., & Wilson, A. (2014). Potential self-regulatory mechanisms of yoga for psychological health. Frontiers in Human Neuroscience, 8, 770. doi:10.3389/ fnhum.2014.00770 PMID:25368562 Gong, H., Ni, C., Shen, X., Wu, T., & Jiang, C. (2015). Yoga for prenatal depression: A systematic review and meta-analysis. BMC Psychiatry, 15(1), 14. doi:10.118612888-015-0393-1 PMID:25652267 Govindaraj, R., Karmani, S., Varambally, S., & Gangadhar, B. N. (2016). Yoga and physical exercise - a review and comparison. International Review of Psychiatry (Abingdon, England), 28(3), 242–253. doi :10.3109/09540261.2016.1160878 PMID:27044898 Greendale, G. A., Huang, M. H., Karlamangla, A. S., Seeger, L., & Crawford, S. (2009). Yoga decreases kyphosis in senior women and men with adult-onset hyperkyphosis: Results of a randomized controlled trial. Journal of the American Geriatrics Society, 57(9), 1569–1579. doi:10.1111/j.1532-5415.2009.02391.x PMID:19682114

172

 Yoga Therapy

Groessl, E. J., Liu, L., Chang, D. G., Wetherell, J. L., Bormann, J. E., Atkinson, J. H., Baxi, S., & Schmalzl, L. (2017). Yoga for military veterans with chronic low back pain: A randomized clinical trial. American Journal of Preventive Medicine, 53(5), 599–608. doi:10.1016/j.amepre.2017.05.019 PMID:28735778 Hagen, I., & Nayar, U. S. (2014). Yoga for children and young people’s mental health and well-being: Research review and reflections on the mental health potentials of yoga. Frontiers in Psychiatry, 5, 35. doi:10.3389/fpsyt.2014.00035 PMID:24765080 Hariprasad, V. R., Koparde, V., Sivakumar, P. T., Varambally, S., Thirthalli, J., Varghese, M., ... Gangadhar, B. N. (2013). Randomized clinical trial of yoga-based intervention in residents from elderly homes: Effects on cognitive function. Indian Journal of Psychiatry, 55(7), 357–363. doi:10.4103/00195545.116308 PMID:24049199 Hariprasad, V. R., Varambally, S., Shivakumar, V., Kalmady, S. V., Venkatasubramanian, G., & Gangadhar, B. N. (2013). Yoga increases the volume of the hippocampus in elderly subjects. Indian Journal of Psychiatry, 55(7), 394–396. doi:10.4103/0019-5545.116309 PMID:24049206 Hayden, J. A., van Tulder, M. W., Malmivaara, A., & Koes, B. W. (2005). Exercise therapy for treatment of non-specific low back pain. Cochrane Database of Systematic Reviews, 3, CD000335. doi:10.1002/14651858.CD000335.pub2 PMID:16034851 Hilfiker, R., Meichtry, A., Eicher, M., Nilsson Balfe, L., Knols, R. H., Verra, M. L., & Taeymans, J. (2018). Exercise and other non-pharmaceutical interventions for cancer-related fatigue in patients during or after cancer treatment: A systematic review incorporating an indirect-comparisons meta-analysis. British Journal of Sports Medicine, 52(10), 651–658. doi:10.1136/bjsports-2016-096422 PMID:28501804 Hood, M. M., & Jedel, S. (2017). Mindfulness-based interventions in inflammatory bowel disease. Gastroenterology Clinics, 46(4), 859–874. doi:10.1016/j.gtc.2017.08.008 PMID:29173527 Horvath, S., Erhart, W., Brosch, M., Ammerpohl, O., von Schonfels, W., Ahrens, M., … Hampe, J. (2014). Obesity accelerates epigenetic aging of human liver. Proceedings of the National Academy of Sciences of the United States of America, 111, 15538-15543. 10.1073/pnas.1412759111 Jayawardena, R., Ranasinghe, P., Chathuranga, T., Atapattu, P. M., & Misra, A. (2018). The benefits of yoga practice compared to physical exercise in the management of type 2 diabetes mellitus: A systematic review and meta-analysis. Diabetes & Metabolic Syndrome, 12(5), 795–805. doi:10.1016/j. dsx.2018.04.008 PMID:29685823 Jeter, P. E., Nkodo, A. F., Moonaz, S. H., & Dagnelie, G. (2014). A systematic review of yoga for balance in a healthy population. Journal of Alternative and Complementary Medicine (New York, N.Y.), 20(4), 221–232. doi:10.1089/acm.2013.0378 PMID:24517304 Jeter, P. E., Slutsky, J., Singh, N., & Khalsa, S. B. (2015). Yoga as a therapeutic intervention: A bibliometric analysis of published research studies from 1967 to 2013. Journal of Alternative and Complementary Medicine (New York, N.Y.), 21(10), 586–592. doi:10.1089/acm.2015.0057 PMID:26196166 Jiang, Q., Wu, Z., Zhou, L., Dunlop, J., & Chen, P. (2015). Effects of yoga intervention during pregnancy: A review for current status. American Journal of Perinatology, 32(06), 503–514. doi:10.1055-00341396701 PMID:25535930

173

 Yoga Therapy

Jonas, W. B., Chez, R. A., Smith, K., & Sakallaris, B. (2014). Salutogenesis: The defining concept for a new healthcare system. Global Advances in Health and Medicine: Improving Healthcare Outcomes Worldwide, 3(3), 82–91. doi:10.7453/gahmj.2014.005 PMID:24944875 Kim, G. S., Kim, E. G., Shin, K. Y., Choo, H. J., & Kim, M. J. (2015). Combined pelvic muscle exercise and yoga program for urinary incontinence in middle-aged women. Japan Journal of Nursing Science, 12(4), 330–339. doi:10.1111/jjns.12072 PMID:25705816 Kim, S. D. (2015). Effects of yoga exercises for headaches: A systematic review of randomized controlled trials. Journal of Physical Therapy Science, 27(7), 2377–2380. doi:10.1589/jpts.27.2377 PMID:26311986 Kvam, S., Kleppe, C. L., Nordhus, I. H., & Hovland, A. (2016). Exercise as a treatment for depression: A meta-analysis. Journal of Affective Disorders, 202, 67–86. doi:10.1016/j.jad.2016.03.063 PMID:27253219 Lauche, R., Hunter, D. J., Adams, J., & Cramer, H. (2019). Yoga for osteoarthritis: A systematic review and meta-analysis. Current Rheumatology Reports, 21(9), 47–47. doi:10.100711926-019-0846-5 PMID:31338685 Lauche, R., Langhorst, J., Lee, M. S., Dobos, G., & Cramer, H. (2016). A systematic review and meta-analysis on the effects of yoga on weight-related outcomes. Preventive Medicine, 87, 213–232. doi:10.1016/j.ypmed.2016.03.013 PMID:27058944 Lin, P. J., Peppone, L. J., Janelsins, M. C., Mohile, S. G., Kamen, C. S., Kleckner, I. R., Fung, C., Asare, M., Cole, C. L., Culakova, E., & Mustian, K. M. (2018). Yoga for the management of cancer treatmentrelated toxicities. Current Oncology Reports, 20(1), 5. doi:10.100711912-018-0657-2 PMID:29388071 Lu, Y. H., Rosner, B., Chang, G., & Fishman, L. M. (2016). Twelve-minute daily yoga regimen reverses osteoporotic bone loss. Topics in Geriatric Rehabilitation, 32(2), 81–87. doi:10.1097/TGR.0000000000000085 PMID:27226695 Lubans, D., Richards, J., Hillman, C., Faulkner, G., Beauchamp, M., Nilsson, M., Kelly, P., Smith, J., Raine, L., & Biddle, S. (2016). Physical activity for cognitive and mental health in youth: A systematic review of mechanisms. Pediatrics, 138(3), e20161642. doi:10.1542/peds.2016-1642 PMID:27542849 Lutgendorf, S. K., & Sood, A. K. (2011). Biobehavioral factors and cancer progression: Physiological pathways and mechanisms. Psychosomatic Medicine, 73(9), 724–730. doi:10.1097/PSY.0b013e318235be76 PMID:22021459 Macy, R. J., Jones, E., Graham, L. M., & Roach, L. (2018). Yoga for trauma and related mental health problems: A meta-review with clinical and service recommendations. Trauma, Violence & Abuse, 19(1), 35–57. doi:10.1177/1524838015620834 PMID:26656487 McCall, T., Satish, L., & Tiwari, S. (2016). History, philosophy, and practice of yoga therapy. In S. B. S. Khalsa, L. Cohen, T. McCall, & S. Telles (Eds.), The principles and practice of yoga in health care (pp. 31–46). Handspring Publishing Limited. McEwen, B. S. (2005). Stressed or stressed out: What is the difference? Journal of Psychiatry & Neuroscience, 30, 315–318. PMID:16151535

174

 Yoga Therapy

Miller, D. B., & O’Callaghan, J. P. (2002). Neuroendocrine aspects of the response to stress. Metabolism: Clinical and Experimental, 51(6), 5–10. doi:10.1053/meta.2002.33184 PMID:12040534 Monro, R., Bhardwaj, A. K., Gupta, R. K., Telles, S., Allen, B., & Little, P. (2015). Disc extrusions and bulges in nonspecific low back pain and sciatica: Exploratory randomised controlled trial comparing yoga therapy and normal medical treatment. Journal of Back and Musculoskeletal Rehabilitation, 28(2), 383–392. doi:10.3233/BMR-140531 PMID:25271201 Moonaz, S. (2019). Current research in yoga and pain. In N. Pearson, S. Prosko, & M. Sullivan (Eds.), Yoga and science in pain care (pp. 37–51). Singing Dragon. Moonaz, S. H., Bingham, C. O. III, Wissow, L., & Bartlett, S. J. (2015). Yoga in sedentary adults with arthritis: Effects of a randomized controlled pragmatic trial. The Journal of Rheumatology, 42(7), 1194–1202. doi:10.3899/jrheum.141129 PMID:25834206 Mooventhan, A., & Nivethitha, L. (2017). Evidence based effects of yoga practice on various health related problems of elderly people: A review. Journal of Bodywork and Movement Therapies, 21(4), 1028–1032. doi:10.1016/j.jbmt.2017.01.004 PMID:29037619 Muehsam, D., Lutgendorf, S., Mills, P. J., Rickhi, B., Chevalier, G., Bat, N., Chopra, D., & Gurfein, B. (2017). The embodied mind: A review on functional genomic and neurological correlates of mind-body therapies. Neuroscience and Biobehavioral Reviews, 73, 165–181. doi:10.1016/j.neubiorev.2016.12.027 PMID:28017838 Newberg, A. B. (2014). The neuroscientific study of spiritual practices. Frontiers in Psychology, 5, 215. doi:10.3389/fpsyg.2014.00215 PMID:24672504 Ni, M., Mooney, K., Richards, L., Balachandran, A., Sun, M., Harriell, K., Potiaumpai, M., & Signorile, J. F. (2014). Comparative impacts of tai chi, balance training, and a specially-designed yoga program on balance in older fallers. Archives of Physical Medicine and Rehabilitation, 95(9), 1620–1628. doi:10.1016/j. apmr.2014.04.022 PMID:24835753 Ornish, D., Brown, S. E., Scherwitz, L. W., Billings, J. H., Armstrong, W. T., Ports, T. A., ... Gould, K. L. (1990). Can lifestyle changes reverse coronary heart disease? The lifestyle heart trial. Lancet, 336(8708), 129–133. doi:10.1016/0140-6736(90)91656-U PMID:1973470 Pascoe, M. C., Thompson, D. R., & Ski, C. F. (2017). Yoga, mindfulness-based stress reduction and stress-related physiological measures: A meta-analysis. Psychoneuroendocrinology, 86, 152–168. doi:10.1016/j.psyneuen.2017.08.008 PMID:28963884 Perna, L., Zhang, Y., Mons, U., Holleczek, B., Saum, K. U., & Brenner, H. (2016). Epigenetic age acceleration predicts cancer, cardiovascular, and all-cause mortality in a German case cohort. Clinical Epigenetics, 8(1), 64–64. doi:10.118613148-016-0228-z PMID:27274774 Rathore, M., & Abraham, J. (2018). Implication of asana, pranayama and meditation on telomere stability. International Journal of Yoga, 11(3), 186–193. doi:10.4103/ijoy.IJOY_51_17 PMID:30233111

175

 Yoga Therapy

Reimers, A. K., Knapp, G., & Reimers, C. D. (2018). Effects of exercise on the resting heart rate: A systematic review and meta-analysis of interventional studies. Journal of Clinical Medicine, 7(12), 503. doi:10.3390/jcm7120503 PMID:30513777 Ross, A., Friedmann, E., Bevans, M., & Thomas, S. (2013). National survey of yoga practitioners: Mental and physical health benefits. Complementary Therapies in Medicine, 21(4), 313–323. doi:10.1016/j. ctim.2013.04.001 PMID:23876562 Saad, M., de Medeiros, R., & Mosini, A. C. (2017). Are we ready for a true biopsychosocial-spiritual model? The many meanings of “spiritual”. Medicines (Basel, Switzerland), 4(4), 79. doi:10.3390/medicines4040079 PMID:29088101 Saeed, S. A., Cunningham, K., & Bloch, R. M. (2019). Depression and anxiety disorders: Benefits of exercise, yoga, and meditation. American Family Physician, 99, 620–627. PMID:31083878 Saoji, A. A., Raghavendra, B. R., & Manjunath, N. K. (2019). Effects of yogic breath regulation: A narrative review of scientific evidence. Journal of Ayurveda and Integrative Medicine, 10(1), 50–58. doi:10.1016/j.jaim.2017.07.008 PMID:29395894 Schmalzl, L., Powers, C., & Henje Blom, E. (2015). Neurophysiological and neurocognitive mechanisms underlying the effects of yoga-based practices: Towards a comprehensive theoretical framework. Frontiers in Human Neuroscience, 9, 235. doi:10.3389/fnhum.2015.00235 PMID:26005409 Schmid, A. A., Van Puymbroeck, M., Altenburger, P. A., Schalk, N. L., Dierks, T. A., Miller, K. K., Damush, T. M., Bravata, D. M., & Williams, L. S. (2012). Poststroke balance improves with yoga: A pilot study. Stroke, 43(9), 2402–2407. doi:10.1161/STROKEAHA.112.658211 PMID:22836351 Semple, R. J. (2019). Review: Yoga and mindfulness for youth with autism spectrum disorder: Review of the current evidence. Child and Adolescent Mental Health, 24(1), 12–18. doi:10.1111/camh.12295 PMID:32677240 Sheffield, K. M., & Woods-Giscombe, C. L. (2016). Efficacy, feasibility, and acceptability of perinatal yoga on women’s mental health and well-being: A systematic literature review. Journal of Holistic Nursing, 34(1), 64–79. doi:10.1177/0898010115577976 PMID:25896571 Snowling, N. J., & Hopkins, W. G. (2006). Effects of different modes of exercise training on glucose control and risk factors for complications in type 2 diabetic patients: A meta-analysis. Diabetes Care, 29(11), 2518–2527. doi:10.2337/dc06-1317 PMID:17065697 Stahl, J. E., Dossett, M. L., LaJoie, A. S., Denninger, J. W., Mehta, D. H., Goldman, R., Fricchione, G. L., & Benson, H. (2015). Relaxation response and resiliency training and its effect on healthcare resource utilization. PLoS One, 10(10), e0140212. doi:10.1371/journal.pone.0140212 PMID:26461184 Sterling, P., & Eyer, J. (1988). Allostasis: A new paradigm to explain arousal pathology. John Wiley and Sons.

176

 Yoga Therapy

Streeter, C. C., Gerbarg, P. L., Saper, R. B., Ciraulo, D. A., & Brown, R. P. (2012). Effects of yoga on the autonomic nervous system, gamma-aminobutyric-acid, and allostasis in epilepsy, depression, and post-traumatic stress disorder. Medical Hypotheses, 78(5), 571–579. doi:10.1016/j.mehy.2012.01.021 PMID:22365651 Streeter, C. C., Jensen, J. E., Perlmutter, R. M., Cabral, H. J., Tian, H., Terhune, D. B., Ciraulo, D. A., & Renshaw, P. F. (2007). Yoga asana sessions increase brain gaba levels: A pilot study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 13(4), 419–426. doi:10.1089/acm.2007.6338 PMID:17532734 Streeter, C. C., Whitfield, T. H., Owen, L., Rein, T., Karri, S. K., Yakhkind, A., Perlmutter, R., Prescot, A., Renshaw, P. F., Ciraulo, D. A., & Jensen, J. E. (2010). Effects of yoga versus walking on mood, anxiety, and brain gaba levels: A randomized controlled MRS study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 16(11), 1145–1152. doi:10.1089/acm.2010.0007 PMID:20722471 Sullivan, M. B., Erb, M., Schmalzl, L., Moonaz, S., Noggle Taylor, J., & Porges, S. W. (2018). Yoga therapy and polyvagal theory: The convergence of traditional wisdom and contemporary neuroscience for self-regulation and resilience. Frontiers in Human Neuroscience, 12, 67. doi:10.3389/fnhum.2018.00067 PMID:29535617 Thind, H., Lantini, R., Balletto, B. L., Donahue, M. L., Salmoirago-Blotcher, E., Bock, B. C., & ScottSheldon, L. A. J. (2017). The effects of yoga among adults with type 2 diabetes: A systematic review and meta-analysis. Preventive Medicine, 105, 116–126. doi:10.1016/j.ypmed.2017.08.017 PMID:28882745 Uebelacker, L. A., & Broughton, M. K. (2016). Yoga for depression and anxiety: A review of published research and implications for healthcare providers. Rhode Island Medical Journal, 99, 20–22. PMID:26929966 van der Kolk, B. A., Stone, L., West, J., Rhodes, A., Med, M. S. W., Emerson, D., ... Spinazzola, J. (2014). Yoga as an adjunctive treatment for posttraumatic stress disorder: A randomized controlled trial. The Journal of Clinical Psychiatry, 75(06), 559–565. doi:10.4088/JCP.13m08561 PMID:25004196 Van Puymbroeck, M., Walter, A. A., Hawkins, B. L., Sharp, J. L., Woschkolup, K., & Urrea-Mendoza, E. (2018). … Schmid, A. A. (2018). Functional improvements in parkinson’s disease following a randomized trial of yoga. Evidence-Based Complementary and Alternative Medicine, •••, 8516351. PMID:29967649 Vasudha, M. S., Manjunath, N. K., & Nagendra, H. R. (2018). Changes in midas, perceived stress, frontalis muscle activity and non-steroidal anti-inflammatory drugs usage in patients with migraine headache without aura following ayurveda and yoga compared to controls: An open labeled non-randomized study. Annals of Neurosciences, 25(4), 250–260. doi:10.1159/000492269 PMID:31000965 Villemure, C., Ceko, M., Cotton, V. A., & Bushnell, M. C. (2014). Insular cortex mediates increased pain tolerance in yoga practitioners. Cerebral Cortex (New York, N.Y.), 24(10), 2732–2740. doi:10.1093/ cercor/bht124 PMID:23696275 Ward, L., Stebbings, S., Cherkin, D., & Baxter, G. D. (2013). Yoga for functional ability, pain and psychosocial outcomes in musculoskeletal conditions: A systematic review and meta-analysis. Musculoskeletal Care, 11(4), 203–217. doi:10.1002/msc.1042 PMID:23300142

177

 Yoga Therapy

Wu, L. L., Lin, Z. K., Weng, H. D., Qi, Q. F., Lu, J., & Liu, K. X. (2018). Effectiveness of meditative movement on COPD: A systematic review and meta-analysis. International Journal of Chronic Obstructive Pulmonary Disease, 13, 1239–1250. doi:10.2147/COPD.S159042 PMID:29713157 Wu, Y., Antony, S., Meitzler, J. L., & Doroshow, J. H. (2014). Molecular mechanisms underlying chronic inflammation-associated cancers. Cancer Letters, 345(2), 164–173. doi:10.1016/j.canlet.2013.08.014 PMID:23988267 Yang, Z. Y., Zhong, H. B., Mao, C., Yuan, J. Q., Huang, Y. F., Wu, X. Y., ... Tang, J. L. (2016). Yoga for asthma. Cochrane Database of Systematic Reviews, 4, 368. PMID:27115477 Zannas, A. S., Arloth, J., Carrillo-Roa, T., Iurato, S., Roh, S., Ressler, K. J., Nemeroff, C. B., Smith, A. K., Bradley, B., Heim, C., Menke, A., Lange, J. F., Brückl, T., Ising, M., Wray, N. R., Erhardt, A., Binder, E. B., & Mehta, D. (2015). Lifetime stress accelerates epigenetic aging in an urban, African American cohort: Relevance of glucocorticoid signaling. Genome Biology, 16(1), 1–12. doi:10.118613059-0150828-5 PMID:26673150 Zeidan, F., Emerson, N. M., Farris, S. R., Ray, J. N., Jung, Y., McHaffie, J. G., & Coghill, R. C. (2015). Mindfulness meditation-based pain relief employs different neural mechanisms than placebo and sham mindfulness meditation-induced analgesia. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 35(46), 15307–15325. doi:10.1523/JNEUROSCI.2542-15.2015 PMID:26586819 Zhang, L., Hu, X. Z., Li, X., Li, H., Smerin, S., Russell, D., & Ursano, R. J. (2014). Telomere length a cellular aging marker for depression and post-traumatic stress disorder. Medical Hypotheses, 83(2), 182–185. doi:10.1016/j.mehy.2014.04.033 PMID:24875221

178

179

Chapter 11

Yoga for Mental Health Disorders: Research and Practice

Bhargav Hemant National Institute of Mental Health and Neurosciences, Bengaluru, India Rashmi Arasappa https://orcid.org/0000-0001-7864-8200 National Institute of Mental Health and Neurosciences, Bengaluru, India Inbaraj G. National Institute of Mental Health and Neurosciences, Bengaluru, India Kaviraja Udupa https://orcid.org/0000-0001-6906-5193 National Institute of Mental Health and Neurosciences, Bengaluru, India Shivarama Varambally https://orcid.org/0000-0003-1420-8606 National Institute of Mental Health and Neurosciences, Bengaluru, India

ABSTRACT This chapter is divided into four subsections: the first section of the chapter provides an update on current evidence for yoga therapy in common mental health disorders, the second section provides brief overview on neurophysiological abnormalities in psychiatric disorders and their relationship with psychological stress, the third section deals with summary of evidence for neurophysiological effects of yoga in mental health disorders, and the last section emphasizes on practical aspects of yoga therapy with details of clinically useful yoga practices for common mental health disorders. The mental health disorders covered in this chapter include depression, anxiety, schizophrenia, child and adolescent psychiatric disorders, substance use disorders, and geriatric psychiatric disorders. DOI: 10.4018/978-1-7998-3254-6.ch011

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga for Mental Health Disorders

INTRODUCTION Psychiatric disorders are common and cause significant disability. As per the recent National Mental Health Survey, the overall weighted prevalence of mental morbidity was 10.6 percent for current and 13.7 percent for a lifetime (Murthy et al., 2017). Anxiety and depressive disorders are two of the most prevalent mental health disorders affecting the general population, with a one-year prevalence of 10-20 percent. As per the US epidemiological survey, anxiety disorders are more common than depressive disorders with a lifetime prevalence of 28.8 percent. Severe mental disorders include Schizophrenia, Major Depressive Disorder, Bipolar Affective Disorder, and Substance use disorders. Usually, psychiatric disorders tend to have a chronic course and considerably affect social, occupational, and personal functioning and increase health care costs. Though pharmacotherapy and psychotherapy form the mainstay of treatment, a significant proportion of patients seek traditional medicine either due to lack of desired response to conventional treatment, the stigma associated with such treatments, or side-effects of medications. This has led to patients employing traditional treatment modalities that could be self-administered, cost-effective, and enhance the sense of self-reliance. Yoga is a way of life that includes mental, physical, and spiritual attributes. According to Indian philosophy, Yoga has been defined as ‘Chitta Vritti Nirodhah’ (Patanjali Yoga Sutras, Chapter 1, Verse 2), a technique to quiet the mind or restrain the modifications of the mind (Iyengar, 1993). Yoga’s major mental health benefits are to calm the mind, enhance concentration, encourage self-acceptance, and reduce stress and anxiety. It has proven benefits in achieving, preserving, and promoting mental health (Balasubramaniam et al., 2013). Yoga-based interventions have claimed to be effective in treating mental health disorders, but the mechanisms involved in the treatment process are not fully understood. In the beginning, yoga research was limited only to investigations into the reduction of stress-related variables and improvement in cardiorespiratory health. The inclusion of new physiological and psychological parameters has improved the research platform to carry out adequate research in the field of mental disorders. The current chapter aims to widen the understanding of yoga in terms of research and practice in mental health disorders.

YOGA THERAPY FOR PSYCHIATRIC DISORDERS: CURRENT EVIDENCE Depression A depressive episode is characterized by feeling pervasively sad, having decreased interest in activities, depressive cognitions, death wishes, suicidal thoughts, ideas of guilt, decreased concentration/memory/ self-confidence/sleep, and appetite. In some cases, sleep and appetite may also be increased. In the past few years, there has been an increased interest in the use of yoga for the management of depressive disorders. In one of the earliest Randomized Controlled Trials (RCT), Sudarshan Kriya Yoga was found to be comparable to anti-depressant medications as a sole treatment in major depression (Janakiramaiah et al., 2000). In another study, yoga produced clinical improvement in depression comparable to medications, without significant adverse effects (Naveen et al., 2013). Yoga has been found to be effective even when used as a stand-alone treatment for mild to moderate depressive episodes. A recent study using yoga as a mono-therapy for depressed patients found that the effect size in reducing depression was large (Prathikanti et al., 2017). In another recent RCT, Iyengar yoga was tested for its impact on suicidal ideations 180

 Yoga for Mental Health Disorders

(SI) in depressed individuals. The preliminary evidence from this study suggests that this intervention may be safe in such patients and may reduce suicidal ideation (Nyer et al., 2018). A review looking at the efficacy of yoga for depression and anxiety concluded that yoga might be helpful for depression and anxiety, although the most substantial evidence was for unipolar depression (Uebelacker et al., 2016). In a systematic review, 12 RCTs were reviewed of which 3 RCTs had a low risk of bias. There was moderate evidence for yoga’s short-term effects compared to usual care concerning the severity of depression. The review concluded that despite methodological drawbacks in the studies, yoga could be considered an add-on treatment option for patients with depressive disorders (Cramer et al., 2013).

Anxiety Disorders Patients with anxiety disorders have symptoms of feeling anxious with worries related to various issues. These symptoms are associated with signs of autonomic arousal, usually. Phobic Disorder, Panic Disorder, Social Anxiety, Generalised Anxiety Disorder (GAD), Obsessive-Compulsive Disorder, and Post Traumatic Stress Disorder (PTSD) come under the rubric of Anxiety disorders. It is generally believed that yoga reduces anxiety, but systematic studies in this area are surprisingly scarce. A review that looked at the use of yoga in managing trauma-related depression, anxiety, PTSD, and physiological stress concluded that these problems could be managed through yoga and meditation (Telles et al., 2012). Six weeks of Kundalini Yoga enhanced Cognitive Behavioural Therapy (Y-CBT) showed improvement in state and trait anxiety, depression, panic, sleep, and quality of life in generalized anxiety disorder patients (Khalsa et al., 2015). Sudarshan Kriya Yoga (SKY) in out-patients with GAD brought about a significant reduction in anxiety symptoms with a response rate of 73 percent and the remission rate of 41 percent (Katzman et al., 2012). A recent systematic review and meta-analysis on RCTs concluded that yoga could be a useful adjunctive intervention for PTSD (Cramer et al., 2018).

Substance Use Disorders Substance use disorders are among the common disorders globally and are associated with considerable social costs. It consists of abuse or dependence on substances like nicotine, alcohol, cannabis, opioid, stimulants, etc. The course of substance use disorders is often protracted with relapses and remissions. In an early study, Shaffer et al. (1997) found that Hatha Yoga therapy was comparable to conventional methadone treatment combined with traditional group psychotherapy. A study in India used Sudarshan Kriya (SK) and pranayama (P) pranayamas are slow yogic breathing techniques aimed at achieving mastery over the breath and mind) in patients with nicotine dependence with cancer after they had completed standard therapy. The authors reported that SK and P helped to control the habit of tobacco consumption in 21 percent of individuals (Kochupillai et al., 2005). In a randomized controlled trial at NIMHANS, 60 patients with alcohol dependence admitted for detoxification were randomized into SK or treatment as usual (TAU) groups. Patients in the SK group had a higher reduction in depression scores compared to the TAU group. Plasma cortisol and ACTH levels dropped in both groups, more so in the SK group. The reduction in depression scores correlated with that of the plasma cortisol level in the SK group (Vedamurthachar et al., 2006). A study using 10-minute yogic breathing exercises in smokers showed acute effects in reducing craving compared to a control group who watched videos (Shahab et al., 2013).

181

 Yoga for Mental Health Disorders

In a study conducted in China, Zhuang et al. (2013) randomized Chinese women undergoing treatment for opioid dependence into an experimental group with yoga as an add on therapy against a control group with routine care. Significant differences were noted in mood and quality of life after six months in the yoga group. This suggests that the improvement in mood states and better affective regulation may help prevent relapse. Dhawan and others (2015) studied the effect of SKY among patients receiving buprenorphine for treatment of opiate dependence. Compared to the treatment-as-usual group, patients in the intervention group had better outcomes in the physical, psychological and environmental quality of life. The current literature recommends yoga as adjuvant therapy for use in substance use disorders to supplement the medically approved treatment approaches. Yoga can be administered both during the acute/detoxification state as well as during long term/ relapse prevention state.

Schizophrenia Schizophrenia is a psychotic illness where the patient loses touch with reality, and the symptoms consist of positive symptoms like delusions, hallucinations, disorganization of thought/ behaviour, negative symptoms like alogia, asocialization apathy, amotivation, anhedonia, and some cognitive deficits. The last decade has seen a substantial increase in the trials related to yoga in schizophrenia, many of them being RCTs. Yoga as an add-on to antipsychotic medications has been useful in improving negative symptoms and socio-occupational functioning (Varambally et al., 2012; Manjunath et al., 2013) and cognitive impairment (Lin et al., 2015; Bhatia et al., 2016). A pilot study has also shown improvement in positive symptoms (Visceglia et al., 2011). Yoga was also found useful in a significant improvement in facial emotion recognition ability, which is an indicator of social cognition. This was also associated with an increase in oxytocin levels in patients with schizophrenia (Jayaram et al., 2013). A recent review that examined the effects of yoga versus standard care in patients with schizophrenia included eight studies. It concluded that yoga may be beneficial to improve symptoms, social functioning, and quality of life but that the available evidence is insufficient to draw strong conclusions and that more systematic research is needed (Broderick et al., 2015).

Childhood Psychiatric Disorders The last two decades have seen a surge in the use of complementary and alternative medicine therapies in the pediatric population as childhood psychiatric illnesses pose a challenge for management through conventional means. This is due to both the nature of childhood psychiatric symptoms and the longterm adverse effects of medications. The maximum literature on the therapeutic application of yoga in child psychiatric disorders has been in Attention Deficit Hyperactivity Disorder (ADHD), followed by Autism Spectrum Disorders. ADHD is one of the most commonly diagnosed neuro-developmental disorders and affects nearly 6-7 percent of children and adolescents. ADHD is characterized by a combination of overactivity and poorly modulated behaviour with marked inattention and a lack of persistent task involvement. A study by Haffner showed that yoga training was superior to the conventional motor training on both test scores on an attention task and parent ratings of Attention Deficit Hyperactivity Disorder symptoms, with effect sizes in the medium-to-high range (0.60–0.97) (Haffner et al., 2006). A 6-week multimodal peer-mediated behavioural program that included yoga was given to 6 - 11 year old children diagnosed with Attention Deficit Hyperactivity Disorder. After six weeks of the program, 90.5 percent of them 182

 Yoga for Mental Health Disorders

showed reductions in the performance-impairment score, a measure of impaired academic performance. This improvement was sustained through the 12 months in 85 percent of the students with weekly one session. Most (92 percent) of the students also had improvements on the Vanderbilt ADHD rating scale scores as assessed by parents (Mehta et al., 2012). Autism spectrum disorders (ASDs) are characterized by two significant areas of difficulty; social communication and repetitive, stereotyped behaviors. A 10-month program of 5-weekly sessions of Integrated Approach to Yoga Therapy (IAYT) and Applied Behavior Analysis (ABA) was given to 6 children with ASD. Improvement in children’s imitation skills, especially pointing to the body, postural, and oro-facial movements were noted. Parents reported change in the play pattern of these children with toys, peers, and objects at home (Radhakrishna et al., 2010). A Get Ready to Learn (GRTL) Yoga Program among Children with Autism Spectrum Disorders was implemented every school day for 16 weeks. These children showed significant differences in total Aberrant Behavior Checklist (ABC)–Community score compared with students in the control group (Koenig et al., 2012). In another study, 30 children who have ASD were given yoga practice, following which there was a significant reduction in the symptoms of autism after three months of practice (Deorari et al., 2014). An exploratory study concluded that structured yoga intervention could be conducted for a large group of ASD children with parents’ involvement, and it can be used as an alternative therapy to reduce the severity of symptoms of ASD children (Narasingharao et al., 2016). A Cochrane based review reviewed the evidence for improvement in core and related symptoms in patients with ASD concluded that there was minimal evidence that yoga interventions may improve core symptoms of ASD, with an Oxford Centre for Evidence-Based Medicine (CEBM) Score of Level 5 (McLeod et al., 2015).

Geriatric Psychiatric Disorders The rapid growth in the geriatric population brings more challenges for the psychiatric fraternity. The geriatric psychiatric issues that pose challenges include mild cognitive impairment, dementia, depression, late-onset psychosis, and late-life issues, including palliative care. Research in geriatric psychiatry has mainly focused on cognitive disorders, which are on the rise, especially in developing countries, with efforts to prevent mild cognitive disorders progressing to dementia. Eight weeks of yoga therapy practice in the elderly showed improvement in perceived stress, mood, sleep, and retrospective memory (Innes et al., 2012). A study from NIMHANS assessed the effects of yoga on the quality of life and sleep quality in healthy elderly individuals living in nursing homes. Subjects in the yoga group had significant improvement in all the domains of QOL and total sleep quality after controlling for the effect of baseline difference in education between the two groups (Hariprasad et al., 2013a). Further, the subjects in the yoga group showed significant improvement in immediate and delayed recall of verbal (RAVLT) and visual memory (CFT), attention and working memory (WMSspatial span), verbal fluency (COWA), executive function (Stroop interference) and processing speed (Trail Making Test-A) compared to the waitlist group at the end of 6 months, after correcting for corresponding baseline score and education (Hariprasad et al., 2013b). As part of this study, seven healthy elderly subjects who had received yoga intervention were taken up for Magnetic Resonance Imaging (MRI) scanning. It was done at baseline and six months after yoga. The analysis showed a small increase in hippocampal volume, indicating that yoga can have neuroplastic effects in the elderly (Hariprasad et al., 2013c). This critical area of research needs further systematic study, as there are no other specific interventions available for the elderly at risk for dementia. 183

 Yoga for Mental Health Disorders

NEUROPHYSIOLOGICAL BASIS OF YOGA IN MENTAL HEALTH DISORDERS Mental illness can be defined as a health condition that changes a person’s thinking, feeling, or behaviour (or all three) and causes distress and difficulty in functioning. There are many different mental illnesses; each affects a person’s thoughts, feelings, and/or behaviours in distinct ways (Bethesda, 2007). Recently, there is an increased interest in the use of a non-pharmacological approach in managing mental disorders. As the Patanjali Sutras (Chapter 1 Verse 2) says: “Yoga is the practice of quieting the mind” (Iyengar, 1993). Positive mental health is “a state of well-being in which every individual realizes his or her abilities, can cope with the normal stresses of life, can work productively and fruitfully, and contribute to his or her community (WHO, 2013)”.

Autonomic Abnormalities in Common Mental Health Disorders Psychiatric and neurological conditions are recognized as fundamentally linked to the central nervous system (CNS) dysfunction and sensorimotor signalling. There is also a growing appreciation of how cognitive, affective, and motivational processes (even self-representation) are based on the body’s internal physiological state, regulated neurally by autonomous nervous system (ANS). However, there remains a lack of research on the neural mechanisms by which the ANS shapes psychological and neurological symptoms. Autonomic dysfunction is often identified as an early sign of many psychiatric disorders, related not only to anxiety disorder but also to depression (Udupa et al., 2007a, JAD; Schumann et al., 2017) and schizophrenia (Cella et al., 2018). The majority of dysfunctions were in terms of increased sympathetic activity and reduced parasympathetic functions (Udupa et al., 2007b). These altered autonomic functions were modulated by different anti-depressant therapies, including yoga and ayurveda (Kishore et al., 2014). Yoga can also be applied to enhance cognitive skills (concentration, memory, and visuomotor speed) among depressed patients. Thus, yoga has beneficial effects in various domains in patients with mental health disorders. A wide range of research evidence has recorded shortened life expectancy in patients with schizophrenia. This excess mortality appears to be caused by cardiovascular pathology and, in particular, coronary heart disease (Osby et al., 2007). This change in cardiac health alters autonomic balance, exhibited in the form of increased heart rate, sweating, salivation, altered pupillary function, and temperature changes (Kraepelin, 1896). The majority of these features indicate increased sympathetic activity, reduced parasympathetic modulation, or both. Heart rate variability (HRV) is the beat-to-beat oscillation of the mean value of RR intervals. It is the result of complex regulatory mechanisms by which the autonomous nervous system regulates heart rate and maintains cardiovascular parameters within physiological limits. The time and frequency domain parameters of HRV show reduced vagal activity in schizophrenia (Bär et al., 2005). Baroreflex sensitivity (BRS) evaluation is a proven method for assessing cardiovascular autonomic regulation. This reflex is one of the homeostatic mechanisms of the body for maintaining blood pressure at almost constant levels. The baroreflex creates a negative feedback loop, in which the heart rate is reflexively reduced by higher blood pressure. In contrast, decreased blood pressure reduces the activation of baroreflexes and increases heart rate. A reduction in this BRS has been reported in cardiovascular disorders and common mental disorders (Bertinieri, 1985). Yoga hypothesized to modulate BRS and related autonomic dysfunctions towards normalization with regular practice.

184

 Yoga for Mental Health Disorders

Stress and Its Relation with Common Mental Health Disorders Stress is a significant factor affecting the mental health of a person irrespective of age. For over two decades, researchers in several fields have been interested in understanding how socially induced stress presents psychological distress, mental illness symptoms, other social dysfunction or health problems. Yoga not only helps the person perceive the stress effectively without entangled factors but also responds without exaggeration of the sympathoadrenal axis. Thus, stress can be effectively countered with yogic training so that a vicious cycle of stress is curtailed. In the light of ample evidence that stress has a causal relationship to many chronic diseases (i.e., cardiovascular conditions, cancer, stroke, diabetics, etc.), yoga can be an appropriate intervention to reduce the negative emotional effects of stress, thus reducing the development of psychosomatic disorders. Although medicine-based remedies provide immediate relief, it exposes the patients to various adverse side effects, such as inflammation and oxidative stress (Trivedi et al., 2006). In contrast, yoga offers promising treatment possibilities for several psychological disorders without potential side effects. Several yoga-based interventions have been developed to relieve symptoms of major depressive and anxiety disorders, autism spectrum, and eating disorders.

Neurophysiological Effects of Yoga in Mental Health Disorders: Current Evidence Many modes of investigations in terms of electrophysiology (EEG, ECG based cardiac autonomic functions), neurochemical measures of hormones, neurotransmitters, and imaging have enabled us to understand the neurophysiological effects of yoga in mental health. These measures have shown how the yogic practitioner achieves higher inhibitory control of the system by enhanced slow-wave activity, parasympathetic functions, and inhibitory neurotransmitter levels. Further, the assessment of salivary alpha-amylase (sAA) activity is a simple method to assess acute sympathetic response or stress. The increase of sAA during stress is directly associated with norepinephrine levels (Rohleder, 2004). Yoga helps in reducing stress reflected in the form of reduced sAA levels (Gururaja et al., 2011). Autonomic dysfunction is most likely the result of long-lasting stressful experiences known to be associated with psychological conditions and an underlying genetic predisposition to autonomic dysfunction observed in relatives of patients (Bär, 2015). Evidence indicates that yoga is effective in mitigating the harmful effects of anxiety, stress, and depression by enhancing positive impact on feel-good neurotransmitters, alpha-wave activity, vagal tone, serum prolactin level, reduced oxidative stress, serum lipids, serum cortisol and the downregulation of the hypothalamic-pituitary-adrenal axis (Madanmohan et al., 2013). Studies show that yoga, along with conventional care, reduces symptoms associated with schizophrenia (hallucination, paranoia, motivation, emotional flattery, and ADHD) (Trivedi et al., 2006). In autistic children, yoga practice can improve self-sense, cognitive abilities, and communicative social behaviour.

Possible Psycho-physiological Mechanisms of Actions of Yoga Various researchers hypothesize that yoga functions by positively influencing the nervous system, cardiovascular system, psycho-neuro-endocrine axis, neurotransmitter levels, molecular and gene expression. The subsequent section would elucidate this process briefly. 185

 Yoga for Mental Health Disorders

Literature suggests that the vagal nerve, which is stimulated by yoga practice, causes an increase in parasympathetic function of the autonomic nervous system and increases GABA activity in the brain (Streeter et al., 2012). Studies comparing gene expression with controls in long-term yoga practitioners show that yoga positively influences gene expression profiles in immune cells (Saatcioglu et al., 2013). The regular practice of yoga has also been shown to induce brain changes, which result in greater activation of the left prefrontal cortex (Davidson et al., 2003; Bhargav et al., 2014). Several studies have shown that yoga-based practices are responsible for forming new neural pathways (Kreiman et al., 2000). Yogic practices affect resonance circuits which increase the thickness of the pre-front medial cortex and insula, particularly on the right side, which contributes to empathy, apprehension, logical and intuitive processing (Kreiman et al., 2000). Such activities often reduce stress, leading to improved mood and reduced emotional distress (Michalsen et al., 2007). The research studies on yoga have focused on autonomic mechanisms (Deepak, 2002a), neurophysiologic mechanisms (Deepak, 2002b), neurochemical mechanisms (Balasubramaniam et al., 2012), cognitive mechanisms (Deepak, 2002b) and spiritual mechanisms. Autonomic models appear to be very useful in explaining cardiorespiratory effects and stress pathophysiologies in diseases such as diabetes, hypertension, and myocardial infarction (Deepak, 2002a). The neurophysiological models help explain the role of yogic treatment in neurologic disorders, including epilepsy (predominant non-stress effects) (Deepak, 2002b). Neurochemical mechanisms appear to work best in explaining the effects of yoga in psychological disorders such as depression, schizophrenia, ADHD, and others. Studies have shown the biochemical basis of yoga in managing depression, i.e., increased brain-derived neurotrophic factors (BDNF) serum levels and reduction in cortisol levels (Naveen et al., 2013a; Naveen et al., 2013b). Similarly, yoga leads to increase in oxytocin levels in patients with schizophrenia, which has been linked to improved social cognition (Jayaram et al., 2013).

YOGA FOR MENTAL HEALTH DISORDERS: PRACTICAL ASPECTS Yoga is recognized by the National Institutes of Health (NIH) National Center for Complementary and Alternative Medicine (NCCAM) as a form of CAM in the category of “mind-body” medicine (NCCAM, 2010). Yoga therapy is defined by the International Association of Yoga Therapists (IAYT) as “the process of empowering individuals to progress toward improved health and well-being through the application of the philosophy and practice of yoga.” (Taylor, 2007).

Yoga Practices for Depression Patanjali yoga sutras describe that chanting and meditating on the sound “Om” can be a useful practice for reducing agitations of the mind (Chitta Vikshepa; PYS: 1.27-28) (Iyengar, 1993). An fMRI study on twelve healthy subjects conducted at NIMHANS, Bengaluru demonstrated that loud chanting of the sound “Om” for 10 seconds (in the ratio 1:2 between “O” and “Mmm”) parts produced significantly higher limbic deactivation than controlled breathing without such sound (Kalyani, 2011). The vibratory component of the humming sound “mmm” towards the end of “Om” chanting may be producing similar effects as those produced by stimulation of the auricular branch of the vagus nerve (which is a known treatment modality for depression). Thus, loud chanting of the sound “Om” may be a useful practice for depression. Other hatha yoga texts have advocated the use of breathing techniques for energizing the 186

 Yoga for Mental Health Disorders

body and mind. E.g., Gheranda Samhita (Chapter 5, Verse 77) describes that performing practice of bellows breath (bhastrika) for three sets with 20 strokes in each set has energizing effects and it prevents diseases (Saraswati, 2012). Sudarshan Kriya Yoga (SKY), which has bellows breathing as its primary component, has been demonstrated to have anti-depressant effects equivalent to those produced by the anti-depressant Imipramine (Janakiramaiah et al., 2000). Authors suggested that the effect of bhastrika may be due to hyperventilation, which may produce similar effects to those produced by the ECT. Another hatha yoga text called Hatha Yoga Pradipika (HYP) suggested that “corpse pose” or shavasana (corpse posture) and trataka (a yogic cleansing technique which involves focussing on a steady candle flame kept at a 1-meter distance in a dark room) help remove tiredness and fatigue (Muktibodhananda, 2012). HYP also describes that the ‘right nostril’ is connected to the ‘Sun channel’ in the body, and the ‘left nostril’ is connected to the ‘Moon channel’; the texts advocate that selective breathing through respective nostrils has activating or relaxing effects on the body. Interestingly, Raghuraj and colleagues (2008) have demonstrated a connection between sympathetic activation and ‘right nostril breathing’ and parasympathetic dominance with selective left nostril breathing. According to yogic understanding, there is over-activity of the ‘Moon channel’ in depression, and hence it is advisable to perform right nostril breathing for 5 minutes three times a day. A patient with depression is slow in his activities and assumes a slumping posture with a closed chest. Thus, it is essential to involve body postures, which generate confidence and energy. The yogic postures which focus on backbends and chest opening e.g., tree pose (vrikshasana), half wheel pose (ardha chakrasana), half-camel pose (ardha ustrasana), fish pose (matsyasana), warrior pose (veerbhadrasana) etc., may thus be useful. At NIMHANS, we have developed and validated a yoga program for depression. This module involves following practices: sun salutations (suryanamaskara), half-wheel posture (ardhachakrasana), camel posture (ustrasana), cobra pose (bhujangasana), right nostril breathing (suryanulomaviloma), victorious breath (ujjayi), bellows breath (bhastrika), skull shining breath (kapalabhati) and om chanting (pranavajapa). Patients with depression often have difficulty concentrating as they are lost in their thoughts. Thus, the therapist needs to maintain good communication and tempo of the yoga class. The therapist should be energetic, kind, and co-operative. He should avoid criticizing the subject and should give positive feedback more often. The therapist should motivate the patient towards yoga practice without pressing towards over-exertion or perfection in yoga class. In the initial phase, it is essential not to keep the practices slow. The practice should be dynamically changed with brief periods of relaxation in between. It is best to focus on physical postures, faster breathing techniques, and loud chanting of mantras in initial phases. As the patient grows in yoga practice, the postures can be maintained for a longer time, and slower breathing practices can be used. It is best to avoid highly contemplative meditative practices that emphasize withdrawal of the senses, at least in the initial phase of treatment.

Yoga Practices for Anxiety Patients with anxiety show signs of sympathetic over-activity; thus, the yoga program should focus on bringing a state of deep relaxation. A patient with anxiety would be more receptive to the yoga practices if, before starting the sessions, the therapist devotes a good amount of time listening to the patient’s difficulties and then explains how the therapy is going to help. For example, the concept of five layers of consciousness (physical layer, energy layer, a layer of emotions, a layer of intellect and layer of bliss) according to yoga philosophy (Panchakosha model = Five-layer model) can be utilized to explain how 187

 Yoga for Mental Health Disorders

yoga therapy connects and harmonizes different layers using yoga techniques, and how it uses physical postures to bring relaxation at the level of body, breathing techniques to slow down and regularise the breath, and meditation to calm down the mind, thereby allowing the layer of intellect and layer of bliss to manifest (Nagarathna & Nagendra, 2010a). In contrast to the practices advised for depression, in anxiety the practices should be gentle, slow and rhythmic. Difficult and risky postures, fast breathing practices and loud chanting may trigger anxiety. The postures should be comfortable with a steady base. Forward bending and movements with gravity are considered more relaxing in the yogic literature, thus postures like moon pose (shashankasana), wind releasing pose (pavanmukatasana), hand to feet pose (padhastasana), and crocodile pose (maksrasana) may be more useful. Similarly, the breathing practices should emphasize slower breathing with prolonged exhalation, e.g., sectional breathing with different mudras focusing on different lobes of the lungs. The breathing should be performed with inhalation: exhalation ratio of 1:3, left nostril breathing (chandra anuloma viloma) to activate the ‘Moon Channel’ (as explained above) and gentle humming breath (bhramari) to relax the mind through sound vibrations. Each breathing practice should be done for 5 minutes, three times a day (Nagarathna & Nagendra, 2010b). Meditative practices can be useful here with more of open monitoring component than intense focus, gentle, relaxing mantras such as “Om mani padme hum” or “Om” or simple, gentle humming “mmm” may also be added. Imagery, which is relaxing and calming (such as imagining cold water, cool moonlight, fresh flowers, etc.), may also be included. Yoga Nidra also can be a useful practice for relaxation in these patients (Rani et al., 2012).

Yoga Practices for Schizophrenia and other Psychotic Disorders Patients with psychotic disorders lose touch with reality. Though patients stabilized on medications can be taught yoga practices quite comfortably, those in the initial phases of illness and with acute exacerbations of psychotic episodes demand more understanding and care from the therapist. The first important thing to note is that a yoga therapist should not treat patients with psychotic illnesses without the supervision of a psychiatrist. Patients should always be encouraged to take medications regularly, and at no point should a therapist encourage the belief of managing the illness with yoga therapy only. The role of yoga therapy in such illnesses should be clearly understood. Yoga therapy can only be used as an add-on to medications, and it will help mostly in addressing the negative symptoms of psychotic illnesses such as lack of motivation, lack of socialization, emotional withdrawal, etc. There is no current evidence to show that yoga practice would reduce the patient’s aggressiveness, delusions, or hallucinations (positive symptoms). The role of caregivers is vital in psychotic illnesses. A therapist should always take the caregivers’ perspective of the patient’s illness independently before beginning the therapy. The therapist should take classes of such patients in the presence of a caregiver. Also, it would be ideal if the caregiver also practices yoga with the patient. The therapist should never enter into any kind of arguments with the patient and should always be kind and empathetic with the understanding that aggressiveness and agitation is a part of their illness. If the patient becomes disruptive on a particular day, the therapist should avoid taking sessions on that day and discuss with the caregivers about the presence of any stressors or triggers which may include lack of proper sleep, rest, food, or reduced compliance with medications. If such behaviour persists, then consultation with a psychiatrist should be sought. The NIMHANS Integrated Centre for Yoga has validated a yoga program for patients suffering from psychotic illnesses (Govindaraj et al., 2016). The program includes the first ten minutes of warm practices such as jogging, twisting, forward-backward bending, tiger breathing, and moon pose breathing. This 188

 Yoga for Mental Health Disorders

is followed by eight rounds of suryanamaskar (sun salutations) (first four rounds fast; next four rounds slow) and then brief relaxation in shavasana with abdominal breathing. After this, patients perform asanas in sitting (ardha-ustrasana, vakrasana), prone (bhujangasana, dhanurasana) and supine positions (sarvangasana, matsyasana). This is followed by fast breathing (bhastrika 20 counts two sets) and slow breathing (alternate nostril breathing). In the end, patients lie down in supine position and chant sounds aa, uu, mm for nine rounds each and then a-u-m together in a single breath for another nine rounds. They are asked to keep their eye open and be attentive of vibrations of the sounds on chest, neck, and head for a, u, and m, respectively. Antipsychotic medications may produce specific side effects such as stiffness, tremors, mask-like facial expressions, slowing down, etc. Some antipsychotics may cause weight gain and insulin resistance. The therapist should try to incorporate yoga practices, which are simple to learn and which may reduce these side effects (Verma et al., 2018). For example, whole-body joint loosening may help reduce the stiffness, dynamic suryanamaskar (sun salutations), and fast breathing practices (e.g., bhastrika or bellows breath) may help reduce weight, right nostril breathing may enhance metabolic rate and help improve cognitive functions. Patients with psychotic illnesses generally should not be advised meditative practices such as focussed or open-monitoring meditation or imagery because these practices may enhance or provoke hallucinations and precipitate acute psychotic episodes. As far as possible, patients should perform practices with eyes open. As the patient may tend to drift into imagination, the yoga therapist should keep interacting with the patient throughout the class. The practices should be intense but straightforward to keep the attention in the present moment; there should be a shift from one practice to another in quick succession with very brief periods of relaxation in between (not more than 2 minutes). During relaxation, the patient may keep the eye open and keep chanting activating mantras (such as suryanamaskar mantra, Gayatri mantra, aa, uu, mm chanting) loudly. Therapists should avoid touching the patient without prior information. The help of caregivers should be taken whenever necessary to modify the postures by touching. The therapist should always try to be interactive and be gentle and kind in her/his approach. Treating patients suffering from psychotic illnesses demands a good amount of patience, temperamental control, and kindness from the therapist.

Yoga Practices for Cognitive Impairment and Dementia Yoga practices need special modifications while adapting them for elderly patients and more so for patients with dementia. The yoga practices should be less intense and easy to learn. Wherever applicable, a chair should be used to support, and the poses should be modified with chair (Litchke et al., 2012; Litchke et al., 2014). The breathing practices should be slow and simple. E.g., alternate nostril breathing can be simplified by replacing it with left nostril breathing, followed by right nostril breathing. Complex meditative practices should be avoided, and loud mantra chanting and positive affirmations should be encouraged during relaxation. The duration of the session should not be too long. Initially, the focus should be on building a good rapport, and yogic games (krida yoga) may be utilized for this. The therapist should start with 30 minutes a day session for an initial 1 week and then gradually increase it to 40 minutes and 50 minutes in weeks two and three. The transition may be even slower if the patient is more agitated or is in the late stage of the disease. The therapist needs to understand the clinical symptoms in different stages of the disease: early, middle and late. The practices then should be modified as per the need. E.g., in the early stages of Alzheimer’s disease (AD), it is common for the patient to be more rigid in his schedule, and therefore it 189

 Yoga for Mental Health Disorders

is important that the same sequence and counts of the posture and breathing must be taught every day without much variation. This will also make it easy for the patient to remember. In the middle stage of the disease, patients are more playful, and it is in this stage that yogic games, laughter-yoga, music, and mantras can be added to the class. In the late stages, the patient may lack initiative and may even have difficulty in seeing and hearing. Thus, in this stage, poses should be done with better physical proximity and increased touch. Breathing practices may be performed with help from the caregiver/therapist. There is a growing body of research evidence that suggests that enrichment of mind-body practices with a multi-sensory approach is associated with better improvements in behaviour and cognition of patients with dementias (Arkin, 2007; Nascimento et al., 2012; Fitzsimmons et al., 2002). Yoga should thus incorporate multi-sensory components of touch, visual modelling to enhance imitation skills, and auditory stimulation through music or rhythmic gesture songs (Mathews et al., 2001). Those in the late stages of the disease may need guidance from therapist through physical touch (in the presence of the caregivers) and guidance in the posture. After initial guidance into the posture for the first few rounds, the patient should be encouraged to do it on his own. Similarly, visually the poses should be shown in such a way that they are connected to some imagery which patient can readily understand, e.g., tree pose can include movement of hands (as if branches are moving in air), or cat pose can be accompanied by forward-backward movements mimicking a cat. This can further include an auditory stimulation of actual meowing like a cat etc. In this way, physical, visual, and auditory stimulations should be incorporated into the yoga practice. Adding some mantras and sounds of aa, uu, mm, which synchronize with specific body movements can be useful. Similarly, the relaxation phase should include absolute positive affirmations such as “I am healthy and strong,” etc., to be pronounced loudly by the participant.

Yoga Practices for Psychiatric Disorders in Children and Adolescents Studies on children with attention difficulties have shown that parents’ involvement is important (Maddigan et al., 2003). Parent-facilitated yoga treatment has been found useful in improving parent-child relationships in children with ADHD, but parents should be trained sufficiently before continuing with home practices (Harrison et al., 2004). Training of at least ten supervised sessions is necessary (with regular supervised booster sessions) before starting home practice. Based on current evidence and our clinical experience, the following advice would be useful for a yoga therapist while delivering yoga therapy to children with attention problems (Kaley-Isley et al., 2010). 1) It is better to conduct classes in groups with same-aged peers (as well as their parents) in the same group as children enjoy them more; 2) The yoga class should not be monotonous: it should have variations in poses and breathing practices which should be enriched with soothing music and calming mantra chanting (Om, A, U, M etc.); 3) The practice should not be too difficult but they should pose some challenge: poses like tree pose (Vrikshasana), sun salutations (suryanamaskara), warrior pose (veerabhadarasana), camel pose (ustrasana), shoulder stand (sarvangasana), moon pose (shashankasana) breathing with humming bhramari chanting may be used; 4) Each session should follow a pattern of faster practices followed by slower practices and then relaxation; 5) Over a period of time children should be taught to maintain a pose for longer and longer time; 6) Focus on breathing should be emphasised during yogasana practice; 7) Kriyas like kapalabhati (or bhastrika if it is difficult to do kapalabhati) and trataka should be added in the yoga program; 8) Yoga class should start with a shorter duration of 20 - 30 minutes and then gradually, over a few weeks the time should be increased by 10 minutes every week upto 60 minutes; 9) Krida yoga (yogic games involving attention and focus) should be utilised to build the rapport with children and as a reward at 190

 Yoga for Mental Health Disorders

the end of the yoga sessions; 10) It is important to note that children should have sufficient sleep and rest before starting yoga otherwise getting up early in the morning for yoga without proper sleep in the previous night may make them more stressed and restless; 11) Therapist should communicate with children in their local language and should try to make the yoga class more interactive and enjoyable for the children by adding interesting facts related to yoga practices, telling stories which generate interest in yoga practices etc; 12) Therapist should be cautious not to compare yoga performances of the children: this may increase the risk of injury and may generate a sense of competition among children which may take away the joy from the practice and make the child more restless; 13) The practices should be done in such a way that they encourage children to cooperate and support each other. For example, making formations of yoga where each child contributes a part of an overall pose, say they make a lotus on the ground where each child becomes a petal (Mandalam yoga). This kind of intervention has been especially useful in children with autism spectrum disorders (Litchke et al., 2018). Evidence shows that approximately 20 sessions of yoga may be necessary to observe the beneficial effects of yoga in children with attention difficulties. For children with autism, a dedicated yoga teacher and parents’ combined effort can lead to an improvement in imitation skills after a regular yoga practice (at least 3 days per week of supervised sessions) of 6-9 months approximately. The benefit is likely to be more if yoga is practiced at home (after training the parents for at least ten sessions) and is supported by the classroom environment as well in between (Kaley-Isley et al., 2010).

Are There Any Adverse Effects Due to Yoga? Cramer et al. (2015) performed a systematic review and meta-analysis of 301 randomized controlled trials (RCTs) of yoga. They found that out of 301 RCTs, only 94 reported an adverse event. They did not find significant differences in frequency of intervention-related, non-serious, or serious adverse events and of dropouts due to adverse events between yoga and exercise or usual care. They concluded that yoga was as safe as exercise or usual care. However, there are case studies that have reported severe side effects of yoga. For example, in a case report, headstand pose (sirsasana) was associated with the development of compressive myelopathy (Ferreira et al., 2013). Another case series suggested a worsening of psychotic symptoms with meditation practices (Walsh et al., 1979). Scientific and yoga literature also warns against emotional disturbances associated with Kundalini awakening (which may resemble acute psychotic conditions, depression and anxiety etc.) mainly if certain yogic practices (Kundalini Yoga) are performed without proper supervision of a yoga teacher (Bhargav et al., 2015; Greenwell et al., 1995). Future studies need to include a proper assessment of adverse effects to scientifically confirm the safety of these interventions.

CONCLUSION The current research evidence suggests that yoga can be used as a safe and useful add-on therapy for several psychiatric disorders. It may also be used as a sole therapy in selected conditions. However, a more substantial evidence base needs to be generated, especially in the form of long-term prospective studies, to clinically establish yoga as a scientific treatment method for psychiatric disorders.

191

 Yoga for Mental Health Disorders

REFERENCES Arkin, S. (2007). Language-enriched exercise plus socialization slows cognitive decline in Alzheimer’s disease. American Journal of Alzheimer’s Disease and Other Dementias, 22(1), 62–77. doi:10.1177/1533317506295377 PMID:17534004 Balasubramaniam, M., Telles, S., & Doraiswamy, P. M. (2013). Yoga on our minds: A systematic review of Yoga for neuropsychiatric disorders. Frontiers in Psychiatry, 3, 117. doi:10.3389/fpsyt.2012.00117 PMID:23355825 Bär, K. J. (2015). Cardiac autonomic dysfunction in patients with schizophrenia and their healthy relatives–a small review. Frontiers in Neurology, 6, 139. doi:10.3389/fneur.2015.00139 PMID:26157417 Bär, K. J., Letzsch, A., Jochum, T., Wagner, G., Greiner, W., & Sauer, H. (2005). Loss of efferent vagal activity in acute schizophrenia. Journal of Psychiatric Research, 39(5), 519–527. doi:10.1016/j.jpsychires.2004.12.007 PMID:15992561 Bertinieri, G., Cavallazzi, A., Ferrari, A. U., Pedotti, A., & Mancia, G. (1985). A new approach to analysis of the arterial baroreflex. Journal of Hypertension. Supplement. Official Journal of the International Society of Hypertension, 3, S79–S81. Bethesda, M., D. (2007). National Institutes of Health. Biological Sciences Curriculum Study. NIH Curriculum Supplement Series [Internet]: National Institutes of Health (US); 2007. Information about Mental Illness and the Brain. Bhargav, H., Jagannathan, A., Raghuram, N., Srinivasan, T. M., & Gangadhar, B. N. (2015). Schizophrenia patient or spiritually advanced personality? A qualitative case analysis. Journal of Religion and Health, 54(5), 1901–1918. doi:10.100710943-014-9994-0 PMID:25543321 Bhargav, H., Nagendra, H. R., Gangadhar, B. N., & Nagarathna, R. (2014). Frontal hemodynamic responses to high frequency yoga breathing in schizophrenia: A functional near-infrared spectroscopy study. Frontiers in Psychiatry, 5, 29. PMID:24715879 Bhatia, T., Mazumdar, S., Wood, J., He, F., Gur, R. E., Gur, R. C., Nimgaonkar, V. L., & Deshpande, S. N. (2017). A randomised controlled trial of adjunctive Yoga and adjunctive physical exercise training for cognitive dysfunction in schizophrenia. Acta Neuropsychiatrica, 29(2), 102–114. doi:10.1017/ neu.2016.42 PMID:27514629 Broderick, J., Knowles, A., Chadwick, J., & Vancampfort, D. (2015). Yoga versus standard care for schizophrenia. Cochrane Database of Systematic Reviews, 9, CD012052. PMID:28956893 Cramer, H., Anheyer, D., Saha, F. J., & Dobos, G. (2018). Yoga for posttraumatic stress disorder–a systematic review and meta-analysis. BMC Psychiatry, 18(1), 72. doi:10.118612888-018-1650-x PMID:29566652 Cramer, H., Lauche, R., Langhorst, J., & Dobos, G. (2013). Yoga for depression: A systematic review and meta‐analysis. Depression and Anxiety, 30(11), 1068–1083. doi:10.1002/da.22166 PMID:23922209

192

 Yoga for Mental Health Disorders

Cramer, H., Ward, L., Saper, R., Fishbein, D., Dobos, G., & Lauche, R. (2015). The safety of Yoga: A systematic review and meta-analysis of randomized controlled trials. American Journal of Epidemiology, 182(4), 281–293. doi:10.1093/aje/kwv071 PMID:26116216 Davidson, R. J., Kabat-Zinn, J., Schumacher, J., Rosenkranz, M., Muller, D., Santorelli, S. F., Urbanowski, F., Harrington, A., Bonus, K., & Sheridan, J. F. (2003). Alterations in brain and immune function produced by mindfulness meditation. Psychosomatic Medicine, 65(4), 564–570. doi:10.1097/01. PSY.0000077505.67574.E3 PMID:12883106 Deepak, K. K. (2002). Neurophysiological mechanisms of induction of meditation: A hypotheticodeductive approach. Indian Journal of Physiology and Pharmacology, 46, 136–158. PMID:12500489 Deepak, K. K. (2002a). The role of autonomic nervous system in rapid breathing practices. In Proceedings: Science of Breath. International Symposium on Sudarshan Kriya, Pranayam and Consciousness (pp. 42-45). New Delhi: All India Institute of Medical Sciences. Deorari, M., & Bhardwaj, I. (2014). Effect of yogic intervention on autism spectrum disorder. Yoga Mimamsa, 46(3), 81. doi:10.4103/0044-0507.159744 Dhawan, A., Chopra, A., Jain, R., & Yadav, D. (2015). Effectiveness of yogic breathing intervention on quality of life of opioid dependent users. International Journal of Yoga, 8(2), 144. doi:10.4103/09736131.154075 PMID:26170596 Ferreira, M. A., & Galvez-Jimenez, N. (2013). Sirsasana (headstand) pose causing compressive myelopathy with myelomalacia. JAMA Neurology, 70(2), 268–268. doi:10.1001/jamaneurol.2013.589 PMID:23407666 Fitzsimmons, S., & Buettner, L. L. (2002). Therapeutic recreation interventions for need-driven dementia-compromised. American Journal of Alzheimer’s Disease and Other Dementias, 17(6), 367–381. doi:10.1177/153331750201700603 PMID:12501484 Govindaraj, R., Varambally, S., Sharma, M., & Gangadhar, B. N. (2016). Designing and validation of a yoga-based intervention for schizophrenia. International Review of Psychiatry (Abingdon, England), 28(3), 323–326. doi:10.3109/09540261.2016.1151404 PMID:27002649 Greenwell, B. (1990). Energies of transformation: a guide to the kundalini process. Shakti River Press. Gururaja, D., Harano, K., Toyotake, I., & Kobayashi, H. (2011). Effect of Yoga on mental health: Comparative study between young and senior subjects in Japan. International Journal of Yoga, 4(1), 7. doi:10.4103/0973-6131.78173 PMID:21654969 Haffner, J., Roos, J., Goldstein, N., Parzer, P., & Resch, F. (2006). The effectiveness of body-oriented methods of therapy in the treatment of attention-deficit hyperactivity disorder (ADHD): Results of a controlled pilot study. Zeitschrift fur Kinder- und Jugendpsychiatrie und Psychotherapie, 34, 37–47. doi:10.1024/1422-4917.34.1.37 PMID:16485612

193

 Yoga for Mental Health Disorders

Hariprasad, V. R., Koparde, V., Sivakumar, P. T., Varambally, S., Thirthalli, J., Varghese, M., ... Gangadhar, B. N. (2013b). Randomized clinical trial of yoga-based intervention in residents from elderly homes: Effects on cognitive function. Indian Journal of Psychiatry, 55(7), S357. doi:10.4103/0019-5545.116308 PMID:24049199 Hariprasad, V. R., Sivakumar, P. T., Koparde, V., Varambally, S., Thirthalli, J., Varghese, M., ... Gangadhar, B. N. (2013a). Effects of yoga intervention on sleep and quality-of-life in elderly: A randomized controlled trial. Indian Journal of Psychiatry, 55(7), S364. doi:10.4103/0019-5545.116310 PMID:24049200 Hariprasad, V. R., Varambally, S., Shivakumar, V., Kalmady, S. V., Venkatasubramanian, G., & Gangadhar, B. N. (2013c). Yoga increases the volume of the hippocampus in elderly subjects. Indian Journal of Psychiatry, 55(7), S394. doi:10.4103/0019-5545.116309 PMID:24049206 Harrison, L. J., Manocha, R., & Rubia, K. (2004). Sahaja yoga meditation as a family treatment programme for children with attention deficit-hyperactivity disorder. Clinical Child Psychology and Psychiatry, 9(4), 479–497. doi:10.1177/1359104504046155 Innes, K. E., Selfe, T. K., Brown, C. J., Rose, K. M., & Thompson-Heisterman, A. (2012). The effects of meditation on perceived stress and related indices of psychological status and sympathetic activation in persons with Alzheimer’s disease and their caregivers: A pilot study. Evidence-Based Complementary and Alternative Medicine, 2012, 927509. doi:10.1155/2012/927509 PMID:22454689 Iyengar, B. K. S. (1993). Light on the yoga sutras of Patanjali. Aquarian/Thorsons. Janakiramaiah, N., Gangadhar, B. N., Murthy, P. N. V., Harish, M. G., Subbakrishna, D. K., & Vedamurthachar, A. (2000). Anti-depressant efficacy of Sudarshan Kriya Yoga (SKY) in melancholia: A randomized comparison with electroconvulsive therapy (ECT) and imipramine. Journal of Affective Disorders, 57(1-3), 255–259. doi:10.1016/S0165-0327(99)00079-8 PMID:10708840 Jayaram, N., Varambally, S., Behere, R. V., Venkatasubramanian, G., Arasappa, R., Christopher, R., & Gangadhar, B. N. (2013). Effect of yoga therapy on plasma oxytocin and facial emotion recognition deficits in patients of schizophrenia. Indian Journal of Psychiatry, 55(7), S409. doi:10.4103/00195545.116318 PMID:24049210 Kaley-Isley, L. C., Peterson, J., Fischer, C., & Peterson, E. (2010). Yoga as a complementary therapy for children and adolescents: A guide for clinicians. Psychiatry, 7, 20. PMID:20877530 Kalyani, B. G., Venkatasubramanian, G., Arasappa, R., Rao, N. P., Kalmady, S. V., Behere, R. V., ... Gangadhar, B. N. (2011). Neurohemodynamic correlates of ‘OM’chanting: A pilot functional magnetic resonance imaging study. International Journal of Yoga, 4(1), 3. doi:10.4103/0973-6131.78171 PMID:21654968 Katzman, M. A., Vermani, M., Gerbarg, P. L., Brown, R. P., Iorio, C., Davis, M., ... Tsirgielis, D. (2012). A multicomponent yoga-based, breath intervention program as an adjunctive treatment in patients suffering from generalized anxiety disorder with or without comorbidities. International Journal of Yoga, 5(1), 57. doi:10.4103/0973-6131.91716 PMID:22346068

194

 Yoga for Mental Health Disorders

Khalsa, M. K., Greiner‐Ferris, J. M., Hofmann, S. G., & Khalsa, S. B. S. (2015). Yoga‐enhanced cognitive behavioural therapy (Y‐CBT) for anxiety management: A pilot study. Clinical Psychology & Psychotherapy, 22(4), 364–371. doi:10.1002/cpp.1902 PMID:24804619 Kishore, R. K., Abhishekh, H. A., Udupa, K., Thirthalli, J., Lavekar, G. S., Gangadhar, B. N., Raju, T. R., & Sathyaprabha, T. N. (2014). Evaluation of the influence of ayurvedic formulation (Ayushman-15) on psychopathology, heart rate variability and stress hormonal level in major depression (Vishada). Asian Journal of Psychiatry, 12, 100–107. doi:10.1016/j.ajp.2014.07.002 PMID:25219946 Kochupillai, V., Kumar, P., Singh, D., Aggarwal, D., Bhardwaj, N., Bhutani, M., & Das, S. N. (2005). Effect of rhythmic breathing (Sudarshan Kriya and Pranayam) on immune functions and tobacco addiction. Annals of the New York Academy of Sciences, 1056(1), 242–252. doi:10.1196/annals.1352.039 PMID:16387692 Koenig, K. P., Buckley-Reen, A., & Garg, S. (2012). Efficacy of the Get Ready to Learn yoga program among children with autism spectrum disorders: A pretest–posttest control group design. The American Journal of Occupational Therapy, 66(5), 538–546. doi:10.5014/ajot.2012.004390 PMID:22917120 Kraepelin, E. (1896). Lehrbuch der psychiatrie. Barth. Kreiman, G., Koch, C., & Fried, I. (2000). Category-specific visual responses of single neurons in the human medial temporal lobe. Nature Neuroscience, 3(9), 946–953. doi:10.1038/78868 PMID:10966627 Lin, J., Chan, S. K., Lee, E. H., Chang, W. C., Tse, M., Su, W. W., ... Khong, P. L. (2015). Aerobic exercise and Yoga improve neurocognitive function in women with early psychosis. NPJ Schizophrenia, 1(1), 15047. doi:10.1038/npjschz.2015.47 PMID:27336050 Litchke, L. G., & Hodges, J. S. (2014). The meaning of” now” moments of engagement in yoga for persons with alzheimer’s disease. Therapeutic Recreation Journal, 48(3), 229. Litchke, L. G., Hodges, J. S., & Reardon, R. F. (2012). Benefits of chair yoga for persons with mild to severe Alzheimer’s disease. Activities, Adaptation and Aging, 36(4), 317–328. doi:10.1080/01924788 .2012.729185 Litchke, L. G., Liu, T., & Castro, S. (2018). Effects of multimodal mandala yoga on social and emotional skills for youth with autism spectrum disorder: An exploratory study. International Journal of Yoga, 11(1), 59. PMID:29343932 Madanmohan, Bhavanani, A. B., Zeena, S., Vithiyalakshmi, L., & Dayanidy, G. (2013). Effects of eight week yoga therapy program on cardiovascular health in hypertensives. Indian Journal of Traditional Knowledge, 12, 535–541. Manjunath, R. B., Varambally, S., Thirthalli, J., Basavaraddi, I. V., & Gangadhar, B. N. (2013). Efficacy of Yoga as an add-on treatment for in-patients with functional psychotic disorder. Indian Journal of Psychiatry, 55(7), S374. doi:10.4103/0019-5545.116314 PMID:24049202 Mathews, R. M., Clair, A. A., & Kosloski, K. (2001). Keeping the beat: Use of rhythmic music during exercise activities for the elderly with dementia. American Journal of Alzheimer’s Disease and Other Dementias, 16(6), 377–380. doi:10.1177/153331750101600608 PMID:11765863

195

 Yoga for Mental Health Disorders

McLeod, G. F., Warren, N. J., Warthen, J., Truleove, J. S., Ross, C. P., & Snook, C. A. (2015). Yoga as an intervention for patients with autism spectrum disorder: A review of the evidence and future directions. Autism-Open Access, 5(3), 155. Mehta, S., Shah, D., Shah, K., Mehta, S., Mehta, N., Mehta, V., Mehta, V., Mehta, V., Motiwala, S., Mehta, N., & Mehta, D. (2012). Peer-mediated multimodal intervention program for the treatment of children with ADHD in India: One-year followup. ISRN Pediatrics, 2012, 419168. doi:10.5402/2012/419168 PMID:23316384 Michalsen, A., Grossman, P., Acil, A., Langhorst, J., Lüdtke, R., Esch, T., ... Dobos, G. (2005). Rapid stress reduction and anxiolysis among distressed women as a consequence of a three-month intensive yoga program. Medical Science Monitor, 11(12), CR555–CR561. PMID:16319785 Muktibodhananda, S. (2012). Hatha yoga pradipika. Sri Satguru Publications. Murthy, R. S. (2017). National mental health survey of India 2015–2016. Indian Journal of Psychiatry, 59(1), 21. doi:10.4103/psychiatry.IndianJPsychiatry_102_17 PMID:28529357 Nagarathna, R., & Nagendra, H. R. (2010a). Yoga for anxiety and depression. Swami Vivekananda Yoga Prakashana. Nagarathna, R., & Nagendra, H. R. (2010b). Yoga for promotion of positive health. Swami Vivekananda Yoga Prakashana. Narasingharao, K., Pradhan, B., & Navaneetham, J. (2016). Sleep disorder, gastrointestinal problems and behaviour problems seen in autism spectrum disorder children and yoga as therapy: A descriptive review. Journal of Clinical and Diagnostic Research: JCDR, 10(11), VE01. doi:10.7860/JCDR/2016/24175.8922 PMID:28050484 Nascimento, C., Teixeira, C. V., Gobbi, L. T., Gobbi, S., & Stella, F. (2012). A controlled clinical trial on the effects of exercise on neuropsychiatric disorders and instrumental activities in women with Alzheimer’s disease. Brazilian Journal of Physical Therapy, 16(3), 197–204. doi:10.1590/S141335552012005000017 PMID:22499405 National Center for Complementary and Alternative Medicine. (2010). What is complementary and alternative medicine? https://nccam.nih.gov/health/whatiscam/ Naveen, G. H., Rao, M. G., Vishal, V., Thirthalli, J., Varambally, S., & Gangadhar, B. N. (2013b). Development and feasibility of yoga therapy module for out-patients with depression in India. Indian Journal of Psychiatry, 55(7, Suppl 3), S350. doi:10.4103/0019-5545.116305 PMID:24049198 Naveen, G. H., Thirthalli, J., Rao, M. G., Varambally, S., Christopher, R., & Gangadhar, B. N. (2013a). Positive therapeutic and neurotropic effects of Yoga in depression: A comparative study. Indian Journal of Psychiatry, 55(7, Suppl 3), S400. doi:10.4103/0019-5545.116313 PMID:24049208 Nyer, M., Gerbarg, P. L., Silveri, M. M., Johnston, J., Scott, T. M., Nauphal, M., ... Fava, M. (2018). A randomized controlled dosing study of Iyengar yoga and coherent breathing for the treatment of major depressive disorder: Impact on suicidal ideation and safety findings. Complementary Therapies in Medicine, 37, 136–142. doi:10.1016/j.ctim.2018.02.006 PMID:29609926

196

 Yoga for Mental Health Disorders

Ösby, U., Correia, N., Brandt, L., Ekbom, A., & Sparén, P. (2000). Mortality and causes of death in schizophrenia in Stockholm county, Sweden. Schizophrenia Research, 45(1-2), 21–28. doi:10.1016/ S0920-9964(99)00191-7 PMID:10978869 Prathikanti, S., Rivera, R., Cochran, A., Tungol, J. G., Fayazmanesh, N., & Weinmann, E. (2017). Treating major depression with Yoga: A prospective, randomized, controlled pilot trial. PLoS One, 12(3), e0173869. doi:10.1371/journal.pone.0173869 PMID:28301561 Radhakrishna, S. (2010). Application of integrated yoga therapy to increase imitation skills in children with autism spectrum disorder. International Journal of Yoga, 3(1), 26. doi:10.4103/0973-6131.66775 PMID:20948898 Raghuraj, P., & Telles, S. (2008). Immediate effect of specific nostril manipulating yoga breathing practices on autonomic and respiratory variables. Applied Psychophysiology and Biofeedback, 33(2), 65–75. doi:10.100710484-008-9055-0 PMID:18347974 Rani, K., Tiwari, S. C., Singh, U., Singh, I., & Srivastava, N. (2012). Yoga Nidra as a complementary treatment of anxiety and depressive symptoms in patients with menstrual disorder. International Journal of Yoga, 5(1), 52. doi:10.4103/0973-6131.91715 PMID:22346067 Rohleder, N., Nater, U. M., Wolf, J. M., Ehlert, U., & Kirschbaum, C. (2004). Psychosocial stress-induced activation of salivary alpha-amylase. Annals of the New York Academy of Sciences, 1032(1), 258–263. doi:10.1196/annals.1314.033 PMID:15677423 Saatcioglu, F. (2013). Regulation of gene expression by yoga, meditation and related practices: A review of recent studies. Asian Journal of Psychiatry, 6(1), 74–77. doi:10.1016/j.ajp.2012.10.002 PMID:23380323 Saraswati, S. N. (2012). Gheranda samhita. Yoga Publication Trust. Schumann, A., Andrack, C., & Baer, K. J. (2017). Differences of sympathetic and parasympathetic modulation in major depression. Progress in Neuro-Psychopharmacology & Biological Psychiatry, 79, 324–331. doi:10.1016/j.pnpbp.2017.07.009 PMID:28710030 Shaffer, H. J., LaSalvia, T. A., & Stein, J. P. (1997). Comparing Hatha Yoga with dynamic group psychotherapy for enhancing methadone maintenance treatment: A randomized clinical trial. Alternative Therapies in Health and Medicine, 3, 57–67. PMID:9210777 Shahab, L., Sarkar, B. K., & West, R. (2013). The acute effects of yogic breathing exercises on craving and withdrawal symptoms in abstaining smokers. Psychopharmacology, 225(4), 875–882. doi:10.100700213012-2876-9 PMID:22993051 Streeter, C. C., Gerbarg, P. L., Saper, R. B., Ciraulo, D. A., & Brown, R. P. (2012). Effects of Yoga on the autonomic nervous system, gamma-aminobutyric-acid, and allostasis in epilepsy, depression, and post-traumatic stress disorder. Medical Hypotheses, 78(5), 571–579. doi:10.1016/j.mehy.2012.01.021 PMID:22365651 Taylor, M. J. (2007). What is yoga therapy? An IAYT definition. Yoga Therapy in Practice, 3(3), 3.

197

 Yoga for Mental Health Disorders

Telles, S., Singh, N., & Balkrishna, A. (2012). Managing mental health disorders resulting from trauma through Yoga: A review. Depression Research and Treatment, 2012, 401513. doi:10.1155/2012/401513 PMID:22778930 Trivedi, M. H., Rush, A. J., Wisniewski, S. R., Nierenberg, A. A., Warden, D., Ritz, L., ... ShoresWilson, K. (2006). Evaluation of outcomes with citalopram for depression using measurement-based care in STAR* D: Implications for clinical practice. The American Journal of Psychiatry, 163(1), 28–40. doi:10.1176/appi.ajp.163.1.28 PMID:16390886 Udupa, K., Sathyaprabha, T. N., Thirthalli, J., Kishore, K. R., Raju, T. R., & Gangadhar, B. N. (2007a). Modulation of cardiac autonomic functions in patients with major depression treated with repetitive transcranial magnetic stimulation. Journal of Affective Disorders, 104(1-3), 231–236. doi:10.1016/j. jad.2007.04.002 PMID:17490754 Uebelacker, L. A., & Broughton, M. K. (2018). Yoga for depression and anxiety: A review of published research and implications for healthcare providers. Focus (San Francisco, Calif.), 16(1), 95–97. doi:10.1176/appi.focus.16104 PMID:32015704 Varambally, S., Gangadhar, B. N., Thirthalli, J., Jagannathan, A., Kumar, S., Venkatasubramanian, G., ... Nagendra, H. R. (2012). Therapeutic efficacy of add-on yogasana intervention in stabilized out-patient schizophrenia: Randomized controlled comparison with exercise and waitlist. Indian Journal of Psychiatry, 54(3), 227. doi:10.4103/0019-5545.102414 PMID:23226845 Vedamurthachar, A., Janakiramaiah, N., Hegde, J. M., Shetty, T. K., Subbakrishna, D. K., Sureshbabu, S. V., & Gangadhar, B. N. (2006). Anti-depressant efficacy and hormonal effects of Sudarshana Kriya Yoga (SKY) in alcohol dependent individuals. Journal of Affective Disorders, 94(1-3), 249–253. doi:10.1016/j. jad.2006.04.025 PMID:16740317 Verma, M., Bhargav, H., Varambally, S., Raghuram, N., & Gangadhar, B. N. (2018). Effect of integrated Yoga on antipsychotic induced side effects and cognitive functions in patients suffering from schizophrenia. Journal of Complementary and Integrative Medicine, 16(1). Visceglia, E., & Lewis, S. (2011). Yoga therapy as an adjunctive treatment for schizophrenia: A randomized, controlled pilot study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 17(7), 601–607. doi:10.1089/acm.2010.0075 PMID:21711202 Walsh, R., & Roche, L. (1979). Precipitation of acute psychotic episodes by intensive meditation in individuals with a history of schizophrenia. The American Journal of Psychiatry, 136(8), 1085–1086. doi:10.1176/ajp.136.8.1085 PMID:380368 World Health Organization. (2013). Mental health: a state of well-being. 2014. Report of the WHO Department of Mental Health. http://origin.who.int/features/factfiles/mental_health/en/ Zhuang, S. M., An, S. H., & Zhao, Y. (2013). Yoga effects on mood and quality of life in Chinese women undergoing heroin detoxification: A randomized controlled trial. Nursing Research, 62(4), 260–268. doi:10.1097/NNR.0b013e318292379b PMID:23715475 Zollman, C., & Vickers, A. (1999). What is complementary medicine? BMJ (Clinical Research Ed.), 319(7211), 693–696. doi:10.1136/bmj.319.7211.693 PMID:10480829

198

199

Chapter 12

Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders: From Suffering to Thriving Elin-Kristin Hem Olsen Drammen District Psychiatric Centre, Norway

ABSTRACT Healing after trauma is a long-lasting process involving the body and mind. The neurobiological foundation of trauma calls for more bodily and sensational, bottom-up regulatory approaches. Yoga has been proposed as a possible adjunctive treatment for trauma, and for more complex forms where talk-therapy have been proven insufficient. This chapter will give a theoretical and evidence-based summary of all (to our knowledge) relevant empirical data on yoga as a treatment for trauma-related disorders and symptoms.

INTRODUCTION Trauma-related disorders cause deep suffering and societal costs worldwide. The WHO World Mental Health surveys have demonstrated a life prevalence of PTSD across nations to be around 6 percent (Kessler & Ustun, 2008). The NICE Guidelines recognize that trauma may lead to other mental diseases or problems (e.g. depression, complex PTSD, addiction). The guidelines are stating that there is a lack of evidence regarding people with PTSD and complex needs. In previous years, many clinical psychologists and scientists have recognized that ordinary «talk» therapy may not be sufficient to treat traumatized patients (Bisson et al., 2013), and also that PTSD is not a sufficient diagnosis to embrace all possible outcomes of traumatic experiences (The National Institute for Health and Care Excellence, 2018; van der Kolk, 2006). The field of neurobiology has contributed to the latest understanding of trauma, especially regarding the acknowledgement of how essential bodily and sensational issues are to trauma patients. Mind-body therapies, such as sensorimotor psychotherapy (Ogden et al., 2006) and yoga and mindfulness, view DOI: 10.4018/978-1-7998-3254-6.ch012

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

the body and the mind as interconnected. These traditions treat mental illnesses in a more holistic way recognizing the importance of also healing the mind through the body (Caplan et al., 2013; Ogden et al., 2006; van der Kolk, 2006). In this perspective trauma treatment is not about treating an abnormal state of mind but rather increase internal awareness, insight and acceptance (Briere, 2015). In addition, the bottom-up approach of yoga is also taking into consideration that the top-down regulatory properties of the prefrontal cortex of is affected in traumatized individuals. In effect, trauma treatment should involve more focus on the body and arousal regulation together with the cognitive and emotional aspects (vander Kolk, 2006). In yoga philosophy, pain is an universal experience of all beings. Pain differs from suffering, which is often underlying mental health issues. The Yoga Sutras of Patanjali guide the practitioner through how to overcome the causes of suffering (e.g. the kleshas). There are in accordance to Patanjali five kleshas, or sources to suffering; ignorance, ego-centeredness, attachment, hatred, and fear of death. Traumatic events have an innate potential for suffering, where yoga can contribute to changes in how one relates to the traumatic reactions, the world and oneself. In essence, one can say it is through acceptance of pain and suffering as part of life, and a non-judgmental stance towards others and one’s reactions to suffering, healing can occur (Bryant, 2009). This chapter will give a theoretical and evidence-based summary of all (to our knowledge) relevant empirical data on yoga as a treatment for trauma-related disorders and symptoms. Interventions with meditation and mindfulness alone, and other body-mind techniques, are not included. Following in-depth reading of articles on the field, 16 studies were found eligible and included in this chapter review (Carter et al., 2013; Culver et al., 2015; Cushing et al., 2018; Descilio et al., 2010; Franzblau et al., 2008; Jindani et al., 2015; Johnston et al., 2015; Mitchell et al., 2014; Reddy et al., 2014; Reinhardt et al., 2017; Rhodes et al., 2016; Stoller et al., 2012; Telles et al., 2010; Thordardottir et al., 2014; Van der Kolk et al., 2014).

Trauma and its Sequelae Trauma is understood as an event which objectively would result in personal distress and an overwhelming psychophysiological emotional response, although many scholars understand the stress response as a trauma (Briere & Scott, 2006). Within the research field of trauma, different traumatic experiences have been categorized into two typologies (Herr, 1991). Type 1 trauma is a single event that occurs abruptly and unexpectedly and often promotes a strong fear response as the person experiences that one or others are at risk of life (APA, 2013). Examples of such traumas may be natural disasters, rape, and robberies. Type 2 trauma often occurs in early development stages, and in a relationship with close caregivers. These traumas can involve everything from the child’s recurring negative and unwarranted criticism, to more serious forms such as sexual and physical abuse. Research on psychological trauma has largely focused on type 1 trauma where PTSD is often the end state. Type 2 trauma has been referred to as relational trauma, often creating more complex symptomatology. In recent times, developmental or early childhood traumas have been used to describe this. This typology is important to highlight as they seem to have different outcomes and implications for treatment (Cloitre et al., 2012).

Posttraumatic Stress Disorder [PTSD] Traumatized individuals are known to respond inadequately to present stimuli which reminds them of the past traumatic event. These reactions must be considered as once necessary for survival in the mo200

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

ment of trauma but no longer are appropriate responses in the present moment (vander Kolk, 2006). The 5th Diagnostic Statistical Manual of Mental Disorders [DSM-5] (APA, 2013) defines PTSD with four groups of symptoms: Intrusion, avoidance, negative alterations in mood and cognitions, and alterations in reactivity and arousal. Previously PTSD was recognized as an anxiety disorder, now it is included in a new diagnosis chapter; Trauma- and stressors-related disorders. The definition includes a variety of physiological and psychological (e.g. emotional, cognitive) reactions following a traumatic event. PTSD is the only diagnosis which has an external source as an inclusion criterion. PTSD was first included in DSM-3 in the early 1980’s, in the aftermaths of the homecoming of returned Vietnam War soldiers. Earlier this diagnosis had been recognized in the literature as “shell shock”. Primarily the diagnosis aimed at adults experiencing single-trauma-events (e.g. type 1 trauma) as war, natural disasters, sexual assault, and accidents (van der Kolk et al., 1996). There is now strong evidence for effective treatment for PTSD. Best practice is considered to be therapies based on cognitive principles such as Cognitive processing Therapy [CPT], Trauma-Focused Behavioral Therapy [TF-CBT], and more exposure based therapies as prolonged exposure therapy [PE] (APA, 2017; Cusack et al., 2016; Watkins et al., 2018; Zammit et al., 2019), and Eye Movement Desensitization and reprocessing [EMDR] (Cusack et al., 2016; Watkins et al., 2018; Zammit et al. 2019). All these treatments are trauma-focused which implies a direct focus on trauma memories, trauma cognitions and emotion regulation. Non-trauma treatments are recommended when a person has not improved through trauma-specific methods, or if they are considered not stable enough to benefit from traditional trauma therapies (Zammit et al., 2019).

Complex PTSD Children exposed to trauma throughout their childhood may develop major impairments in several functional areas; biology, affect-regulation, dissociation, behavioral regulation, cognition, attachment and self-concept (Cloitre et al., 2009; van der Kolk et al., 1996). An extensive study conducted by van der Kolk and colleagues (1996) found that the complexity of how a person adapt to trauma may be more complex than the PTSD-model comprised, and that symptoms as dissociation, difficulties in affect-regulation and somatization can both be interrelated and independent of other PTSD symptoms (e.g. intrusive memories and disordered arousal and emotions). The last few decades there has been an extensive development in the understanding of childhood trauma, especially through the Adverse Childhood Experience-Study [ACE] (Felitti et al., 1998). This study has provided extensive correlational evidence on how childhood trauma as neglect, abuse, dysfunctional family environment holds great potential to induce persistent negative effects on children’s developmental health, and in the utmost consequence, how traumatized children become physically and mentally ill adolescents and adults. Therefore, scientists have called for a new diagnosis of trauma (e.g. Complex PTSD), which takes into consideration the width and more complex forms of possible symptoms caused by traumatic experiences (Cloitre et al., 2012; Green et al., 2000; Herman, 1992; van der Kolk, 2005). In 2012, Cloitre with colleagues was a part of a task force to develop guidelines on how to treat complex PTSD/developmental trauma. They argue that regular trauma therapies address emotional and cognitive processes which demands cognitive capacity and stability, which many of those with more complex trauma symptoms do not present. Many of these patients have great difficulties to comprehend and verbalize their symptomatic emotions and thoughts. In addition, recent developed trauma theories have underlined how trauma reactions become physiological conditioned; the body keeps the score (van 201

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

der Kolk, 2006, 2015; Ogden et al., 2006). The task force is calling for treatment principles beyond traditional trauma therapy (Cloitre et al., 2012), including more body-centered approaches (Ogden et al., 2006, van der Kolk, 2006). For a critical review of the current treatment- guidelines of Complex PTSD, see de Jongh with colleagues (2016).

Symptoms Beyond PTSD Historically the connection between trauma experiences and other psychiatric symptoms and diagnosis has been understated, even though how people react to a traumatic experience can differ extensively. The Colombo Twin and Singleton Study illustrated that the association between trauma and PTSD is not unique, and there are high rates of other diagnosis associated with trauma exposure (Dorrington et al., 2014). This notion is important when investigating whether yoga can work on trauma specific symptoms. Traumatic experiences may result in a wide-range of trauma-related disorders which often co-occur with PTSD or act on their own, as mood-disorder, anxiety disorders, eating disorders, somatoform disorders (Ogden et al., 2006), substance and alcohol abuse, and psychotic symptoms (Briere & Scott, 2006), while others experience the traumatic event and the psychological adaptations to it as becoming a part of who they are, and develop a personality disorder (Briere, 2015). The DSM-5 (American Psychiatric Associations [APA], 2013) has to a greater extent acknowledged the various reactions which can follow a traumatic event, and mood reactions have been implemented in the PTSD-model. It can be argued that trauma-related symptoms should be understood on a continuum, especially when considering the high comorbidity rate in PTSD.

Neurobiological Model of Traumatic Reactions Modern trauma theories and therapies are in a great extent founded on neurobiological findings in relation to brain development (Chen & Baram, 2016). Early childhood trauma has shown to have greater potential for lasting and extensive consequences due to the brain’s plastic properties early in the brain development (Teicher et al., 2016; van der Kolk, 2006). The neurobiological effects of trauma events alter processing of emotions and memory and can give rises to negative cognitions. Emotional processing theory emphasizes the role of the fear structure in the brain, including the HPA-axis, amygdala and hippocampus (Chen & Baram, 2016; Foa & Kozak, 1986). In PTSD the function of these structures is altered and become pathological. Simplified, the underlining of the theory is that the fear system is hyperactive and at the same time the regulating qualities of the prefrontal cortex is reduced, causing the traumatized individual to physically and emotionally overreact in situations associated with the traumatic event (Chen & Baram, 2016; Teicher et al., 2016; vander Kolk, 2006). This causes the long-lasting hyperarousal seen in traumatized individuals. The changes in memory processing and storage are thought to give rise to flashbacks and nightmares, and the re-experience of emotions related to the traumatic event. Associated memories operate as triggers which work automatically, unconsciously and bodily (Brewin, Dalgleish, & Joseph, 1996). Therefore, a person with trauma reactions can react with intense fear as if the trauma is happening in the present moment, often leading to the development of a core belief that the world in general is a dangerous place (Ehlers & Clark, 2000). Negative cognitive beliefs are known to be a result of traumatic experiences. Everyday life is being appraised through the lenses of specific experiences associated with the traumatic events. In sum the alteration of the emotional, memory and

202

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

cognitive system, causes the person to use coping strategies which can be harmful and prolong the duration of disease (ex. avoidance, social isolation, substance use, memory suppression).

Trauma and Yoga Interventions: The Status of the Scientific Field Yoga is a philosophical system which provides guidelines on how to live a holistic healthy life, both physically, emotionally, cognitively, and behaviorally. The most traditional understanding is based on Patanjali’s Yoga Sutra (Bryant, 2009), where yoga is the eight-limbed path to self-revelation and way of living. These 8 limbs are: yamas (ethical standards and personal integrity), niyamas (self-discipline and self-reflection), asana (the physical poses), pranayama (breath control), pratyahara (sensory withdrawal and objectively observance of ourselves), dharana (concentration), dhyana (meditation), and samadhi (state of ecstasy). Yoga as a whole practice is complex, and it can be hard to conceptualize it in research (Büssing et al., 2012; Elwy et al., 2014). There are a variety of different yoga styles are which based on different yoga philosophies or traditions. Many of these have roots in Patanjali’s eight limbs yoga tradition, although they do not contain all the eight limbs. Generally, one can say that yoga is physical poses (asanas), breathing exercises (pranayama), and meditation and relaxation (e.g. yoga nidra), and in addition the more philosophical and spiritual parts. In Western yoga traditions, the poses (asanas) are more prominent in a yoga practice than breathing techniques (pranayama) or meditation, all though these are a part of it as well. Breathing techniques and meditation follow and enhance the physical practice, and the three parts are intertwined. Hatha yoga (Hatha = Force) is a collection of practices which has a primary focus on the physical aspects of yoga. Hatha yoga has both physical and psychological effects, and it strengthen the mind-body connection hypothesized to be important in treating trauma. The breathing practices have a focus on different breathing techniques and using the breath as a mediational tool. Yoga breathing is thought to influence trauma symptoms through the autonomic nervous system (Telles et al., 2013). Some yoga forms have been specially designed to address trauma. One example is Sensory-enhanced hatha yoga (Stoller et al., 2012) which applies the use of props to enhance sensory stimuli (e.g. straps and wood blocks). The yoga form is developed for individuals exposed to extreme stress (e.g. war). The props are thought to enhance deep touch pressure and proprioceptive input to balance the nervous system for more relaxed and steady mood state.

Healing Trauma Through Yoga Healing after trauma is a long-lasting process involving the body and mind. The neurobiological foundation of trauma calls for more bodily and sensational, bottom-up regulatory approaches. Yoga has been proposed as a possible adjunctive treatment for trauma (van der Kolk et al., 2014), and for more complex forms where talk-therapy have been proven insufficient (Emerson, 2015). Yoga has proven to have positive outcomes on trauma-related symptoms, including depression, anxiety, stress, sleep issues, regulatory issues (Macy et al., 2018; Telles et al., 2012). This effect has been shown across different populations and trauma typologies; Military veterans (Cushing et al., 2018; Johnston et al., 2015; Mitchell et al., 2014; Reddy et al., 2015; Seppälä et al., 2014; Stoller et al., 2012), tsunami (Descilio et al., 2010) and flood survivors (Telles et al., 2010), earthquake survivors (Culvers et al., 2015; Telles et al., 2007; Thordardottir et al., 2014) women suffered from domestic violence (Franzblau et al., 2014), women with chronic PTSD (Price et al., 2017; Rhodes et al., 2016; van der Kolk et al., 2014), and homogenous group with PTSD (Jindani et al., 2015). 203

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

A systematic review of 25 randomized-controlled studies (RCT) suggests that yoga practice affects stress-related symptoms and more specifically improves the regulation of the sympathetic nervous system and hypothalamic-pituitary-adrenal system [HPA-axis], (Ross & Thomas, 2010). There is evidence that yoga has a positive influence on trauma related stress symptoms after natural disasters (Telles et al., 2007; Telles et al., 2010). The relation between yoga and stress seems to be mediated by enhancement of positive affect, and inhibition of posterior hypothalamus and cortisol secretion involved in the HPAaxis (Riley & Park, 2015). Yoga has also been shown to decrease symptoms related to depression and anxiety in different populations (Pascoe & Bauer, 2015). Yoga also seems to improve emotion regulation skills both in healthy individuals and for individuals with mental illness (Menezes et al., 2015). These improvements may further enhance the ability for self-compassion, increase mindfulness, body awareness and self-efficacy, and promote quality of life (Caplan, et al., 2013). Some benefits of yoga particularly address cognitions relating to trauma experiences. People with PTSD have great difficulties recognizing and relate to their inner sensations and feelings as they represent signs of danger and harm (Caplan et al., 2013; van der Kolk, 2006). Through yoga and body awareness traumatized individuals can cognitively reconstruct these automatically assumption. Sensations and emotions associated with the traumatic reactions can be differentiated and give rise to a corrective that they are not static phenomena. Hence, present an opportunity to challenge the trauma-induced cognitions of being “shattered”, “broken” or “the world is a dangerous place” (van der Kolk, 2006). One can also assume that yoga, as a philosophical practice would have the capacity to change persistent negative patterns of thought, behavior, and emotion.

Trauma-Sensitive Yoga for Complex Trauma Few studies have looked at childhood trauma, which often result in more complex trauma reactions (van der Kolk, 1996). Women and men exposed for interpersonal trauma as sexual or physical abuse may have more trouble relating to their body. People with developmental trauma experiences may have a harder time verbalizing their feelings and sensations, resulting in an experience of the traumatic event captured in the body. Yoga may help unlock these bodily experiences. Trauma-Sensitive Yoga (TSY), developed by David Emerson and Elizabeth K. Hopper (2011), is specialized for individuals who have experienced complex traumatization. Guidelines for trauma-sensitive yoga (Emerson et al., 2009) emphasize that trauma reactions may occur in different postures. Instructors allow each participant to feel safe in every posture, and at the same time they will be given an opportunity to withdraw from some highly emotional postures (e.g. associated with the trauma). They describe the main goal as letting people “tolerate that they are having a body” (Emerson et al., 2009, pp. 172). This contrasts with the notion of liking one’s body. TSY can be a stand-alone intervention or a part of an integrated therapeutic treatment. A review emphasizes that there is tentative evidence for TSY to improve symptoms of PTSD and depression, anxiety in individuals with traumatic experiences (Nolan, 2016). A qualitative study of 31 sexually or physically abused women during childhood used trauma-sensitive yoga as an intervention (West et al., 2016). The researcher found five central themes among the women’s experience of yoga; Grace and compassion, relation, acceptance, centeredness, and empowerment (G.R.A.C.E). Grace and compassion relate to the experience of be gentle and patient toward one’s body and the whole healing process. Relation tap into the ability to attune to internal sensations and external relationship. The women reported being more accepting of their body, self, history and the life as it is. The aspect of centeredness contains the shift towards more in place in the moment with both thinking 204

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

and feeling. The women also felt more empowered, being more in control of one’s life and trusting that they could cope with future obstacles (West et al., 2016). This study is an important contribution to the understanding of how yoga may improve individuals’ self-worth and self-believe.

SPECIFIC EFFECTS OF YOGA ON TRAUMA-RELATED SYMPTOMS Posttraumatic Stress Overall, the results of yoga improving symptoms of PTSD are promising, and most of the reviewed studies show that yoga may be as effective as other active interventions (e.g. dance, exposure) and are better than waiting lists. Out of the 15 studies reviewed in this chapter, 14 of them have investigated symptoms of posttraumatic stress or PTSD as primary outcome after a yoga intervention (Carter et al., 2013; Culver et al., 2015; Descilio et al., 2010; Jindani et al., 2015; Johnston et al., 2015; Mitchell et al., 2014; Price et al., 2017; Reddy et al., 2014; Reinhardt et al., 2018; Rhodes et al, 2016; Telles et al., 2010; Thordardottir et al., 2014; Van der Kolk et al., 2014). Different instruments and versions of an instrument have been used to investigate the effects of yoga on PTSD-symptoms. All though the evidence is promising, a recent meta-analysis refers to only weak recommendations for yoga as an adjunctive treatment for PTSD (Cramer et al., 2018). The review supports primarily physical yoga interventions, and the evidence regarding breathing techniques is considered low.

Depression and Anxiety Different reviews of the literature find that yoga can enhance positive affect and decrease symptoms of anxiety and depression (Balasubramaniam et al., 2013; Macy et al., 2018; Pascoe & Bauer, 2015). The studies of yoga and trauma-related symptoms are not necessarily focusing on depression and anxiety following trauma, but they do show promising results for yoga as a complementary treatment for these psychological disorders. In many studies measures of depression and anxiety are included in addition to measures of posttraumatic symptoms (Cushing et al., 2018; Descilio et al., 2012; Franzblau et al., 2008; Jindani et al., 2015; Mitchell et al., 2014; Rhodes et al., 2014; Seppälä et al., 2014). This emphasizes the notion that the effects of trauma experiences are diverse and widespread, illustrated by high comorbidity rates. A few of the studies observed a difference between the intervention group and the control group (Seppälä et al., 2014; Stoller et al., 2012). Some of the other studies found that both groups improved from pretest to posttest on symptoms of depression (Descilio et al. 2010; Franzblau et al., 2008; Jindani et al., 2015; Mitchell et al., 2014; Thordardottir et al., 2015; Van der Kolk et al., 2014) and anxiety (Jindani et al., 2015; Mitchell et al., 2014).

Quality of Life Even though it is important to recognize that the experience of trauma may cause PTSD, depression and anxiety, it is also important to acknowledge that individuals who have experienced trauma may have a good life despite challenges. Measurements of quality of life address the positive resources in being a human. Even though yoga does not heal trauma related disorders on its own it may result in improvements in quality of life. Four of the studies have investigated whether this is the case (Carter et al., 2013; 205

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Descilio et al., 2010; Stoller et al., 2012; Thordardottir et al., 2014). Decsilio and colleagues (2010) found that both the yoga intervention alone and together with exposure therapy improved the subject’s quality of life (p < 0.01) measured by the GHQ-12 questionnaire. Two of the studies did not find any difference between the control group and the intervention group, or any difference for each of the group from pretest to posttest (Carter et al., 2013; Thordardottir et al., 2014). Since only three studies measured quality of life, and they used different instruments as well, one cannot conclude that yoga improves quality of life or not. Future studies should include measurements of quality of life.

Mindfulness Trait Mindfulness is an attentional state where one is present in the moment in a non-judgmental and accepting manner (Kabat-Zinn, 1990). Human beings can be mindful in every second of their lives. The trait of mindfulness makes it possible to observe internal and external stimuli without acting out on those experiences or avoid them. A systematic review of the effects of mindfulness-based interventions on PTSD showed promising results, especially on mindfulness trait and acceptance of trauma memories (Banks et al., 2015). Mindfulness trait is theorized to have its effect on trauma-related symptoms through a shift to awareness instead of avoidance, improve emotion regulation, and increase tolerance for symptoms of posttraumatic stress (Folette et al., 2006). Three studies in this review examined whether yoga can enhance mindfulness trait (Cushing et al., 2018; Jindani et al., 2015, Johnston et al., 2015). Two of the studies assessed mindfulness trait with the Five-Facet Mindfulness Questionnaire (FFMQ). The FFMQ is a self-report questionnaire with 39 items. There are five components addressed by the FFMQ; observing, describing, acting with awareness, non-judging of inner experience, and non-reactivity to inner experience (Baer et al., 2008). The last study used the Mindful Attention Awareness Scale (MAAS). MAAS is used to measure mindfulness as attention awareness in everyday experiences through 15 questions rated on a 6-point scale (Black et al., 2012). The military service personnel in the study of Cushing and colleagues (2018) improved significantly on mindfulness, whilst Johnston and colleagues (2015) did not find improvements in mindfulness trait after the yoga intervention (p = 0.181), neither did the sample in the study of Jindani and colleagues (2015). These studies are not providing enough evidence for yoga to improve mindfulness trait among individuals with trauma-related symptoms or disorders.

Alcohol and Drug Abuse Posttraumatic stress may in many cases lead to excessive drinking or other drug abuse to avoid traumatic feelings and memories. It is therefore of interest to explore whether yoga may reduce this kind of risk behavior in individuals with PTSD or other trauma-related disorders. Reddy with colleagues (2014) investigated the effect of yoga on alcohol and drug abuse in 38 women with PTSD. The Alcohol Use Disorder Identification Test (AUDIT) and Drug Use Abuse Disorder Identification Test (DUDIT) were administered. The women were randomized either to the yoga intervention or to the control group. They were all tested at baseline, immediately after treatment, and 1 month after treatment. A linear mixed model showed that there was no difference between the intervention group and the control group in reducing substance abuse, though the mean scores for both AUDIT and DUDIT decreased in the yoga group and AUDIT scores for the control group. The authors conclude that yoga «may play a role in attenuating the symptoms of PTSD, reducing risk of alcohol and drug use» (Reddy et al., 2014), even though the results are scarce and based on a small sample size. In another study of 25 Vietnam veterans, who either 206

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

received a yoga intervention or were on a waiting list, their alcohol abuse was assessed (Carter et al., 2013). Neither the difference in mean scores from pretest to posttest nor between the two groups reached statistically significant levels. Regarding the results from these two studies one cannot conclude that yoga interventions have an impact on substance use or abuse. Future studies should further investigate the issue of substance abuse with larger samples and more robust research designs.

Resilience Resilience is in a broad term defined as the psychological and physical capacity to “bounce back” from adversity (Resnick et al., 2011). A person who is resilient will have the ability to keep on living a good life despite adversity, have a balanced view of life, would find a meaning to live, believe in oneself, and recognize that some experiences in life have to be handled alone (Wagnild, 2009). Resilience is associated with the yogic philosophical stance of acceptance, and especially acceptance of pain as an inevitable part of life. Suffering is a mental state where one fear the pain and act on that fear by avoiding the emotional pain and the situation leading to it. Individuals who are able to adapt to pain and challenges in such a way tend to exhibit more optimism and positive emotions, and are less vulnerable to stress, including traumatic stress (Resnick et al., 2011). Both Johnston and colleagues (2015) and Jindani and colleagues (2015) investigated whether a yoga intervention enhances resilience among individuals with PTSD, using the Resilience Scale. The Resilience Scale consists of five characteristics; 1) perseverance, 2) equanimity, 3) meaningfulness, 4) self-reliant, and 5) existential aloneness (Wagnild, 2009; Wagnild & Young, 1990). Johnston and colleagues (2015) did not find any difference between the yoga intervention and the treatment benchmark. Jindani and colleagues (2015) on the other hand found that the yoga group improved significantly compared to the control group. Still there are needed more research on whether yoga have effect on resilience.

Physiological Measures Trauma-related symptoms have shown to have physiological correlates (Kelly et al., 2018). Yoga may improve the physiological substrate underlying posttraumatic stress through both bottom-up and top-down processes (Taylor et al., 2010). Streeter with colleagues (2012) hypothesize that posttraumatic stress is a condition of stress, low heart rate-variability (HRV), and low gamma amino-butyric acid (GABA), and that yoga has its’ effect on these factors. In their study, Telles with colleagues (2010) measured HRV and breath rate. They did not find a statistical difference between the yoga group and the waiting list-group Another study measured startle response and respiration (Seppälä et al., 2014) using BIOPAC electrodes and belt. They found a statistically significant difference between the yoga group and the control group at 1-year follow-up between reduction in PTSD-symptoms and improvements in respiration. More research is needed to find the physiological basis for the effects of yoga. The autonomic nervous system and its regulatory characteristics seems to be specially involved (Pascoe & Bauer, 2015; Telles et al., 2012; Streeter et al., 2012). Streeter with colleagues (2007) found that yoga also may affect neurotransmitter systems correlated with posttraumatic symptoms (e.g. GABA). As the mind and the body is interconnected, yoga must affect each level of functioning.

207

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Insomnia Yoga can improve sleep (Khalsa, 2004), and low sleep quality is prevalent among people with posttraumatic stress disorder. One study explored whether yoga improves insomnia among individuals with PTSD (Jindani et al., 2015). Insomnia is a severe sleeping disorder often leading to decline in everyday functioning. The results showed a significant difference between the yoga group and the waitlist group on the Insomnia Severity Index. The results are promising regarding improving sleep quality in PTSD.

Affect Regulation Affect regulation is a capacity which people with trauma related disorders struggles with (van der Kolk et al., 2014). The Inventory of Altered Self Capacities (IoASC) identifies problems in affect regulation and emotional control. Van der Kolk and colleagues (2014) investigated affect regulation with IoASC. They did not find a significant difference between the yoga group and the control group, although there was a difference there illustrated by a small to medium effect size. Culver and colleagues (2015) did not find any statistically significant difference between the groups at follow-up on the Strengths and Difficulties Questionnaire measuring emotional and behavioral difficulties. Van der Kolk and colleagues (2014) conclude that yoga can be a tool for traumatized individuals to “tolerate physical and sensory experience associated with fear and helplessness and to increase emotional awareness and affect tolerance” (pp. 1).

LIMITATIONS OF THE SCIENTIFIC REALM OF YOGA AND TRAUMA Although the evidence is promising for yoga in trauma treatment, many of the studies have several limitations. Previous reviews have concluded that the field of yoga lack a robust empirical foundation in that most of the studies are of low methodological quality (Cramer et al., 2018; Elwy et al., 2014). The limitations are, as follows: 1. As one can see it is difficult to compare yoga interventions. The heterogeneity of yoga results in some implications and limitations regarding how extensive findings across different empirical studies can be interpreted and compared. 2. Summary of these limitations are: Small sample sizes, lack of randomization and adequate control groups, heterogeneity of yoga interventions and outcome variables, overlap between the constructs of mindfulness, meditation and yoga, lack of measures on long-term effects, and the absence of replication studies 3. Research on yoga are somewhat varying in showing that yoga is an effective intervention for traumarelated disorders or symptoms. In some studies, the yoga group had greater improvements than the passive control group. When the control group is active there seems to be less of a difference between the groups. 4. Overall, the sample sizes are small, and this may result in low power to detect difference between intervention and control groups. Only one study had a sample size larger than 100 subjects (Descilio et al., 2010). The other studies had a sample size ranged from 9 to 80 subjects. Small sample sizes reduce the generalizability of the study.

208

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

5. Homogenous samples also reduce the generalizability of the results. Homogeneity is demonstrated in type of trauma, age, gender, level of education or ethnicity. 6. Out of the 16 studies included, 11 of them are RCT-studies (Carter et al., 2013; Culver et al., 2015; Cushinger et al., 2016; Franzblau et al., 2008; Jindani et al., 2015; Mitchell et al., 2014; Reddy et al., 2014; Reinhardt et al., 2017; Seppälä et al., 2014; Stoller et al., 2012; van der Kolk et al., 2014). All of these have control groups, mostly assessment only or a waitlist group. To best investigate whether yoga has an effect one should include both an active control group (e.g. dance or other therapies) and a passive control group (e.g. waiting list). Only three of the studies included had this design (Culver et al., 2014; Descilio et al., 2010; Franzblau et al., 2008). The study of Descilio and colleagues (2010) did not use randomization to allocate the subjects to the different conditions. The active control groups were exposure therapy, giving testimony, and aerobic/dance. Exposure therapy and giving testimony are therapeutic, while aerobic/dance is more health promoting and has similar aspects to yoga. Whether the design include therapeutic interventions as a control condition rely on the research question. If the question concerns yoga as an alternative to therapeutic interventions, then the study should compare the yoga intervention with trauma related treatments (e.g. exposure therapy). If yoga is seen as an adjunctive part of trauma treatment one should compare it with other interventions that have similar characteristics (e.g. dance). The ideal design should implement both of these control conditions in addition to passive control groups, which can control for placebo effects. 7. The optimal strategy to reduce the risk of treatment bias in the different conditions is to blind both the participants and the researchers. In the case of yoga study this practice is not possible. 8. Primarily, self-report instruments are used to assess different outcomes. Self-report instruments are easy to administer and analyze, although the subjectivity of the answers has a limited value. To assess PTSD, CAPS is the best instrument. It is more time consuming as it involves clinical evaluations. As previously mentioned only three studies used CAPS to assess PTSD. Most of the other studies used PCL, a self-report instrument. To make the process of comparison across the field easier, researchers should attempt to use the same or similar instruments.

YOGA AND TRAUMA: A FUTURE PERSPECTIVE Even though most of the studies have small sample sizes, there is an indication that yoga may improve trauma-related disorders such as anxiety, depression, and PTSD. Since yoga is a time-effective and have low costs, it is an alternative approach when more time consuming and expensive treatment is not available or possible. Also, yoga may be an alternative to more complex needs and symptomatology where ordinary psychotherapy has not proven to be sufficient. It is difficult to conclude anything regarding the other measures such as resilience, mindfulness trait, quality of life, substance abuse, physiological measures, and affect regulation. However, collectively the evidence of yoga indicates that yoga may be an effective adjunctive treatment. Modern theories of trauma have recognized the importance of bodily and sensational approaches to trauma-related symptoms, in addition to or independent treatment. Research on yoga have contributed to insight on how psychological disorders take place in the body, and in which sense bottom-up approaches can be a purposeful entrance to treat psychiatric disorders and symptoms. Future research should seek to integrate yoga as an adjunctive and independent treatment to psychotherapy. More research on the more 209

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

complex forms of trauma is needed. Another topic of interest is yoga as a preventive tool to enhance resilience and coping strategies in children and adolescence. Receiving yoga in early childhood may vaccinate children against later traumatic reactions. As yoga is a low-cost intervention it suitable to be used in a more broader manner in the therapeutic landscape, especially, for those who live in a context where they do not have access to sufficient mental health care (ex. war zones, refugee camps, orphan homes). Yoga has a great potential to overcome the barriers of language and verbalization in therapy as everyone has a body. Future research should not underestimate the power of the more philosophical and spiritual parts of yoga and look further into how these aspects may be operate as therapeutic tools similar to regular psychotherapies.

REFERENCES Adkins, J. W., Weathers, F. W., McDevitt-Murphy, M., & Daniels, J. B. (2008). Psychometric properties of seven self-report measures of posttraumatic stress disorder in college students with mixed civilian trauma exposure. Journal of Anxiety Disorders, 22(8), 1393–1402. doi:10.1016/j.janxdis.2008.02.002 PMID:18436427 American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (5th ed.). American Psychological Association. (2017). Clinical Practice Guideline for the Treatment of Posttraumatic Stress Disorder (PTSD) in Adults. American Psychological Association. Baer, R. A., Smith, G. T., Lykins, E., Button, D., Krietemeyer, J., Sauer, S., ... Williams, J. M. (2008). Construct validity of the five facet mindfulness questionnaire in meditating and non-meditating samples. Assessment, 13, 27–45. doi:10.1177/1073191105283504 PMID:16443717 Balasubramaniam, M., Telles, S., & Doraiswamy, M. P. (2013). Yoga on our minds: A systematic review of yoga for neuropsychiatric disorders. Frontiers in Psychiatry, 3, 3. doi:10.3389/fpsyt.2012.00117 PMID:23355825 Banks, K., Newman, E., & Saleem, J. (2015). An overview of the research on mindfulness‐based interventions for treating symptoms of posttraumatic stress disorder: A systematic review. Journal of Clinical Psychology, 71(10), 935–963. doi:10.1002/jclp.22200 PMID:26192755 Bisson, J. I., Roberts, N. P., Andrew, M., Cooper, R., & Lewis, C. (2013). Psychological therapies for chronic post-traumatic stress disorder (PTSD) in adults. Cochrane Database of Systematic Reviews, 12, CD003388–CD003388. doi:10.1002/14651858.CD003388.pub4 PMID:24338345 Black, D. S., Sussman, S., Johnson, C. A., & Milam, J. (2012). Psychometric assessment of the mindful attention awareness scale (MAAS) among Chinese adolescents. Assessment, 2012(1), 42–52. doi:10.1177/1073191111415365 PMID:21816857 Briere, J., & Scott, C. (2006). Principles of trauma therapy: A guide to symptoms and, evaluation and treatment. SAGE Publications, Inc. Bryant, E. F. (2009). The yoga sūtras of patañjali: A new edition, translation and commentary. North Point Press.

210

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Büssing, A., Michalsen, A., Khalsa, S. B. S., Telles, S., & Sherman, K. J. (2012). Effects of yoga on mental and physical health: A short summary of reviews. Evidence-Based Complementary and Alternative Medicine, 2012, 165410–165410. doi:10.1155/2012/165410 PMID:23008738 Caplan, M., Portillo, A., & Seely, L. (2013). Yoga Psychotherapy: The integration of western psychological theory and ancient eastern yogic wisdom. Journal of Transpersonal Psychology, 45, 139–158. Carter, J. J., Gerbarg, P. L., Brown, R. P., Ware, R. S., D’Ambrosio, C., Anand, L., ... Katzman, M. A. (2013). Multi-component yoga breath program for Vietnam veteran Post Traumatic Stress Disorder: Randomized controlled trial. Journal of Traumatic Stress Disorders & Treatment, 2, 1–10. doi:10.4172/23248947.1000108 Chen, Y., & Baram, T. Z. (2016). Toward understanding how early-life stress reprograms cognitive and emotional brain networks. Neuropsychopharmacology, 41(1), 197–206. doi:10.1038/npp.2015.181 PMID:26105143 Clark, C. J., Lewis-Dmello, A., Anders, D., Parsons, A., Nguyen-Feng, V., Henn, L., & Emerson, D. (2014). Trauma-sensitive yoga as an adjunct mental health treatment in group therapy for survivors of domestic violence: A feasibility study. Complementary Therapies in Clinical Practice, 20(3), 152–158. doi:10.1016/j.ctcp.2014.04.003 PMID:25129883 Cloitre, M., Courtois, C. A., Ford, J. D., Green, B. L., Alexander, P., Briere, J., … van der Hart, O. (2012). The ISTSS expert consensus treatment guidelines for complex PTSD in adults. Retrieved from https:// www.istss.org/ISTSS_Main/media/Documents/ISTSS-Expert-Concesnsus-Guidelines-for-ComplexPTSD-Updated-060315.pdf Cloitre, M., Stolbach, B. C., Herman, J. L., van der Kolk, B., Pynoos, R., Wang, J., & Petkova, E. (2009). A developmental approach to complex PTSD: Childhood and adult cumulative trauma as predictors of symptom complexity. Journal of Traumatic Stress, 22(5), 399–408. doi:10.1002/jts.20444 PMID:19795402 Cramer, H., Anheyer, D., Saha, F., & Dobos, G. (2018). Yoga for posttraumatic stress disorder - a systematic review and meta analysis. BMC Psychiatry, 18(1), 72–72. doi:10.118612888-018-1650-x PMID:29566652 Culver, K., Whetten, K., Boyd, D. L., & O’Donnell, K. (2015). Yoga to reduce trauma-related distress and emotional and behavioral difficulties among children living in orphanages in Haiti: A pilot study. Journal of Complementary Medicine, 21(9), 539–545. doi:10.1089/acm.2015.0017 PMID:26090580 Cusack, K., Jonas, D. E., Forneris, C. A., Wines, C., Sonis, J., Middleton, J. C., Feltner, C., Brownley, K. A., Olmsted, K. R., Greenblatt, A., Weil, A., & Gaynes, B. N. (2016). Psychological treatments for adults with posttraumatic stress disorder: A systematic review and meta-analysis. Clinical Psychology Review, 43, 128–141. doi:10.1016/j.cpr.2015.10.003 PMID:26574151 Cushing, R. E., Braun, K. L., Alden, S. W., & Katz, A. R. (2018). Military tailored yoga for veterans with posttraumatic stress disorder. Military Medicine, 183, 223–231. doi:10.1093/milmed/usx071 PMID:29415222

211

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

de Jongh, A., Resick, P. A., Zoellner, L. A., van Minnen, A., Lee, C. W., Monson, C. M., Foa, E. B., Wheeler, K., Broeke, E., Feeny, N., Rauch, S. A. M., Chard, K. M., Mueser, K. T., Sloan, D. M., van der Gaag, M., Rothbaum, B. O., Neuner, F., de Roos, C., Hehenkamp, L. M. J., ... Bicanic, I. A. (2016). Critical analysis of the current treatment guidelines for complex PTSD in adults. Depression and Anxiety, 33(5), 359–369. doi:10.1002/da.22469 PMID:26840244 Descilo, T., Vedamurtachar, A., Gerbarg, P. L., Nagaraja, D., Gangadhar, B. N., Damodaran, B., Adelson, B., Braslow, L. H., Marcus, S., & Brown, R. P. (2010). Effects of a yoga breath intervention alone and in combination with an exposure therapy for post-traumatic stress disorder and depression in survivors of the 2004 South-East Asia tsunami. Acta Psychiatrica Scandinavica, 121(4), 289–300. doi:10.1111/j.16000447.2009.01466.x PMID:19694633 Dorrington, S., Zavos, H., Ball, H., McGuffin, P., Rijdijk, F., Sirib, S. A., ... Hotof, M. (2014). Trauma, post-traumatic stress disorder and psychiatric disorders in a middle-income setting: Prevalence and comorbidity. The British Journal of Psychiatry, 205(5), 383–389. doi:10.1192/bjp.bp.113.141796 PMID:25257062 Ehlers, A., & Clark, D. M. (2000). A cognitive model of posttraumatic stress disorder. Behaviour Research and Therapy, 38(4), 319–345. doi:10.1016/S0005-7967(99)00123-0 PMID:10761279 Elwy, A. R., Groessl, E. J., Eisen, S. V., Riley, K. E., Maiya, M., Lee, J., Sarkin, A., & Park, C. L. (2014). A systematic scoping review of yoga intervention components and study quality. American Journal of Preventive Medicine, 47(2), 220–232. doi:10.1016/j.amepre.2014.03.012 PMID:24996759 Emerson, D., & Hopper, E. (2011). Overcoming trauma through yoga. North Atlantic Books. Emerson, D., Sharma, R., Chaudry, S., & Turner, J. (2009). Yoga therapy in practice. Trauma-sensitive yoga: Principles, practice, and research. International Journal of Yoga Therapy, 19(1), 123–138. doi:10.17761/ijyt.19.1.h6476p8084l22160 Felitti, V., Anda, R., Nordenberg, D., Williamson, D., Spitz, A., Edwards, V., Koss, M. P., & Marks, J. (1998). Relationship of childhood abuse and household dysfunction to many of the leading causes of death in adults - The adverse childhood experiences (ACE) study. American Journal of Preventive Medicine, 14(4), 245–258. doi:10.1016/S0749-3797(98)00017-8 PMID:9635069 Follette, V., Palm, K. M., & Pearson, A. N. (2006). Mindfulness and trauma: Implications for treatment. Journal of Rational-Emotive & Cognitive-Behavior Therapy, 24(1), 45–61. doi:10.100710942-006-0025-2 Franzblau, S. H., Echeverria, S., Smith, M., & Van Cantfort, T. E. (2008). A preliminary investigation of the effects of giving testimony and learning yogic breathing techniques on battered women’s feelings of depression. Journal of Interpersonal Violence, 23(12), 1800–1808. doi:10.1177/0886260508314329 PMID:18319369 Herman, J. L. (1992). Complex PTSD – A syndrome in survivors of prolonged and repeated trauma. Journal of Traumatic Stress, 5(3), 377–391. doi:10.1002/jts.2490050305 Herr, J. (1991). Childhood traumas: An outline and overview. The American Journal of Psychiatry, 148(1), 10–20. doi:10.1176/ajp.148.1.10 PMID:1824611

212

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Jindani, F., Turner, N., & Khalsa, S. B. S. (2015). A yoga intervention for posttraumatic stress: A preliminary randomized control trial. Evidence-Based Complementary and Alternative Medicine, 2015, 351746–351746. doi:10.1155/2015/351746 PMID:26366179 Johnston, J. M., Minami, T., Greenwald, D., Li, C., Reinhardt, K., & Khalsa, S. B. S. (2015). Yoga for military service personnel with PTSD: A single arm study. Psychological Trauma: Theory, Research, Practice, and Policy, 7(6), 555–562. doi:10.1037/tra0000051 PMID:26010108 Kabat-Zinn, J. (1990). Full catastrophe living: Using the wisdom of your mind and body to face stress, pain, and illness. Delacorte. Karanicolas, P. J., Farrokhyar, F., & Bhandari, M. (2010). Blinding: Who, what, when, why, how? Canadian Journal of Surgery, 53, 345–348. PMID:20858381 Kelly, U. A., Evans, D. D., Baker, H., & Noggle Taylor, J. (2018). Determining psychoneuroimmunologic markers of yoga as an intervention of person diagnosed with PTSD: A systematic review. Biological Research for Nursing, 20(3), 343–351. doi:10.1177/1099800417739152 PMID:29130314 Kessler, R. C., & Ustun, T. B. (2013). The WHO world mental health surveys: global perspectives on the epidemiology of mental disorders. Cambridge University Press. Khalsa, S. B. (2004). Treatment of chronic insomnia with yoga: A preliminary study with sleep-wake diaries. Applied Psychophysiology and Biofeedback, 29(4), 269–278. doi:10.100710484-004-0387-0 PMID:15707256 Menezes, C. B., Dalpiaz, N. R., Kiesow, L. G., Sperb, W., Hertzberg, J., Oliveira, A. A., & Ventura, D. F. (2015). Yoga and emotion regulation: A review of primary psychological outcomes and their physiological correlates. Psychology & Neuroscience, 8(1), 82–101. doi:10.1037/h0100353 Mitchell, K. S., Dick, A. M., DiMartino, D. M., Smith, B. N., Niles, B., Koenen, K. C., & Street, A. (2014). A pilot study of a randomized controlled trial of yoga as an intervention for PTSD symptoms in women. Journal of Traumatic Stress, 27(2), 121–128. doi:10.1002/jts.21903 PMID:24668767 Nolan, C. R. (2016). Bending without breaking: A narrative review of trauma-sensitive yoga for women with PTSD. Complementary Therapies in Clinical Practice, 24, 32–40. doi:10.1016/j.ctcp.2016.05.006 PMID:27502798 Ogden, P., Minton, K., & Pain, C. (2006). Trauma and the body: A sensorimotor approach to psychotherapy. Norton & Company, Inc. Ogden, P., Pain, C., & Fisher, J. (2006). A sensorimotor approach to the treatment of trauma and dissociation. The Psychiatric Clinics of North America, 27(1), 263–279. doi:10.1016/j.psc.2005.10.012 PMID:16530597 Pascoe, M. C., & Bauer, I. E. (2015). A systematic review of randomized control trials on the effects of yoga on stress measures and mood. Journal of Psychiatric Research, 68, 270–282. doi:10.1016/j. jpsychires.2015.07.013 PMID:26228429

213

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Price, M., Spinazzola, J., Musicar, R., Turner, J., Suvak, M., Emerson, D., & van der Kolk, B. A. (2017). Effectiveness of an extended yoga treatment for women with chronic posttraumatic stress disorder. Journal of Alternative and Complementary Medicine (New York, N.Y.), 23(4), 300–309. doi:10.1089/ acm.2015.0266 PMID:28121466 Reddy, S., Dick, A. M., Gerber, M. R., & Mitchell, K. (2014). The effect of a yoga intervention on alcohol and drug abuse risk in veteran and civilian women with posttraumatic stress disorder. Journal of Alternative and Complementary Medicine (New York, N.Y.), 20(10), 750–756. doi:10.1089/acm.2014.0014 PMID:25211372 Resnick, B., Galik, E., Dorsey, S., Scheve, A., & Gutkin, S. (2011). Reliability and validity testing of the physical resilience measure. The Gerontologist, 51(5), 643–652. doi:10.1093/geront/gnr016 PMID:21402647 Rhodes, A., Spinazzola, J., & van der Kolk, B. (2016). Yoga for adult women with chronic PTSD: A long term follow-up study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 22(3), 189–196. doi:10.1089/acm.2014.0407 PMID:26863321 Riley, K. E., & Park, C. L. (2015). How does yoga reduce stress: A systematic review of mechanisms of change and guide to future inquiry. Health Psychology Review, 9(3), 379–396. doi:10.1080/174371 99.2014.981778 PMID:25559560 Seppälä, E. M., Nitchke, J. B., Tudorascu, D. L., Hayes, A., Goldstein, M. R., Nguyen, S. T., ... Davidson, R. J. (2014). Breathing-based meditation decreases Posttraumatic Stress Disorder symptoms in U.S. military veterans: A randomized controlled longitudinal study. Journal of Traumatic Stress, 27(4), 397–405. doi:10.1002/jts.21936 PMID:25158633 Stoller, C. C., Greuel, J. H., Cimini, L. S., Fowler, M. S., & Koomar, J. A. (2012). Effects of sensoryenhanced yoga on symptoms of combat stress in deployed military personnel. The American Journal of Occupational Therapy, 66(1), 59–68. doi:10.5014/ajot.2012.001230 PMID:22389940 Streeter, C. C., Gerbarg, P. L., Saper, R. B., Ciraulo, D. A., & Brown, R. P. (2012). Effects of yoga on the autonomic nervous system, gamma-aminobutryic-acid, and alleostasis in epilepsy, depression, and post-traumatic stress disorder. Medical Hypotheses, 78(5), 571–579. doi:10.1016/j.mehy.2012.01.021 PMID:22365651 Streeter, C. C., Jensen, J. E., Perlmutter, R. M., Cabral, H. J., Tian, H., Terhune, D. B., Ciraulo, D. A., & Renshaw, P. F. (2007). Yoga asana sessions increase brain GABA levels: A pilot study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 13(4), 419–426. doi:10.1089/acm.2007.6338 PMID:17532734 Taylor, A. G., Goehler, L. E., Galper, D. I., Innes, K. E., & Bourguignon, C. (2010). Top-down and Bottom-up mechanisms in mind-body medicine: Development of an integrative framework for psychophysiological research. Explore (New York, N.Y.), 6(1), 29–41. doi:10.1016/j.explore.2009.10.004 PMID:20129310

214

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Teicher, M. H., Samson, J. A., Anderson, C. M., & Ohasi, K. (2016). The effects of childhood maltreatment on brain structure, function and connectivity. Nature Reviews. Neuroscience, 17(10), 652–666. doi:10.1038/nrn.2016.111 PMID:27640984 Telles, S., Naveen, K. V., & Dash, M. (2007). Yoga reduces symptoms of distress in tsunami survivors of the Andaman Islands. Evidence-Based Complementary and Alternative Medicine, 4(4), 503–509. doi:10.1093/ecam/nem069 PMID:18227918 Telles, S., Singh, N., & Balkrishna, A. (2012). Managing mental health disorders resulting from trauma through yoga: A review. Depression Research and Treatment, 2012, 1–9. doi:10.1155/2012/401513 PMID:22778930 Telles, S., Singh, N., Joshi, M., & Balkrishna, A. (2010). Post-traumatic stress symptoms and heart rate variability in Bihar flood survivors following yoga: A randomized controlled study. BMC Psychiatry, 10(1), 18–18. doi:10.1186/1471-244X-10-18 PMID:20193089 The National Institute for Health and Care Excellence. (2018). Post-traumatic stress disorder. Retrieved from: https://www.nice.org.uk/guidance/ng116 Thordardottir, K., Gudmundsdottir, R., Zoëga, H., Valdimarsdottir, U. A., & Gudmundsdottir, B. (2014). Effects of yoga practice on stress-related symptoms in the aftermath of an earthquake: A community-based controlled trial. Complementary Therapies in Medicine, 22(2), 226–234. doi:10.1016/j.ctim.2014.01.008 PMID:24731893 van der Kolk, B. A. (2005). Developmental trauma disorder: Toward a rational diagnosis for children with complex trauma histories. Psychiatric Annals, 35(5), 401–408. doi:10.3928/00485713-20050501-06 van der Kolk, B. A. (2006). Clinical Implications of Neuroscience Research in PTSD. Annals of the New York Academy of Sciences, 1071(1), 277–293. doi:10.1196/annals.1364.022 PMID:16891578 van der Kolk, B. A. (2015). The body keeps the score: Brain, mind, and the body in the healing of trauma. Penguine Books. van der Kolk, B. A., Pelcovitz, D., Roth, S., Mandel, F. S., Mcfarlane, A., & Herman, J. I. (1996). Dissociation, somatization, and affect dysregulation: The Complexity of adaption to trauma. The American Journal of Psychiatry, 153(7), 83–93. doi:10.1176/ajp.153.7.83 PMID:8659645 van der Kolk, B. A., Roth, S., Pelcovitz, D., Sunday, S., & Spinazzola, J. (2005). Disorders of extreme stress: The empirical foundation of a complex adaptation to trauma. Journal of Traumatic Stress, 18(5), 389–399. doi:10.1002/jts.20047 PMID:16281237 van der Kolk, B. A., Stone, L., West, J., Rhodes, J., Emerson, D., Suvak, M., & Spinazzola, J. (2014). Yoga as an adjunctive treatment for Posttraumatic Stress Disorder: A randomized controlled trial. The Journal of Clinical Psychiatry, 75(06), 559–565. doi:10.4088/JCP.13m08561 PMID:25004196 Wagnild, G. (2009). A review of the resilience scale. Journal of Nursing Measurement, 17(2), 105–113. doi:10.1891/1061-3749.17.2.105 PMID:19711709 Wagnild, G., & Young, H. M. (1990). Resilience among older women. Image—the Journal of Nursing Scholarship, 22(4), 252–255. doi:10.1111/j.1547-5069.1990.tb00224.x PMID:2292448

215

 Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Watkins, L. E., Sprang, K. R., & Rothbaum, B. O. (2018). Treating PTSD: A review of evidencebased psychotherapy interventions. Frontiers in Behavioral Neuroscience, 12, 258–258. doi:10.3389/ fnbeh.2018.00258 PMID:30450043 West, J., Liang, B., & Spinazzola, J. (2016). Trauma sensitive yoga as a complementary treatment for posttraumatic stress disorder: A qualitative descriptive analysis. International Journal of Stress Management, 24(2), 173–195. doi:10.1037tr0000040 PMID:28458503

216

Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

APPENDIX 1 Table 1. Overview of the included studies in the scoping review Study

Carter et al., 2013           RCT

Culver et al., 2015 RCT

Sample

Treatment groups

N = 31 Male Veterans from the Vietnam War Chronic, treatmentresistant PTSD 1. Sudarshan Kriya Yoga (SKY) Age (mean): 58 All participants used 2. Control: Wait-list medications Comorbidity (e.g. anxiety, suicidality, ADHD, depression)

N = 76 Children living in two different orphanages in Haiti Age (mean): 11.10 Female (N= 32) Male (N = 44) PTSD (10%); Partial PTSD (38%) Multiple trauma experiences (43%) Earthquake as index trauma (43%); seeing someone being beaten up/shot/killed (8%) as index trauma

Cushing et al., 2018 Pre-post test

N = 18 Veterans from 9/11-conflicts with PTSD or sub-PTSD Women: 9 Men: 9 Age: 26 – 62

Franzblau et al., 2008 RCT

N = 40 10 in each condition Women exposed to domestic violence Afro-American (N=20) European American (N = 20) Age: 18-45

3. Hatha Yoga 4. Control: Dance/ Aerobic 5. Control: Wait-list

Intervention

Sudarshan Kriya Yoga consist of 4 different meditative breathing techniques. Including psychoeducation about stress reduction and group processes

1. The Yoga program was based on Hatha yoga and adjusted to the cultural frame of Haitian children. The intervention included postures, breathing techniques, and mindfulness meditation. 2. Aerobic dance was kid-friendly and involved routines choreographed to Latino music.

Vinyasa yoga

     Trauma-informed yoga based on the WAE-protocol suited for veterans. Provided by certified teachers.

1. Yoga breathing techniques (pranayama) 2. Testimony 3. Yoga breathing + Testimony 4. Control: Waiting-list

Yoga breath techniques includes controlling and being mindful towards one’s breath, and simple yoga postures (asanas) Testimony involves the women telling their story about their trauma

Duration

Outcome measures

Findings

Effect sizes: d = 2.9 (Difference between pretest and 6-month posttest scores 1. Primary outcome: on CAPS for the Yoga 2. PTSD group) Statistically significant 3. CAPS 4. PCL-M differences between 5. Secondary outcomes: the groups, with the 22 h over 5 yoga group showing      Depression succeeding days      CES-D more improvements, with 2 h per session      Quality of life in CAPS (p = 0.04),      WHOoL (Australian PCL-M (p = 0.00) and CES-D (p = 0.01) version)      Alcohol abuse scores. No statistically      AUDIT significant difference between the groups on AUDIT, and WHOoL scores

Limitations Small sample size No comparison between groups on 6 months follow-up since the control group received the intervention No active control conditions No blinding of participants Comorbidity No adjustments for multiple comparisons Medication given simultaneously Convenience and selfsample sampling

1. 2 yoga sessions per week for 8 succeeding weeks with 45 min per session 2. 2 aerobic dance sessions per week for 8 succeeding weeks with 45 min per session

Medium/small sample size Quasi-randomization (e.g. randomization within the two sites) Baseline differences as No statistically possible confounding significant difference variables 1. Primary outcome: between the groups on Regression to the mean 2. PTSD/trauma related PTSD-scores (p = 0.19) as a possible explanation symptoms or the EMBD-scores (p of improvements (those      UCLA PTSD-RI = 0.13) with high PTSD-scores 3. Secondary outcome: Post-hoc test showed a improve more than those 4. Emotional and statistically significant with low scores) behavioral difficulties difference between No blinding of 5. SDQ the yoga group and participants the wait-list group (p Not assessing for = 0.05) complex PTSD Lack of measures of long-term effects Low generalizability beyond children living in orphanages

     1 h session per week for 6 succeeding weeks

     1. Primary outcome:      2. PTSD      3. PCL-M      Secondary outcome:      Depression      PHQ-8      Anxiety      BAI      Sleep      PSQI      Mindfulness      MAAS

Effect size (Cohen’s d): 1.93 (PCL-M) 1.08 (PSQI) 1.32 (PHQ-8) 1.32 (BAI) - 1.00 (MAAS) Statistically significant improvements in PTSD and mindfulness measures from pre to post test. Clinically improvements in symptoms of anxiety and depression.

Small sample size Lack of control group No randomization Low generalizability Not representative to the population Lack of long-term follow-up Possible confounding variables

Depression BDI-II

Statistically significant difference between yoga + testimony and the control group (p = 0.009), and yoga and the control group (p = 0.031) but not between testimony and yoga (p = 0.111).

Small sample size Difference in length of interventions Self-selected sample Difference in pretest scores; mild BDI-scores in the Yoga-group and severe scores in the other groups. This may lead to regression to the mean. No blinding

1. Y: 1,5 h 2. T: 1,5 h 3. Y+T: 3 h 4. C: 4 days

continued on following page

217

Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Table 1. Continued Study

Jindani et al., 2015 RCT

Johnston et al., 2015

Mitchell et al., 2014 RCT

Price et al., 2017

Sample

N = 80 Female (71) Male (9) Age: 18-64; 41 (mean) PTSD Homogenous trauma experiences

N = 12 Military service personnel PTSD-diagnosis Age: 36-63 White (50%) Black (50%) Female (9%) Male (92%)

Treatment groups

1. Kundalini yoga 2. Control: Wait-list

Kripalu Hatha Yoga

Intervention

Kundalini Yoga (KY) includes classic yoga postures and physical exercises, breathing techniques, meditation, cultivation of mind-body awareness, and deep relaxation. Guidelines for traumasensitive yoga were implemented

Kripalu Hatha Yoga includes breathing techniques, physical postures, and aspects of mindfulness.

N = 38 Female military veterans and civilian Complex trauma 1. Kripalu Hatha Yoga (multiple incidents) 2. Control: Assessment PTSD (70, 7%) only Threshold- PTSD (22%) Comorbidity, especially depression (34,2%)

Kripalu Hatha Yoga includes breathing techniques, physical postures, and aspects of mindfulness. Guidelines for trauma-sensitive yoga was implemented. The control group was offered 12 yoga sessions after 1-month follow-up assessment.

N=9

Hatha yoga includes different physical postures, breathing techniques, and mindfulness meditation. Trauma-informed yoga class.

Hatha yoga

Duration

Outcome measures

1. Primary outcome: 2. PTSD 3. PCL-17 4. Secondary outcomes: 5. Resilience 6. Resilience Scale 7. Positive and negative affect 90 min weekly for 8 8. PANAS succeeding weeks 9. Mindfulness 15 min optional 10. FFMQ home practice 11. Insomnia 12. ISI (Insomnia Severity Index) 13. Depression/Anxiety 14. DASS-21 15. Stress 16. PSS 17. 18.

20 sessions 2 times a week for 90 min each 15 min home practice (recommended)

Findings

Limitations

Small to moderate between-groups effect sizes (0.09 - 0.25). Statistically significant differences between the groups with the yoga group showing greater improvements in scores of PTSD, perceived stress, insomnia, positive and negative effect, stress and anxiety. Both groups improved statistical significantly from pre- to posttest. .

Medium/small sample size 30% drop-out rate in the yoga group (0% in the control group) No active control conditions No blinding Lack of measures of long-term effects Low generalizability beyond Caucasian women

Effect sizes (Cohen’s d): 0.768 (CAPS); 0.290 1. Primary outcome: (RS); 0.429 (FFMQ) 2. PTSD Benchmarking:      CAPS Statistical significance 3. Secondary outcomes: lower CAPS-scores      Resilience for the yoga group      Resilience Scale compared to the (RS) treatment-benchmark      Mindfulness (d = -0.306), and      FFMQ higher than the wait-list control-benchmark (d = 0.612)

Small sample size No control conditions Benchmarking not based on reliable meta-analysis Medication or psychological treatment not an exclusion criterion High comorbidity rate Lack of reliability and validity measures of the RC and FFMQ

1. 12 sessions à 75 min      12 assessment sessions once a week in groups of 4-5

1. Primary outcome:      PTSD      PCL-C 2. Secondary outcome: 3. Anxiety      STAI      Depression      CES-D

Small effect sizes for between group effects (Cohen’s d): 0.20 (PCL); 0.15 (CES-D); 0.12 (STAI-S), 0.10; (STAI-T) No statistically significant difference between the groups on either outcome variable Both the Yoga and the control group had statistically significant improvements on PCLscores and marginally on CES-D scores, but only the control group on STAI-Trait scores

Small sample size Low power High drop-out rate (N=12) Lack of blinding No long-term effects Low generalizability Lack of active control condition High comorbidity rate

     1 h session per week for 20 succeeding weeks + home practice

     Primary outcomes:      PTSD      CAPS      Davidson trauma scale (DTS)      Dissociation      DES

Effect sizes (Cohen’s d): - 3.02 (CAPS) - 1.23 (DTS) - 1.41 (DES) Clinically significant reduction in PTSD and dissociative symptoms between pre- and post-test

Small sample size Low power Lack of control group Lack of randomization Low generalizability Possible confounding variables

continued on following page

218

Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Table 1. Continued Study

Reddy et al., 2014 RCT

Reinhardt et al, 2018 RCT

Rhodes et al. 2016 RCT Follow-up van der Kolk et al., 2014

Seppälä et al., 2014 RCT

Sample

Treatment groups

Intervention

Duration

Primary outcome: PTSD PCL-M PCL-C CAPS IES

Significant decreases in both active and control group, with no between-group differences.

Medium/small sample size No active control conditions No blinding Intervention confused with different techniques other than yoga

1. Hatha yoga includes different physical postures, breathing techniques, and mindfulness meditation. Trauma-informed yoga class. 1 h session per 2. Women’s health week for 10 education is an succeeding weeks interactive teaching style which seeks to increase knowledge about different health areas and improve the women’s self-efficacy.

     1. Primary outcome:      2. PTSD      3. CAPS      4. Dissociation      5. DES      6. Secondary outcome:      7. Depression      8. BDI      9. Regulatory capacity      10. IASC      11. Effects of cumulative trauma exposure      12. SLESQ

No statistically significant difference between the baseline study and the follow-up study, except for the predictor of frequency of post-study yoga practice. Further yoga practice improved symptoms of PTSD and depression

Small sample size Lack of a passive control condition (e.g. assessment only) Self-sample sampling Comorbidity as a confounding variable Difficult to isolate effects to the randomized control condition Low generalizability beyond USA and highly educated women with chronic PTSD

Sudarshan Kriya Yoga consist of 4 different meditative breathing techniques. Group discussion and simple stretching followed each SKY session

Effect sizes (Cohen’s d): d = 1.00 (PCL-M); 0.99 (MASQ); 1.30 and 1.07 (anxiety); 0.43 and 0.87 (depression); 1. Primary outcome: 0.51 (startle light);      PTSD 0.63 (startle dark); 1.22      PCL-M (military (respiration) version) Statistically significant 2. Secondary outcomes: difference between      Depression & the groups at 1 year Anxiety follow up with the      MASQ yoga group showing      Startle-response more improvements      BIOPAC electrode in PCL-M-scores (p      Respiration = 0.001) MASQ     BIOPAC respiration scores (p = 0.008). belt Statistical significance correlation between PTSD-symptoms and respiration-scores (p = 0.011)

N = 51 US Military personnel and veterans

Kripalu Hatha Yoga includes breathing techniques, physical 90 min twice postures, and aspects of      1. Kripalu Hatha yoga a week for 10 mindfulness. Trauma     2. Assessment only sensitive yoga guidelines. succeeding weeks Elements of DBT and mindfulness incorporated in the intervention.

N=49 (60 in the original study) Women with Chronic, treatment resistant PTSD Age (mean): 42.8 White: 71.4% Single: 46.9% College graduates: 71.4% Full-time employment: 38.8%

     1. Sudarshan Kriya Yoga (SKY)      2. Control: Wait-list

Limitations

1. Primary outcomes:      Alcohol abuse AUDIT      Drug abuse      DUDIT 2. Secondary outcome: 3. PTSD perception and management      Researchers own developed questionnaire

Kripalu Hatha Yoga includes breathing techniques, physical postures, and aspects of mindfulness. Traumasensitive yoga guidelines. 12 sessions à Elements of DBT and 75 min mindfulness incorporated in the intervention. The control group was offered yoga after last assessment.

1. Kripalu Hatha Yoga 2. Control: Assessment only

N = 21 U.S. military male veterans PTSD

Findings

Small sample size Low power No active control No statistically significant difference conditions between groups on Low pretest scores (noneither primary outcome diagnostic) which may indicate limited space for variables, The yoga group improvement reported improvement Intervention confused in the PTSD perception with different techniques and management scores other than yoga compared to the control No reliability or validity group (p = 0.001) measures of the PTSDquestionnaire No blinding

N = 38 Veteran (N = 9) and civilian (N = 29) women PTSD-symptoms Age: 18-65

     1. Hatha yoga      2. Control: Women’s health education      3.           (Yoga post-study)

Outcome measures

21 h distributed on 7 succeeding day with 3 h per session

Small sample size Not controlled for comorbidity Lack of demographic information No active control conditions No blinding Group discussion as a possible confounding variable High relative drop-out rate (N = 5)

continued on following page

219

Yoga as a Healing Modality of Trauma-Related Symptoms and Disorders

Table 1. Continued Study

Sample

Treatment groups

Stoller et al., 2012 RCT

Telles et al., 2010

N = 22 Male survivors after the Bihar flood (directly influenced; loss of friends, relatives, and property) 1. Patanjali Yoga Settled in a refugee 2. Control: Wait-list camp Normal health Farmers, physical labor Less than 7-year education level Age (mean): 31.5

N = 66 Icelandic residents exposed to an earthquake Male (7%) Women (93%) Higher education (30%) Employed (83%) Other physical activities 2-3 times a week (35%) Thordardottir Age: 1. Hatha Yoga et al., 2014 20-67 2. Control: Wait-list Other therapy or tried yoga prior to the intervention (2/3) Earthquake as primary trauma (Yoga: 42%; control; 44%) Mild scores on depression and anxiety ; low scores on quality of life; moderate to high scores for PTSD

220

Duration

Includes postures and the use of props (e.g. straps and wood blocks) which enhances deep touch pressure and 1. Sensory-enhanced 21 sessions à 75 proprioceptive input Hatha Yoga min during a 3 to balance the nervous weeks period      Control: No treatment system for a more relaxed and steady state. Developed for individuals exposed to extreme stress (e.g. war).

N = 70 Deployed military personnel Not necessary PTSD, but a stressful life situation (e.g. war) Female: N=22 Male: N = 48 Age (mean): 32

N = 64 Women with chronic, treatment-resistant PTSD (CAPS scores) Van der Kolk Age: et al., 2014 18-58 RCT White (78.1%) College graduate (73.4%) Employed (59.4%)

Intervention

1. Hatha yoga 2. Control: Women’s health education

The Yoga intervention included loosening exercises, physical postures, breathing techniques, and relaxation.

Hatha yoga includes different physical postures, breathing techniques, and mindfulness meditation

Outcome measures

1. Primary outcome: 2. Combat stress 3. Anxiety: STAI 4. Sensory processing: AASP 5. Secondary outcome:      Quality of life      Quality of Life Survey (developed by the authors)

1. Primary outcome: 2. Emotional distress 3. Visual analog scale (fear, anxiety, disturbed 7 succeeding days sleep, and sadness) with 1 h per session 4. Secondary outcomes: The control group 5. Heart rate variability was offered (HRV) learning yoga after 6. Polygraph (4 last assessment channeled)      Breath rate      Polygraph (4 channeled)

1. Primary outcome:      Stress-related symptoms      PSS-10 2 times à week for      PDS 6 succeeding weeks      Depression with 60 min per      BDI-2 session      Anxiety      BAI      Quality of life      Icelandic Quality Scale

1. Hatha yoga includes different physical postures, breathing techniques, and mindfulness meditation. Trauma-informed yoga class. 1 h session per 2. Women’s health week for 10 education is an succeeding weeks interactive teaching style which seeks to increase knowledge about different health areas and improve the women’s self-efficacy.

Findings

Limitations

Medium sample size Skewed randomization with those high in Statistically significant sensory sensitivity difference between randomized to the control condition the groups on STAItrait (p < 0.001) and Lack of demographic STAI-state scores (p < data on what kind of war exposure and length of 0.001) with the Yogagroup improving more service compared to the control Lack of measures of long-term group. No statistically Quality of Life Survey significant difference not been tested for between the groups for validity or reliability Lack of active control sensory processing condition Only self-report for the physiological parameters Small sample size The use of VAS instead of standardized measurements Lack of measures of long-term effects No reported effect sizes Purposive sampling No statistically Reporting of t-test significant difference results (positive) instead between the groups on of the ANOVA in the either variable discussion of whether yoga has an effect over the control group (Bonferroni-correction may lead to false negatives)

No reporting of effect sizes No statistically significant difference between the groups Both groups showed statistically significant improvements on stress and PTSD related symptoms, and depression between pre- and posttest (p < 0.01)

Effect sizes: CAPS d = - 0.41 (group difference); d = 1.07 (yoga); d = 0.66 (control) DTS      1. Primary outcome: d = - 0.34 (group      2. PTSD difference); d = - 0.52           CAPS (yoga); d = - 0.29           Davidson (control) Trauma Scale BDI-II      3. Secondary No statistically outcomes: significant difference;           Depression d = -0.60 (yoga); d =           BDI-II 0.39 (control)           Affect regulation IoASC           Inventory of d = - 0.31 (group Altered Self-Capacities difference); d = - 0.44 (IoASC) (yoga); d = 0.03 (control) Statistically significant differences between the group with the yoga group improving more on CAPS, DTS and IoASC scores.

Medium sample size No randomization (allocation based on place of residence) No active control conditions Self-sample sampling Other trauma experiences as «noise» Lack of measure of longterm effects No blinding (the researcher both instructor and analyzer) Low response rate on the PDS-scale Few yoga sessions

Medium sample size Lack of a passive control condition (e.g. assessment only) Self-sample sampling Lack of measures of long-term effects Comorbidity as a confounding variable Low generalizability beyond USA and highly educated women with chronic PTSD

221

Chapter 13

Clinicians Applying Yoga Principles and Practices in Pain Care:

An Evidence-Informed Approach Shelly Prosko PhysioYoga, Canada Matthew J. Taylor MyRehab, LLC, USA

ABSTRACT This chapter provides a practical introduction and guidelines for clinicians and researchers to appreciate the relationship between two enormous topics: pain and yoga. The emphasis is on practicality for both audiences with frequent citation to recent related publications and their deeper citations. The intention to fuel the imaginations of both clinicians and researchers occurs with a foundational understanding of how yoga, pain, and pain care are related. The authors’ experience in the clinic using yoga for people with pain and future areas for clinical and research inquiry are included. A list of techniques for the clinic, their clinical rationale, and two case reports ground the material and invite additional reflection for the psychophysiological applications of yoga in pain care.

INTRODUCTION The current global pain crisis is in urgent need of an integrative pain care strategy (Foreman, 2014). While pharmacology will always be a part of care, might there be more cost effective, empowering and accessible technologies with fewer adverse side effects that can play an important role? The practices and principles of yoga and its psychophysiology are an important component of just such an integrative approach. DOI: 10.4018/978-1-7998-3254-6.ch013

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Clinicians Applying Yoga Principles and Practices in Pain Care

This chapter consolidates the perspectives and current research evidence from the recent white paper on yoga therapy and pain by the International Association of Yoga Therapists (IAYT) (Pearson et al., 2020), as well as the published textbook Yoga and Science in Pain Care (Pearson et al., 2019). Other chapters in this book provide thorough descriptions of the psychophysiology that is mentioned in this chapter. This chapter focuses on providing clinicians and researchers perspectives informed by published evidence and clinical mastery for current and possible future applications of yoga in pain care. Therefore, where citations are available, they will be offered. When assertions are made without citation, they are the clinical expertise of the authors. The objectives of this chapter are to: 1. 2. 3. 4.

Offer an introduction and guidelines for appreciating the relationship between pain and yoga. Describe the evidence base, emphasizing practicality and relevance for clinicians and researchers. Demonstrate applications from the authors’ clinical experience. Fuel the reader’s imagination from a foundational understanding of the relationship for future areas for clinical and research inquiry.

BACKGROUND Yoga, and its subset, yoga therapy, can be part of addressing the shortage of integrative solutions to the current public health pain crisis. The problems offered below are not so much a criticism of current care, but set the context and orientation to describe the opportunities available for yoga therapy to complement and augment pain care. An applicable list of the current limitations in pain care relevant to yoga and the pain crisis are (Pearson et al., 2020): 1. Root causes often not addressed: The primary problem of the pain crisis is often misrepresented as an “opioid/addiction/drug” issue. The complex, emergent phenomenon of pain defies simple, single issue categorization (Foreman, 2014). Rather, it should be asked, “What are the root causes of chronic pain and suffering, and how do we address and prevent them?” 2. Limited attention to prevention of persistent pain: When people in primary care present with pain, thorough and comprehensive management is required (Gewandter et al., 2015). Effective pain care includes prevention of persistent or chronic pain (ibid). 3. Lack of comprehensive and effective long-term management and self-management: There is a need for effective pain care/comprehensive management approaches for people in pain (Clauw et al., 2019). The bias in the U.S. is for single intervention solutions that tend to be high cost, of very limited efficacy, and relatively high risk while also failing to address social, behavioral and spiritual determinants, as well as the structural/cultural barriers to care. Most approaches are not patient-centered, nor do they provide simple, effective long-term management and self-management options (ibid). 4. Lack of options for vulnerable populations: A large portion of those in pain crisis are in vulnerable populations without adequate care and support for their diverse needs (Simon, 2012). 5. Lack of pain literacy: Weak or no pain literacy by stakeholders (consumers, providers, legislators, insurers, and governments) who have not or can not keep up with the rapid advancements in pain science (Louw et al., 2011 & 2019; Blickenstaff et al., 2016). 222

 Clinicians Applying Yoga Principles and Practices in Pain Care

6. Limited awareness of yoga in pain care: Very limited public awareness of the value that yoga and yoga therapy can provide in pain care. Consider the following brief description of the contemporary understanding of pain and pain science, and all of its complexity before considering yoga therapy’s role. Pain is now understood to be an emergent output phenomenon of ever increasing complexity (Foreman, 2014) that goes far beyond being within just the ‘skin or skull’ of the individual in pain to include social and systemic environmental factors. Long gone should be the simplistic mechanistic models of the last century (this has not happened, though) of pain being an input from injured tissue, but now understood to be an output (Butler & Moseley, 2003). Contemporary pain definitions continue to evolve. The International Association for the Study of Pain (IASP) is currently undergoing a comprehensive review of their pain definition, with leading pain researchers and thought leaders having difficulty coming to a consensus (Taylor, 2019). Williams and Craig (2016) describe pain as “...a distressing experience associated with actual or potential tissue damage, with sensory, emotional, cognitive and social components”. The international definitions will require regular adjustments far beyond where they are today as pain science continues to rapidly evolve. For example, Stilwell and Harman (2019) offer a broader evolving framework that will continue to change that they presently term the “5 E process” of pain: embodied; embedded; enacted; emotive; and extended. The authors fully describe the 5 E’s of the process, the historical development of previous frameworks, as well the limitations of their description citing ongoing discoveries and interdisciplinary breakthroughs that make change inevitable. As is apparent by just those five words in the 5 E process of pain, the depth and complexity of defining and describing pain is beyond the scope of this chapter and the reader is referred to that article for further description. However, what is germane to this chapter is that any solution to the pain crisis must move beyond the many myopic care models currently available, to a care model that addresses environmental, systemic, and structural influences on the pain crisis. Yoga in its breadth and depth as a lifestyle health science has long considered this wider net of what leads to suffering. That relationship of yoga to pain and suffering is the next consideration. Yoga therapy is a subset of yoga in general (Bhavanani et al., 2019). The contemporary practice of yoga generally focuses on the postures as exercise. However, yoga’s rich history and tradition is that of a comprehensive system intended to alleviate suffering (Feuerstein, 1998). Yoga practices include more than postures, such as breathing practices, non-reactivity training, concentration, ethical practices, and meditation. As a wisdom tradition, yoga combines philosophical as well experiential exploration around pain and suffering (Taylor, 2018). Yoga teachings and practices are not static, and today continue to evolve. The current biopsychosocial-spiritual (BPSS) understandings and treatment approaches that apply to client populations with persistent pain tend to mirror the multi-dimensional approaches of yoga to pain and suffering. This relationship of practices of yoga is explained further after describing yoga therapy’s relationship with comprehensive integrative pain management. There is also a general lack of awareness and understanding of how yoga therapy is differentiated from the practices most people associate with contemporary yoga (such as general group asana-focused yoga classes) and of its benefits to health and well-being including pain care (Bhavanani et al., 2019). Yoga therapy is an evolving profession with a growing evidence base to include nearly 500 studies in 2018 (Gothe et al., 2019). The IAYT has a scope of practice, a PubMed-indexed research journal, accredits schools of yoga therapy, and certifies individual therapists and continuing education providers. Globally there are other 223

 Clinicians Applying Yoga Principles and Practices in Pain Care

groups in various levels of development of the profession. Raising awareness for stakeholders and consumers is a shared responsibility among these groups as awareness will be critical for yoga therapy to assume its role in integrative pain management as a collaborative partner that offers safe and effective options for people in pain. This chapter is one more step in that larger process. The antidote to the above pain problem list as noted requires a consensus of best pain care practices that will continue to develop and change rapidly. In the United States, one such recent consensus is Integrative Pain Care Policy Congress’s Comprehensive Integrative Pain Management (CIPM). This collective of united representatives from over 65 organizations and agencies - including licensed and certified health care professionals, public and private payers, people living with pain, federal agencies, purchasers of healthcare, researchers, policymakers, and policy experts. The IAYT participated as a member of this group. At the 2017 inaugural meeting, they supported this consensus definition of CIPM taking into consideration many international definitions of pain (World Health Organization, IASP, etc) and contemporary pain care evidence to arrive at: Comprehensive, integrative pain management includes biomedical, psychosocial, complementary health, and spiritual care. It is person-centered and focuses on maximizing function and wellness. Care plans are developed through a shared decision-making model that reflects the available evidence regarding optimal clinical practice and the person’s goals and values (IHPC 2017). The Congress believes, by this definition, that no one group can properly advance this integrative care alone (Clauw et al., 2019). “Comprehensive integrative” is also a level of consciousness, requiring changes from the simplistic, linear thinking that contributes to and sustains the pain crisis to adopting more complex thought and integrative systems-based awareness as Taylor (2019) describes thoroughly. This chapter is based on that framework as it reflects the current best practices around pain care. It is of particular interest that the CIPM definition also mirrors the aspects of a definition of yoga therapy. Pearson et al. (2020) summarize the commonalities in Table 1 [bracketing/underline/italics/ numbering ours to highlight crossover of CIPM and yoga therapy]: Table 1. CIPM and Yoga Therapy Definition Comparison Comprehensive, integrative pain management includes 1) biomedical, 2) psychosocial, 3) complementary health, and 4) spiritual care. It is 5) person-centered + focuses on maximizing function and wellness. Care plans are developed through a shared decision-making model that reflects the available evidence regarding optimal clinical practice and the person’s goals and values.

Yoga therapy [adjunctive & complementary health practice] is the process of empowering individuals [person-centered] to progress [No silver bullet, requires a care plan] toward improving health and well-being [focused on max function and wellness, not pathologies] through the application of the philosophy and practice [biomedical, psychosocial, spiritual, person-valued] of yoga [5000 yr old psychospiritual tradition] (Bhavanani et al., 2019)

Source: (Adapted from Pearson, Prosko, Sullivan & Taylor, 2020).

This understanding of yoga for pain care as being person-centered, focused on wellness, and utilizing a shared-decision making process grounds this chapter. An important note for clinicians and researchers to realize is that currently much of what has been published (primarily in modern yoga) is based on the current dominant mechanistic cultural model, rather than this more in-depth definition and aspiration of

224

 Clinicians Applying Yoga Principles and Practices in Pain Care

a broader understanding of yoga. See Pearson et al. (2019) for a thorough discussion on this topic, and know that most research has been symptom-focused by methodological and funding demands (Eaves et al., 2019). Further, often what is sold and taught as yoga therapy has been what Bhavanani (2011) describes as ‘yogopathy’. Yogopathy tends to concentrate on treating the symptoms of a diagnosis. In contrast, yoga therapy tends to focus on wellness and salutogenesis while finding and uprooting the cause(s) of pain and suffering. Yoga therapy and yogopathy both serve a purpose, but it is a critical distinction to identify which is being considered in our search for how to best support people living in pain both in the clinic and research setting (Bhavanani et al., 2019). With that distinction in mind, yoga therapy is not a simplistic symptom modifier, rather it utilizes a biopsychosocial-spiritual model that both emphasizes and influences each domain of health. The following is an excerpt from Pearson et al. (2020) that describes these domains and the relationship to yoga: 1. Mental and physical health are cultivated through the synchronistic application of yoga practices. Top-down and bottom-up processes are both employed through the practices of movement, breathing techniques, meditation, awareness practices, and ethical inquiry to promote systemic mind-body regulation and resilience. These practices have been shown to help to promote autonomic nervous system regulation, strengthen interoceptive skills, foster positive psychological states, increase health and resilience of the body, and enhance prosocial-behavioral attributes as well. 2. The spiritual wisdom tradition of yoga reflects commonly held features of spiritual care that holds well-being for pain care includes connectedness (personal, interpersonal, existential); purpose, such as dharma, where the person finds a greater sense of harmony with life; and via ethical inquiry through the yamas and niyamas. These support healthy and adaptive relationships with oneself (body, emotions, thoughts, beliefs), others (cultivating better quality relationships), and with life in general (fostering meaning-making and purpose). This spiritual focus of yoga is reflected in the literature where it has been demonstrated that yoga practitioners both shift to spiritual intentions with prolonged practice and that shifting to a spiritual focus can lead to higher levels of well-being compared to yoga students who hold solely on physical intentions. 3. Social well-being (the broader complex context of pain) arises from yoga practices through the cultivation of qualities such as patience, acceptance, compassion toward self and other, as well as the realization of the connection between all people as part of yoga philosophy. Yoga’s use of group classes can also facilitate more meaningful relationships with the community to enhance social integration and well-being. Yoga therapy will not solve the current complex pain crisis, but hopefully this brief summary opens new appreciation for the many ways it can contribute. Asking, “What causes the suffering in this complex problem?” is the discerning feature of yoga therapy and its unfolding process of discovery. How yoga therapy can make a meaningful contribution to solving the overlapping public health crises of chronic pain, opioid misuse and overdose, and mental health disorders is addressed in the next section. There are practical, feasible, and accessible opportunities that yoga can offer for clinicians and researchers.

225

 Clinicians Applying Yoga Principles and Practices in Pain Care

Solutions and Recommendations How Yoga Can Support Pain Care The last section hinted at how yoga therapy may support the six problems initially listed in the pain crisis. Here is a direct outline of how yoga therapy may help address these current problems in pain care, realizing there are also many more problems to address but these are most germane to this focus. The current crisis is not only taking a toll on those in pain, their families and support systems, but also on those providing pain care in the present systems. Table 2 lists the problem and a postulation based on clinical expertise, research evidence, and consumer needs and values. The “consumers’ needs” as part of evidenced based care and person-centered care need to include the providers’ needs, as well as those in pain. Table 2. Pain crisis problems and rationale for yoga therapy Pain Crisis Problem

Rationale for Yoga Therapy

Root causes often not addressed: The root of the crisis is pain, not the symptoms of addiction, opioid abuse, suicide and overdose.

Focuses on the large and small sources of suffering and pain, including the individual, but also wider system sources.

Limited attention to prevention of persistent pain: Lack of focus on wellness and prevention first, and lack of effective early pain care when needed.

Focuses on wellness, health creation and lifestyle behavioral changes grounded in an updated/edited personal narrative of meaning, purpose and connection.

Lack of comprehensive and effective long-term management and self-management: Still searching for effective pain management approaches knowing single intervention approaches have failed. Focused on “fixing” and managing pain vs caring for the individual with pain. The person in pain plays a more passive role.

An ongoing, iterative process of uncovering related and central sources of suffering, with patient-centered and valued care that includes individualized practices and modifications of long-term, empowered self management. The provider facilitates processes of change within the person in pain. Evidence of effectiveness of yoga for pain care is promising and growing (Moonaz, 2019).

Lack of options for vulnerable populations: Current high tech/low-value care systems leave large numbers of vulnerable populations without access or local care.

Provides both individualized and group access with low tech/highvalue person-centered care that can be delivered in remote and underserved environments.

Lack of pain literacy: Widespread pain science illiteracy in many stakeholders, including the majority of providers. (Louw et al., 2011 & 2019; Blickenstaff et al., 2016).

Ongoing practices and philosophy teach “eco” literacy concepts and experiences, that translate naturally to principles of evolving pain science such as the 5 E process describes (Stilwell & Harman, 2019).

Limited awareness of yoga in pain care: Cultural and research limitations in awareness of what yoga and yoga therapy are and how they relate to pain care.

Both top-down public education from organizations and providers, as well as growing consumer demand based on experience and needs.

It is worth noting that this shift of focus and these above rationales together with the health provider’s personal practice of yoga can also help address providers’ symptoms of burnout, promote caregiver health, well being, resilience and career longevity and enhance compassionate pain care (Prosko, 2019a). What is offered next is the evidence and practices that are part of this shift. An unheralded benefit of this shift is that when one does the practices, and even as they teach the practices, they can’t help but reflect within themselves (reflexivity as part of the clinical mastery of evidence-based medicine), and also experience their own shift in physiological and emotional states as well as insights. First, what is the research evidence that informs our clinical experience and serves as a basis for further research inquiry?

226

 Clinicians Applying Yoga Principles and Practices in Pain Care

Research Evidence In the past decade, the supporting research evidence has grown exponentially and continues to expand (Gothe et al., 2019). This extensive list of the evidence for yoga and pain and numerous pain associated conditions and their co-morbidities, exceeds the limits of this chapter. The reader is pointed toward the other chapters in this book, as well as the extended appendix in the open access white paper by Pearson, Prosko, Sullivan & Taylor (2020). That appendix lists by condition/diagnosis within each of the five facets of the CIPM definition with over 80 citations. That foundation of evidence informs the following use of the assessments and technologies below.

Evidenced-based Practices for the Clinic There are simple, practical ways of introducing specific yoga assessments, techniques, principles, and philosophy into pain care. This section offers a summary of some of the most common practices and their clinical rationale. The next section of case reports offers examples in context, and the final section describes future considerations for clinicians and researchers. Table 3 lists a range of applications for people in pain in the clinic setting. It is loosely structured in order of complexity, starting with more simplistic and less involved practices to more complex, involved and potentially more challenging practices. The practices are in line with contemporary pain science rationale. Given the relatively nascent incorporation of yoga into pain care, the limitations in methodological rigor and the current gaps in pain research due to the complexity of pain, there is little or no data to support specific techniques, doses, dosages, etc. The techniques in Table 3 are drawn from over 20 years of clinical expertise of each author (clinical expertise being one of the three foundations of evidence-based medicine) (Sackett et al., 1996) and research evidence (another of the three foundations) in varying degrees. Keep in mind this is not an exhaustive list of techniques, rather it is meant to offer some of the more common examples used in yoga therapy pain care. Greater depth of description is available in Pearson, Prosko & Sullivan (2019). Clinicians reviewing this table can see instances of a yogopathic or symptom-modifying line of reasoning for application (i.e. certain pranayama techniques such as alternate nostril breathing have good evidence for changing the autonomic state). There will be other instances, when viewed from outside a yoga therapy paradigm of inquiry, where for yoga novice clinicians it may be difficult, if not impossible, to comprehend such concepts as: how do mudras or chanting support pain management when there’s essentially no evidence to support these claims, only conjecture? In such instances, a good first step for the clinician or researcher would be to seek out first-hand experiential learning from a qualified instructor. Then with that first-person context, initiate study in the manner in which those interventions are assessed from a yoga therapy paradigm, as well as reference back to this textbook and Pearson, Prosko & Sullivan (2019) to bridge the pyschophysiological correlates. Keep in mind that these practices and potential outcomes are context dependent. The individual’s sense of safety (and therefore pain experience) may be influenced by their beliefs and expectations about the technique, current physiological state or mood, the surrounding environment (light, sound, temperature), the position they or the therapist is in, quality of the therapeutic alliance, therapist’s language and tone of voice, and even the therapist’s current physiological state, beliefs and expectations about the practices offered.

227

 Clinicians Applying Yoga Principles and Practices in Pain Care

Table 3. Yoga techniques and clinical rationale for pain management Technique

Clinical Rationale for Pain Care

Breath awareness/interoceptive attention (Gard et al., 2014)

Sustaining attention to sensory and movement experiences associated with breathing to reinforce top-down ability and bottom-up sensory remapping for enhanced motor planning. Can also identify distorted sensory mapping, habitual holding/bracing patterns, and altered breath patterns that may potentially be perpetuating the pain experience.

Directed breathing/imagery (Wilson-Mendenhall et al., 2019; Villemure et al., 2013)

Actively directing the breath along with both sensory and motor imagery. Can include facilitating motor recruitment of related somatic components such as intercostals, diaphragm, pelvic floor etc.

Intention setting (Vlaeyen & Crombez, 2020)

Focusing on an intention can offer direction and support attention and goals, influencing top-down processing and pain modulation.

Meditation/dhyana (Hilton et al., 2017; Zeidan & Vago., 2016)

Multiple forms with strong evidence as a valuable role in areas important for pain self-management: self-regulation (cognitive and emotional), self-efficacy, acceptance, reduced catastrophization. From simple attention to formal sitting/supported practices that don’t require movement, meditation may be a helpful starting point for those with kinesiophobia and an accessible practice during flare-ups.

Foundational postures: Seated and/or standing mountain (Moseley & Butler, 2017)

Exploring novel foundational posture options promoting stability without rigidity. Optimize dynamic postural strategies with the integration of the pelvic floor, respiratory diaphragm and glottis/thoracic outlet. Identify and reduce excessive muscular effort, using imagery of a mountain support from the base vs. military protective bracing/guarding. Use of language in line with pain science and informed by yamas and niyamas to promote sense of safety, support and trust vs fear, fragility and brokenness. Breath cues can be used to enhance dynamic postural stability without rigidity.

Visualization/motor imagery (Breckenridge et al., 2019; Yap & Lim., 2019)

Especially helpful to use in place of movements that may elicit pain or people with kinesiophobia. Numerous motor imagery techniques can be used such as visualizing movement prior to actual movement; visualize the movement without pain and/or with specific positive qualities of individual’s choice (smooth, easeful, peaceful movement); move opposite non-painful body part, then just visualize moving painful body part with those same qualities. Motor imagery allows for pre-motor cortex firing while providing novel input into the danger detection system without system sensing danger message thereby reducing threat resulting in potential pain modulation. Can then add the gentle movement after visualizing it first.

Yoga nidra/guided imagery (Lazaridou et al., 2019)

Training non-reactivity to varying stimuli, to include addressing fear-avoidance behaviours, trauma triggers, and developing a witness/observer role to moment to moment sensory experiences, as well as inducing/promoting deep relaxation response.

Pranayama: variety of breathing exercises (Jafari et al., 2017; Schmalzl et al., 2018; Telles et al., 2013; Saoji et al., 2017; Prosko, 2019b)

Multiple practices that regulate/manipulate the breath (rate, depth, length, varying motor recruitment patterns, etc) to influence pain via autonomic and central nervous systems for self-regulation, developing top-down resilience to adverse states and fostering relaxation response and bottom-up reinforcement. Focused attention/concentration via breathing facilitates discovering patterns of ideation related to breath, pain and habit.

Emotional body scan (McClintock et al., 2019; Strigo & Craig 2016; Craig, 2003)

Guided imagery exercise to explore somatic/sensory responses to varying emotional states, how they first present, their fluidity, and to gain top-down control in order to respond vs react habitually to the circumstances as in the past. Contributes to enhancing self-regulation, reducing catastrophization and rumination.

Identifying postural holding patterns in asana and daily postures (Moseley & Butler, 2017)

In both classical asana and the individual’s activities of daily living postures, explore unhelpful or inefficient patterns where they expend wasted effort, clench, guard or brace, then explore movement options without the patterns, linking back to insights from emotional body scan and foundational postures as above.

Mindful movement/supported postures (Carson et al., 2016)

Explore more dynamic movement habits, both classical and functional, as well as in restorative/ sustained supported postures, slowing down and attending inward, then noting insights in order to release old motor patterns, report related emotional states, or ask for support (both literally and metaphorically). Integrate more challenging movement to restore recovery of function, in line with yoga philosophy and current views in pain science.

Yama and niyama instruction/integration (Sullivan et al., 2018)

Share in modern terminology as they relate to pain science and use to inform the assessment process, treatment planning, and the language used by the provider. Can be used to explore personal ethical framework, historical antecedents to unexamined habits, reconceptualize pain, and search out new alternatives in line with personal preferences and beliefs. Language informed by yamas/niyamas may modulate pain by enhancing sense of safety, trust and therapeutic alliance between provider and individual in pain.

Vocalization/chant/mantra (Sullivan et al., 2018)

Enhance sensory and motor control of the glottis and other speech functions. Discover sensory and motor smudges in sound production, use improved awareness to track habits in daily life, and practice articulating their voice both literally and metaphorically in culturally appropriate drills and daily function as that relates to pain science.

Mudras (Taylor, 2013)

Simple and often easily accessible practices to complement other practices or use in isolation (individually or in series such as bija series, etc). Integrates a somatic experience often without flare-up. Also enhances awareness around patterns of holding from the large hand cortical maps and provides insight into how those habits influence breathing, posture, ideation, and emotional states.

Dharana/attention training (Gard et al., 2014; Schmalzl et al., 2015)

Top-down regulatory control in the current attention directing environments of media, technology, and internal narratives.

228

 Clinicians Applying Yoga Principles and Practices in Pain Care

How clinicians actually apply these tools in the clinic is the subject of extensive training programs including clinical mentorships. The following case reports are presented to glimpse such an application from two broad scenarios: 1) The first as a stand alone yoga therapy service for a person in pain, as can be the case for those with limited resources or in remote locations; and, 2) As part of an integrative team approach.

Case Reports Case Report A: Yoga Therapy as a Stand-alone Service *Ben is a 64 year old male with longstanding persistent/chronic low back pain for approximately the past 40 years. He had previously heard from various media sources that yoga could help back pain. He happened to see an advertisement for yoga therapy in his community and he was curious if it might help him, so he decided to give it a try. Narrative review: Ben’s main complaint is his constant, persisting lower back pain, generalized across both sides but often varying in location and intensity. He can’t recall any mechanism of injury and states he’s had back pain for most of his adult life. He has had three back surgeries (two partial lumbar discectomies; L4-5 fusion during third and final surgery). The last surgery was over 10 years ago. He states the surgeries “didn’t really help” and that his pain continues to persist and his function has declined over the years. He states he is never pain-free and cannot remember being free from pain for over 20 years. He feels weak, particularly in his legs, with low energy throughout the day. Aggravating factors: 1. standing or walking for over five minutes 2. bed mobility 3. transfers: getting in/out of chair, bed or vehicle, up/down off floor (fear/anxiety of falling as legs sometimes “give out” when back spasms unexpectedly) 4. activities of daily living requiring forward bending or twisting 5. sleeping: wakes up with pain approximately three to four times each night 6. stress: pain exacerbated when he worries and feels overwhelmed about finances or unfinished projects around the house that he needs to complete for his family. At the time of assessment, Ben said that his pain controlled him. He did not have strategies for pain reduction other than his pain medication that was minimally, occasionally and only temporarily effective. He expressed a desire to “one day be able to wean off” his medication. He is currently under the care of a family physician. Past medical history: Past history of general anxiety. No other significant health concerns reported. Medications: tramadol-acetominophen, pregabalin, indomethacin. Past interventions: Over the years, Ben has tried numerous interventions such as physical therapy, chiropractic, acupuncture, surgery, medications, counseling, massage therapy, and was part of a pain rehabilitation return-to-work program on three occasions (after each back surgery). He states that his past physical therapy and psychology interventions were beneficial, but were temporary and overall did not change his pain nor his function.

229

 Clinicians Applying Yoga Principles and Practices in Pain Care

Occupational and psychosocial-spiritual status: Ben was forced to retire early from a job he enjoyed because of the limitations from his pain. He expresses feeling “useless” and “not productive” since being off work and that he also misses the camaraderie of his coworkers/friends. He realized while he was telling his story that he was feeling somewhat isolated from his friends and unable to participate in his usual recreational activities such as fishing, golfing and hunting. He also feels guilty for not being able to contribute to and perform more of the household chores and duties and feels ashamed that sexual intimacy with his wife has declined since he is afraid of back spasms and pain. He is not affiliated with any religion or spiritual tradition, but when asked “what brings you joy and makes you feel alive, connected and peaceful?” he said he felt most connected and happy when he spends time in nature and with his new granddaughter or when he hears a good song. He has a loving and supportive wife, with two adult children that have recently moved away from home. He reports his memory has declined and his brain has become ‘foggy’ over the past few years. He is a non-smoker and rarely drinks alcohol. He typically eats three healthy meals per day, but says his pain has reduced his appetite and joy for food over the years. Physical observations: 1. gait and general mobility/transfers performed with excessive abdominal bracing, guarding, general muscle tension, and protective patterns with reduced breath and body awareness and proprioception 2. reduced mobility throughout entire spine and pelvis when performing all movements 3. shallow and rapid breath pattern at 18 breaths/minute with reduced abdominal movement during quiet breathing and reduced costal expansion during more engaged breathing, with excessive use of accessory muscles of respiration on inspiration 4. frequent breath holding and jaw clenching elicited when performing most movements 5. poor balance reaction strategies due to overall rigidity and reduced mobility of trunk, pelvis and ankles 6. finds comfort in: prone lying with folded blanket under hips and shins; supine lying with bolster for support under thighs and folded blanket under head; and supported child’s pose with bolster for supporting trunk and folded blanket roll at ankles. Ben’s Goals: 1. participate in a regular home yoga practice independently and use as part of a pain reduction and coping strategy 2. improve overall body strength and tolerance to walking and standing (at least 30 min duration) 3. improve function, including: able to bend over to put on shoes/socks; get in and out of bed/vehicle safely with ease and without fear/anxiety; fish, hunt, golf; engage in sexual intimacy with wife with ease and without fear/anxiety of pain; perform household chores/duties like light yard work, putting up a shelf and organizing the garage. Yoga therapy program: Ben attended 14, 1.5 hour yoga therapy sessions over 13 months. The frequencies of sessions were initially at once per week, then reduced to once every 8-10 weeks as needed. The program consisted of a variety of yoga techniques (as described below), pain education, self-reflection, developing skill in daily strategic planning, and ongoing inquiry/assessment, all of which were informed

230

 Clinicians Applying Yoga Principles and Practices in Pain Care

by and in line with yoga teachings and current pain science. The following are some examples of what was included over the 13 month program: Breath and body interoceptive awareness practices were performed in Ben’s most comfortable positions with adequate props for optimal support to ensure he felt safe in order to minimize threat to his systems. Breath awareness practices included having him simply observe various qualities of the breath including how it moves his body. Guided body scans, visualizations, yoga nidra, and various body tension release techniques that heightened interoceptive awareness were also used. Pranayama such as alternate nostril breathing, three-part breath and extending the exhalation along with visualizations, proved to be valuable for reducing his pain, muscle tension and guarding (especially surrounding the low back and trunk), and promoting generalized calming effects. Foundational postures such as mountain were explored along with various postures (modified as needed) that instilled a sense of confidence, strength and grounding (warrior series, tree, triangle) along with concepts of trust, contentment, self-compassion, and courage to surrender and let go of extraneous effort or thoughts that brought more tension, resistance or pain to the experience. Mindful movement practices that included or simulated functional movements of Ben’s most meaningful activities of daily living that he wanted to address were also included. Cueing, language and use of guidelines that were in line with yoga teachings and pain science (Pearson, 2016) were key to guiding the movements to include more ease, grace, mobility, and coordination with more confidence and to help reduce Ben’s fear of movement, body tension, bracing, guarding and protective habits, ultimately resulting in less pain. Sessions always started with a centering breath and body awareness meditation and ended with supported savasana (the corpse pose). The program also included pain education and resources, developing strategies to include activities that brought him joy (time in nature, music), social connection (he contacted a friend to start joining him on short nature walks and he joined a community yoga class) and gratitude journaling. 14 sessions and 13 months later: 1. Ben states “the pain doesn’t control me anymore” and he feels more like himself 2. improvements in scores in three outcome questionnaires: Fremantle Back Awareness Questionnaire, Depression Anxiety Stress Scale, and the Tampa Scale of Kinesiophobia (30 percent, 20 percent and 75 percent improvement respectively) 3. significantly improved mobility: reduced guarding, bracing, fear and pain; and more confidence and ease with general mobility and activities of daily living, with improved quality of movement through trunk and pelvis 4. more easeful and efficient breath pattern during movement, with calmer breath at rest at 12 breaths per minute 5. able to walk one hour per day on average, almost daily, in nature and often with a friend 6. engaging in regular sexual activities, but still reports his stamina is low 7. improved sleep: averaging six hours of uninterrupted sleep 8. states he now has better control of his pain, using a combination of practices, reframing his thoughts, and also gaining a deeper understanding into his capabilities and is better at setting boundaries. At times when he is stressed, he cannot control or reduce it as much as he desires. 9. his back still feels sore in the morning, but improves after morning yoga routine 10. able to return to nine holes of “easy” golf but hasn’t tried fishing or hunting 231

 Clinicians Applying Yoga Principles and Practices in Pain Care

11. 12. 13. 14.

able to perform light yard work and small chores around the home unable to perform heavy yard work or go for walks for more than one hour successfully starting to taper off pain medications under physician supervision independently and safely engages in daily home yoga practice and attends community yoga class once per week; states he thoroughly enjoys both and attributes his success in being able to control his pain to his new awareness, insights and yoga practice. 15. Ben substantially attained his set goals. Detailed assessment and progress notes were sent to the physician on a regular basis over the course of the program. He and his physician were pleased with the outcomes. Ben states, “Yoga helps me practice letting go and move more freely and I can feel that immediately helps my pain. I never realized how tense I used to be all over. Now I am more aware and know when I’m tense, and I can change it and then my pain changes. It’s amazing how much more in tune I am with my body.” He states that “yoga has taught me how I can help myself, and not just rely on others to fix my pain.” *Name has been changed in order to protect confidentiality

Case Report B: Yoga Therapy as Part of an Integrative Team The team consisted of a pain interventionist, psychologist, physical therapist, chest surgeon/plastic surgeon, pastor, and yoga therapist. History: A 51 year-old single widow with past medical history of post-mastectomy pain syndrome with a persistent complaint of bilateral neck, shoulder, chest and arm pain, left worse than right for seven years presented. Surgical history includes left breast mastectomy for breast cancer seven years previous, followed by a failed reconstruction which necessitated a second reconstruction with TRAM flap procedure and right breast reduction to match left. She then underwent a C5-C6 discectomy three years later. Social emotional history includes husband dying of AIDS in mid-1980’s with her health being tracked by epidemiologists at the CDC for 10 years as he was one of the early AIDS victims. Two weeks after being released from observation her breast cancer was diagnosed. Pain experience: She was also undergoing a work-up for pain pumps to modulate her chronic pain experience. She described the pain in her neck and shoulders which is most notable upon static sitting or standing beyond 20 minutes, usually lasting all day. Intermittent increases in pain were noted throughout the day after prolonged phone and computer work. Pain was at a 7/10 on the left and 5/10 level on the right. She also reported radiating pain, numbness, and paresthesia proximally and distally around the left chest and upper abdomen. Past treatment experience: She had undergone more than four previous attempts at physical therapy for her upper quarter pain over the past five years with poor results and increased pain. She was referred by her pain management physician to begin a course of yoga therapy intervention designed to manage her chronic bilateral upper quarter complaints. Patient goals: “To decrease pain so that I can work full-time, to include increasing my static sitting tolerance beyond 20 minutes and improve my health through better tolerance of a regular exercise program.” Employment: Patient was working part time in a phone sales environment. In the past she was a voice instructor and assistant television producer for top soap operas in a major metropolitan area.

232

 Clinicians Applying Yoga Principles and Practices in Pain Care

Recreational activities: Prior to her surgeries, she engaged in theater, dance, hiking, and needle point. Now socially isolated in a small apartment with her pet dog. General health: Good Medications: oxycodone and acetaminophen and pregabalin. Treatment course: She attended eight yoga therapy sessions over the course of two months which included restorative yoga postures to open the anterior chest, sensory attention to her chest wall with her own hands to integrate chest wall/abdomen; directed breathing (guided imagery = prana vidya/bhavana) to numerous areas that were difficult to perceive, and slow progression from supine postures to seated practices to lengthen her spine and increase upper extremity mobility. All postures had a focus on core awareness/interoception and respective bases of support when supine, standing or sitting. Other yoga therapeutics included breath awareness exercises, pranayama to slow and expand her breath, open hand mudras, asana to sense pelvis and thoracic cage and visualization of re-establishing communication with core, breast reconstructions and thoracic cage. Integrative care: Physical therapy (PT) provided gentle soft tissue mobilization and intervertebral joint mobilization of the cervico-thoracic spine and soft tissue of the thoracic cage with active engagement of patient visualizing the anatomy after studying a spine beforehand. During the postures and PT she was directed to study her thoughts, sensations, and emotional responses and then journal the patterns and insights post-activity. These activities were shared with her psychologist who worked with her life narrative and discussed with her pastor. She was able to continue with the prescribed movement programs with excellent adherence as she didn’t experience the severe flares of earlier disembodied rehab attempts. Insights: She also incorporated the knowledge from ergonomics from PT but not in the formerly instructed rigid bracing, but soft “mountain” pose in her workspace and home, especially noting cued stressors and postural reactions as assigned by her yoga therapist. She was given self-care postures and self-reflection to be done at her desk, and options of self-management if she sensed moving into fear or panic to perform at home and to replace watching TV news in the evenings. She attended a weekly gentle yoga class for people living with pain and reported she’d struck up friendships with two other students noting how good it felt to have social connection again. Results: After eight sessions, improvement in trunk awareness and flexibility was enhanced in quantity and reduced fear, and an overall reduction in her cervical and shoulder symptoms was noted, however, the relief that she experienced was brief in duration and her symptoms still prohibited her from completing her daily work activities. A re-examination demonstrated an improvement in trunk ROM and flexibility by 40 percent overall. She reported improvement in cervical pain which was at a 3-5/10 level at work and resting symptoms at a 2/10 level. Due to her persistent symptoms and inability to work full time, arrangements were made to proceed with the pain pump. Quality of life changes: 1. Patient reported an enhanced sense of efficacy in using her breathing techniques and improved postural holding habit awareness to modulate her pain, rest in bed when unable to sleep and control her anxiety over the future. 2. Most notably, she had learned to use the feedback from her body to either change her behavior or request changes from others in circumstances where she would have tried to “push” through painful scenarios.

233

 Clinicians Applying Yoga Principles and Practices in Pain Care

3. She also noted that her introspection on thought patterns behind her postural holding had demonstrated to her how much she felt dis-spirited in her current vocation and she began actively seeking employment possibilities back in theater coaching and education. 4. She related that when she visualized that type of future it deepened her breath, extended her spine, and opened her chest while easing her pain. Her sense of well-being, direct spiritual support in the breath, new friends, and a renewed search for purpose all gave her hope for the future despite her pain.

FUTURE RESEARCH DIRECTIONS Future Inquiry into Yoga for Pain Care This section offers recommendations for clinicians and researchers with specific actionable steps rather than generally and simply suggesting the obvious, which is that “more randomized controlled trials” are needed in research and “ongoing yoga education and training” is needed clinically.

Clinicians Initiate or continue a personal yoga practice beyond asana. Such a practice can prime a clinician for enhanced resonance, compassion, empathy, and prevent burnout (Taylor, 2015; Prosko 2019). Explore practices with colleagues or in an inter-professional professional development setting. Read and practice from additional yoga and pain resources. Start small and observe your and the client’s responses, modifying to adapt to the circumstances. Don’t forget to pause and be in awe at how our exploration of psychophysiology reveals the complex wonder of life and calls each of us to humility.

Research Currently there are many knowledge gaps and research limitations surrounding yoga and pain with numerous research challenges facing integrative investigations. Further inquiry is needed by researchers in collaboration with clinicians, with these following points to consider: 1. Yoga doesn’t help everyone nor is it effective for all pain conditions. 2. There are many yoga practices, styles and techniques: Which are effective? In what context? For which individuals? 3. What skill level is necessary for yoga therapists to attain the same results as the experts used in research? 4. How can future studies control for both researchers’ and subjects’ expectations and strong bias towards yoga being effective and how will results vary from current research? 5. More replication studies are needed to impact conclusions of most systematic reviews and meta-analyses. 6. More standardization of outcome measurements is needed for comparison between studies. 7. With standardization of outcomes, further inquiry into dose and duration of effects is needed. 234

 Clinicians Applying Yoga Principles and Practices in Pain Care

8. Qualitative and mixed methods research including a robust exploration for methodologies that can examine the essence of the yoga therapy health creation paradigm rather than the disease model. How does this get reported within current publisher’s limitations? 9. Longitudinal studies to capture long-term effects are needed. The old adage that “Yoga is strong, but slow medicine” highlights a key void in the literature due to time constraints related to funding and completing studies that would capture such data. These and other issues therefore significantly limit what can be exaggerated claims about the positive effects of yoga on people living in persistent pain. For a comprehensive literature review, including knowledge gaps and research limitations, and a commentary on the current state of the evidence surrounding yoga and pain, the authors encourage the reader to review the chapter Current Research in Yoga and Pain by Moonaz (2019) in Yoga and Science in Pain Care by Pearson et al., 2019. Additional communication from both clinicians and researchers is welcomed. There are many unknowns at this point and so many possibilities, as well as new emerging research methods for inquiring into complex, integrative models of care that this chapter could not begin to allow description or discussion of them all.

CONCLUSION There is immense potential value of yoga as a viable, safe, practical, accessible, evidence-informed, and effective option for people in pain. The long standing practices and teachings of yoga, backed by its psychophysiological effects offer hope, inspiration and possibility for people in pain, their clinicians, and those looking to bring forward the next generation of research in pain. This chapter is merely a pause in the long quest to fulfill Patanjali’s Yoga Sutra Chapter 2, Sutra 16: Heyam-Dukham-Anagatam … “Pain that has not yet come is avoidable” (Carrera, 2006 p 120).

REFERENCES Bhavanani, A. B. (2011). Are we practicing yoga therapy or yogopathy? Yoga Therapy Today, 7, 26–28. Bhavanani, A. B., Sullivan, M., Taylor, M. J., & Wheeler, A. (2019). Shared foundations for practice: The language of yoga therapy. Yoga Therapy Today, 2019, 44–47. Blickenstaff, C., & Pearson, N. (2016). Reconciling movement and exercise with pain neuroscience education: A case for consistent education. Physiotherapy Theory and Practice, 32(5), 396–407. doi:1 0.1080/09593985.2016.1194653 PMID:27356079

235

 Clinicians Applying Yoga Principles and Practices in Pain Care

Breckenridge, J. D., Ginn, K. A., Wallwork, S. B., & McAuley, J. H. (2019). Do people with chronic musculoskeletal pain have impaired motor imagery? a meta-analytical systematic review of the left/right judgment task. The Journal of Pain, 20(2), 119–132. doi:10.1016/j.jpain.2018.07.004 PMID:30098404 Butler, D. S., & Moseley, G. L. (2003). Explain Pain. Noigroup Publications. Carrera, J. (2006). Inside the yoga sutras. Integral Yoga Publications. Carson, J. W., Carson, K. M., Jones, K. D., Lancaster, L., & Mist, S. D. (2016). Mindful yoga pilot study shows modulation of abnormal pain processing in fibromyalgia patients. International Journal of Yoga Therapy, 26(1), 93–100. doi:10.17761/1531-2054-26.1.93 PMID:27797669 Clauw, D. J., Essex, M. N., Pitman, V., & Jones, K. D. (2019). Reframing chronic pain as a disease, not a symptom: Rationale and implications for pain management. Postgraduate Medicine, 131(3), 185–198. doi:10.1080/00325481.2019.1574403 PMID:30700198 Craig, A. D. (2003). Interoception: The sense of the physiological condition of the body. Current Opinion in Neurobiology, 13, 500–505. doi:10.1016/S0959-4388(03)00090-4 PMID:12965300 Eaves, E. R., Hsu, C. W., DeBar, L. L., Livingston, C. J., Ocker, L. E., McDonald, S. J., Dillon-Sumner, L., & Ritenbaugh, C. (2019). Whole systems within whole systems: The oregon health plan’s expansion of services for back and neck pain. Journal of Alternative and Complementary Medicine (New York, N.Y.), 25(S1), S61–S68. doi:10.1089/acm.2018.0431 PMID:30870022 Feuerstein, G. (1998). The yoga tradition. Hohm Press. Foreman, J. (2014). The Global Pain Crisis. Oxford University Press. Gard, T., Noggle, J. J., Park, C. L., Vago, D. R., & Wilson, A. (2014). Potential self-regulatory mechanisms of yoga for psychological health. Frontiers in Human Neuroscience, 8, 770. doi:10.3389/ fnhum.2014.00770 PMID:25368562 Gewandter, J. S., Dworkin, R. H., Turk, D. C., Farrar, J. T., Fillingim, R. B., Gilron, I., ... Walcob, G. A. (2015). Research design considerations for chronic pain prevention clinical trials: IMMPACT recommendations. Pain, 156(7), 1184–1197. doi:10.1097/j.pain.0000000000000191 PMID:25887465 Gothe, N. P., Khan, I., Hayes, J., Erlenback, E., & Damoiseaux, J. S. (2019). Yoga effects on brain health: A systematic review of the current literature. Brain Plasticity, 5(1), 105–122. doi:10.3233/BPL-190084 PMID:31970064 Hilton, L., Hempel, S., Ewing, B. A., Apaydin, E., Xenakis, L., Newberry, S., Colaiaco, B., Maher, A. R., Shanman, R. M., Sorbero, M. E., & Maglione, M. A. (2017). Mindfulness meditation for chronic pain: Systematic review and meta-analysis. Annals of Behavioral Medicine, 51(2), 199–213. doi:10.100712160016-9844-2 PMID:27658913 Integrative Health Policy Consortium. (2017). National Strategy for Comprehensive Integrative Pain Management Gaining Momentum at 2nd Annual Integrative Pain Care Policy Congress. Retrieved from http://www.ihpc.org/national-strategy-for-comprehensive-integrative-pain-management-gainingmomentum-at-2nd-annual-integrative-pain-care-policy-congress/

236

 Clinicians Applying Yoga Principles and Practices in Pain Care

Jafari, H., Courtois, I., Van den Bergh, O., Vlaeyen, J. W. S., & Van Diest, I. (2017). Pain and respiration: A systematic review. Pain, 158(6), 995–1006. doi:10.1097/j.pain.0000000000000865 PMID:28240995 Lazaridou, A., Koulouris, A., Devine, J. K., Haack, M., Jamison, R. N., Edwards, R. R., & Schreiber, K. L. (2019). Impact of daily yoga-based exercise on pain, catastrophizing, and sleep amongst individuals with fibromyalgia. Journal of Pain Research, 12, 2915–2923. doi:10.2147/JPR.S210653 PMID:31802932 Louw, A., Diener, I., Butler, D. S., & Puentedura, E. J. (2011). The effect of neuroscience education on pain, disability, anxiety, and stress in chronic musculoskeletal pain. Archives of Physical Medicine and Rehabilitation, 92(12), 2041–2056. doi:10.1016/j.apmr.2011.07.198 PMID:22133255 Louw, A., Vogsland, R., Marth, L., Marshall, P., Cox, T., & Landers, M. (2019). Interdisciplinary pain neuroscience continuing education in the veterans affairs: Live training and live-stream with 1-year followup. The Clinical Journal of Pain, 35(11), 901–907. doi:10.1097/AJP.0000000000000756 PMID:31433321 McClintock, A. S., McCarrick, S. M., Garland, E. L., Zeidan, F., & Zgierska, A. E. (2019). Brief mindfulness-based interventions for acute and chronic pain: A systematic review. Journal of Alternative and Complementary Medicine (New York, N.Y.), 25(3), 265–278. doi:10.1089/acm.2018.0351 PMID:30523705 Moonaz, S. (2019). Current Research in Yoga and Pain. In N. Pearson, S. Prosko, & M. Sullivan (Eds.), Yoga and science in pain care: Treating the person in pain (pp. 37–51). Singing Dragon. Moseley, G. L., & Butler, D. S. (2017). Explain pain supercharged. Noigroup Publications. Pearson, N. (2016). Yoga Therapy. In D. L. Thompson & M. Brooks (Eds.), Integrative pain management. Handspring. Pearson, N., Prosko, S., & Sullivan, M. (Eds.). (2019). Yoga and science in pain care: Treating the person in pain. Singing Dragon. Pearson, N., Prosko, S., Sullivan, M., & Taylor, M. J. (2020). White Paper: Yoga Therapy and Pain: How Yoga Therapy Serves in Comprehensive Pain Management, and How It Can Do More. International Journal of Yoga Therapy, 10. Prosko, S. (2019a). Compassion in pain care. In N. Pearson, S. Prosko, & M. Sullivan (Eds.), Yoga and science in pain care: Treating the person in pain (pp. 235–256). Singing Dragon. Prosko, S. (2019b). Breathing and pranayama in pain care. In N. Pearson, S. Prosko, & M. Sullivan (Eds.), Yoga and science in pain care: Treating the person in pain (pp. 141–157). Singing Dragon. Sackett, D. L., Rosenberg, W. M., Gray, J. A., Haynes, R. B., & Richardson, W. S. (1996). Evidence based medicine: What it is and what it isn’t. British Medical Journal, 312(7023), 71–72. doi:10.1136/ bmj.312.7023.71 PMID:8555924 Saoji, A. A., Raghavendra, B. R., & Manjunath, N. K. (2017). Effects of yogic breath regulation: A narrative review of scientific evidence. Journal of Ayurveda and Integrative Medicine, 10(1), 50–58. doi:10.1016/j.jaim.2017.07.008 PMID:29395894

237

 Clinicians Applying Yoga Principles and Practices in Pain Care

Schmalzl, L., Powers, C., & Henje Blom, E. (2015). Neurophysiological and neurocognitive mechanisms underlying the effects of yoga-based practices: Towards a comprehensive theoretical framework. Frontiers in Human Neuroscience, 9, 235. doi:10.3389/fnhum.2015.00235 PMID:26005409 Schmalzl, L., Powers, C., Zanesco, A. P., Yetzb, N., Groessl, E. J., & Saron, C. D. (2018). The effect of movement-focused and breath-focused yoga practice on stress parameters and sustained attention: A randomized controlled pilot study. Consciousness and Cognition, 65, 109–125. doi:10.1016/j.concog.2018.07.012 PMID:30099318 Simon, L. S. (2012). Relieving pain in America: A blueprint for transforming prevention, care, education, and research. Journal of Pain & Palliative Care Pharmacotherapy, 26(2), 197–198. doi:10.3109/ 15360288.2012.678473 Stilwell, P., & Harman, K. (2019). An enactive approach to pain: Beyond the biopsychosocial model. Phenomenology and the Cognitive Sciences, 18(4), 637–665. doi:10.100711097-019-09624-7 Strigo, I. A., & Craig, A. D. (2016). Interoception, homeostatic emotions and sympathovagal balance. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 371(1708), 20160010. doi:10.1098/rstb.2016.0010 PMID:28080968 Sullivan, M. B., Erb, M., Schmalzl, L., Moonaz, S., Noggle Taylor, J., & Porges, S. W. (2018). Yoga therapy and polyvagal theory: The convergence of traditional wisdom and contemporary neuroscience for self-regulation and resilience. Frontiers in Human Neuroscience, 12, 67. doi:10.3389/fnhum.2018.00067 PMID:29535617 Sullivan, M. B., Moonaz, S., Weber, K., Taylor, J. N., & Schmalzl, L. (2018). Toward an explanatory framework for yoga therapy informed by philosophical and ethical perspectives. Alternative Therapies in Health and Medicine, 24, 38–47. PMID:29135457 Taylor, M. J. (2007). What is yoga therapy? An IAYT definition. Yoga Therapy in Practice, 3, 3. Taylor, M. J. (2013). The Science of Mudras. In J. Le Page & L. Le Page (Eds.), Mudras for Healing. Integrative Yoga Therapy. Taylor, M. J. (Ed.). (2015). Fostering creativity in rehabilitation. Nova Science Publishers. Taylor, M. J. (2018). Yoga Therapy as a Creative Response to Pain. Singing Dragon. Taylor, M. J. (2019). The current state(s) and theor(y)ies on pain management (sic). In N. Pearson, S. Prosko, & M. Sullivan (Eds.), Yoga and science in pain care: Treating the person in pain (pp. 52–66). Singing Dragon. Telles, S., Singh, N., & Puthige, R. (2013). Changes in P300 following alternate nostril yoga breathing and breath awareness. BioPsychoSocial Medicine, 7(1), 11. doi:10.1186/1751-0759-7-11 PMID:23721252 Villemure, C., Ceko, M., Cotton, V. A., & Bushnell, M. C. (2013). Insular cortex mediates increased pain tolerance in yoga practitioners. Cerebral Cortex (New York, N.Y.), 24(10), 2732–2740. doi:10.1093/ cercor/bht124 PMID:23696275

238

 Clinicians Applying Yoga Principles and Practices in Pain Care

Vlaeyen, J. W. S., & Crombez, G. (2020). Behavioral conceptualization and treatment of chronic pain. Annual Review of Clinical Psychology, 16(1), 1. doi:10.1146/annurev-clinpsy-050718-095744 PMID:31821023 Williams, A. C., & Craig, K. D. (2016). Updating the definition of pain. Pain, 157(11), 2420–2423. doi:10.1097/j.pain.0000000000000613 PMID:27200490 Wilson-Mendenhall, C. D., Henriques, A., Barsalou, L. W., & Feldman Barrett, L. (2019). Primary interoceptive cortex activity during simulated experiences of the body. Journal of Cognitive Neuroscience, 31(2), 221–235. doi:10.1162/jocn_a_01346 PMID:30277431 Yap, B. W. D., & Lim, E. C. W. (2019). The effects of motor imagery on pain and range of motion in musculoskeletal disorders: A systematic review using meta-analysis. The Clinical Journal of Pain, 35(1), 87–99. doi:10.1097/AJP.0000000000000648 PMID:30222613 Zeidan, F., & Vago, D. R. (2016). Mindfulness meditation–based pain relief: A mechanistic account. Annals of the New York Academy of Sciences, 1373(1), 114–127. doi:10.1111/nyas.13153 PMID:27398643

ADDITIONAL READING Bussing, A., Ostermann, T., Ludtke, R., & Michalsen, A. (2012). Effects of yoga interventions on pain and pain-associated disability: A meta analysis. The Journal of Pain, 13(1), 1–9. doi:10.1016/j. jpain.2011.10.001 PMID:22178433 Chang, D. G., Holt, J. A., Sklar, M., & Groessl, E. J. (2016). Yoga as a treatment for chronic low back pain: A systematic review of the literature. Journal of Orthopedics & Rheumatology, 3, 1–8. PMID:27231715 Cramer, H., Klose, P., Brinkhaus, B., Michalsen, A., & Dobos, G. (2017). Effects of yoga on chronic neck pain: A systematic review and meta-analysis. Clinical Rehabilitation, 31(11), 1457–1465. doi:10.1177/0269215517698735 PMID:29050510 Cramer, H., Lauche, R., Haller, H., & Dobos, G. (2013). A systematic review and meta-analysis of yoga for low back pain. The Clinical Journal of Pain, 29(5), 450–460. doi:10.1097/AJP.0b013e31825e1492 PMID:23246998 Cramer, H., Lauche, R., Langhorst, J., & Dobos, G. (2013). Yoga for rheumatic diseases: A systematic review. Rheumatology, 52(11), 2025–2030. doi:10.1093/rheumatology/ket264 PMID:23934220 Engel, G. L. (1977). The need for a new medical model: A challenge for biomedicine. Science, 196(4286), 129–136. doi:10.1126cience.847460 PMID:847460 Langhorst, J., Klose, P., Dobos, G. J., Bernardy, K., & Hauser, W. (2013). Efficacy and safety of meditative movement therapies in fibromyalgia syndrome: A systematic review and meta-analysis of randomized controlled trials. Rheumatology International, 33(1), 193–207. doi:10.100700296-012-2360-1 PMID:22350253

239

 Clinicians Applying Yoga Principles and Practices in Pain Care

Lovibond, S. H., & Lovibond, P. F. (1995). Manual for the depression anxiety stress scales (2nd ed.). Psychology Foundation. Miller, R. P., Kori, S., & Todd, D. (1991). The Tampa scale: A measure of kinesiophobia. The Clinical Journal of Pain, 7(1), 51–52. doi:10.1097/00002508-199103000-00053 Prosko, S. (2016). Case report: Yoga therapy for an individual with persistent pain. Yoga Therapy Today, 54, 34–38. Schmalzl, L., Powers, C., & Henje Blom, E. (2015). Neurophysiological and neurocognitive mechanisms underlying the effects of yoga-based practices: Towards a comprehensive theoretical framework. Frontiers in Human Neuroscience, 9, 235. doi:10.3389/fnhum.2015.00235 PMID:26005409 Sullivan, M., Moonaz, S., Weber, K., Taylor, J., & Schmalzl, L. (2017). Toward an explanatory framework for yoga therapy informed by philosophical and ethical perspectives. Alternative Therapies in Health and Medicine, 24, 1. PMID:29135457 Wand, B. M., Catley, M. J., Rabey, M. I., O’Sullivan, P. B., O’Connell, N. E., & Smith, A. J. (2016). Disrupted self-perception in people with chronic low back pain. Further evaluation of the Fremantle back awareness questionnaire. The Journal of Pain, 17(9), 1001–1012. doi:10.1016/j.jpain.2016.06.003 PMID:27327235 Ward, L., Stebbings, S., Cherkin, D., & Baxter, G. D. (2013). Yoga for functional ability, pain and psychosocial outcomes in musculoskeletal conditions: A systematic review and meta-analysis. Musculoskeletal Care, 11(4), 203–217. doi:10.1002/msc.1042 PMID:23300142

KEY TERMS AND DEFINITIONS Biopsychosocial: A model developed by George L. Engel in 1977, attempting to explain and describe the various domains of a human being: biological, psychological, and social, in order to allow for a more holistic approach to healthcare, research and education. Biopsychosocial-Spiritual: See Biopsychosocial. The biopsychosocial-spiritual model includes the spiritual care domain of the human and its substantial body of evidence. Chronic Pain: Pain that persists longer than what would typically be expected for a person to experience within the context of a certain situation. The reader is encouraged to review the evolving definitions from the International Association for the Study of Pain. Integrative: An awareness of distinctions/differences between the multiple aspects of the human experience and how they relate and influence one another. Pain: A subjective, complex, conscious, biopsychosocial-spiritual experience and phenomenon with an ever-evolving definition that remains controversial among leading pain researchers and thought leaders. Pain is perhaps something best to be described, rather than defined. Pain Care: Any aspects of care surrounding people in pain, including, but not exclusive to the following: the type, accessibility, feasibility, quantity and quality of care or support from the provider, organization, systems, payers; family and social support systems; self-care within the person in pain

240

 Clinicians Applying Yoga Principles and Practices in Pain Care

and within the provider; best practice guidelines; policies; pain education; and resources for people in pain, families and health providers. Pain Management: See Pain Care. Any aspects of management surrounding people in pain. The word “management” is traditional and suggests an ability to control an emergent, extremely complex phenomenon contrasted with “care” denoting action but with a bit less clinical hubris. Persistent Pain: See Chronic Pain. The term persistent pain is often used interchangeably with chronic pain, however in patient care understood to generate a reduced nocebo effect. Salutogenesis: A term coined by scientist Aaron Antonovsky in 1979 that can be translated as “health creation.” The salutogenic model of health looks at the factors that progress people towards positive health and what causes well-being and thriving, rather than focusing on what causes disease or pathology. Yoga Therapy: There are multiple definitions, historical contexts and modern adaptations as defined in Table 1. Yogopathy: The use of yoga technologies to address symptoms related to a medical diagnosis without the teachings of yoga or incorporation of a broader yogic lifestyle.

241

242

Chapter 14

Yoga and Lifestyle Factors in Heart Disease M. Mala Cunningham University of Virginia, USA

ABSTRACT This chapter reviews cardiovascular risk factors and presents a comprehensive yoga-based intervention for impacting on heart health. Lifestyle interventions and yoga have been shown to positively impact on heart disease, stress levels, inflammation, vagal tone, and homeostasis. The Cardiac Medical Yoga Lifestyle Change program along with the five-point model for heart health (BREAD) provides heart patients with a comprehensive approach for impacting on their disease process and assists patients in developing the necessary strategies and skills for changing their lifestyle.

INTRODUCTION Sir William Osler (1849-1919), is considered the founding father of modern medicine. He taught that “the good physician treats the disease; the great physician treats the patient who has the disease” (Brainy Quote, Osler, 2018). Osler was an early proponent of lifestyle medicine and viewed the mind as an important factor in understanding the etiology and progression of disease. The concept of focusing on “treating the patient who has the disease” is of particular importance in working with heart patients. It has long been established that the factors of stress, personality and lifestyle all play an important role in the development and progression of heart disease. Healthcare providers and patients working together to include psychosocial and lifestyle interventions will enable a more informed decision on treatment options for positively impacting on and potentially reversing heart disease. Lifestyle risk factors that contribute to coronary heart disease include smoking, lack of physical activity, obesity, diet, high blood pressure, high cholesterol, high blood sugar, diabetes, and excessive alcohol (Steptoe et al., 2007; Klainin‐Yobas et al., 2015). Specific behavioral and emotional characteristics such as social isolation, chronic stress, depression, anxiety, and hostility also have the potential to impact on the development and progression of heart disease (Su & Chang, 2018). Medical interventions such as drugs and surgery are certainly important interventions in dealing with acute and chronic symptoms of DOI: 10.4018/978-1-7998-3254-6.ch014

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga and Lifestyle Factors in Heart Disease

cardiovascular disease (CVD), but it is also critical for heart patients to address their lifestyle, behavioral, and psychosocial risk factors, and to participate in strategies for decreasing risks contributing to heart disease. In developing guidelines and strategies for the reduction of risk factors both lifestyle factors and overactivation and under-activation of the autonomic nervous system (stress) are important to assess. When looking at the autonomic nervous system through the lens of the yogic concept of the gunas (qualities) there are similarities that emerge. Sullivan et al. (2018) describe the yogic gunas in correlation to Porges’ (2009) Polyvagal Theory (PVT) and the autonomic nervous system. According to the yogic conceptualization of the gunas, chronic over-activation of the rajasic quality (sympathetic nervous system) or over-activation of the tamasic quality (parasympathetic nervous system) causes a lack of homeostasis (sattvic quality) within the system (Sullivan et al., 2018). Rajas (the innate tendency or quality that drives motion, in Sanskrit) is one of the three gunas or attributes, a philosophical concept developed by Indian philosophy; the other two qualities being sattva (goodness, balance) and tamas (destruction, chaos). Any over-activation of the sympathetic and parasympathetic nervous system contributes to a neurobiological process that stimulates stress-related hormones and chemicals within the body including excess cortisol and glucose (Sullivan et al., 2018). Co-morbid risk factors such as smoking, stress, obesity, diabetes, dyslipidemia and hypertension can contribute to a cascade of events that potentially impact on endothelial dysfunction and an inflammatory response that lead to the development of CVD (Tawakol et al., 2017). Emotional and psychological stress factors are critical components that can contribute to various coronary problems such as arrhythmias, myocardial ischemia, and left ventricular contractile dysfunction. The American Heart Association supports the practice of yoga for lowering blood pressure, improving circulation, muscle tone, respiration and stress (Healthy Living, 2018). As such, the relationship between the yogic conceptualization of the gunas and Porges’ Polyvagal Theory provides a foundation for understanding how yoga therapy can impact on and influence neurobiological mechanisms and contribute to increased healing, self-regulation and resiliency through physical, psychological and behavioral changes (Porges, 2009).

RISK FACTORS The term cardiovascular disease is an umbrella term used to describe the various conditions affecting the heart and circulatory system, i.e. coronary artery disease, stroke, heart attack and aortic disease. Such conditions are affected by risk factors. There are two kinds of risk factors - modifiable and nonmodifiable. Modifiable risk factors include lifestyle and behavioral choices that can be impacted on through changes in diet, exercise, stress, and smoking. Non-modifiable risk factors such as family history, sex, age, and ethnicity can’t be changed but can be impacted upon. The development and progression of cardiovascular disease increases in relation to the number of risk factors that are present (McCarthy, 2015). Participating in a healthy lifestyle and reducing risk factors becomes critical in the management of cardiovascular disease.

243

 Yoga and Lifestyle Factors in Heart Disease

Non-Modifiable Risk Factors Ethnicity and Race There are many factors that influence CVD related to ethnicity and race. Socioeconomic issues and cultural differences in diet and lifestyle choices are among factors that impact on the development of CVD. In the U.S., after factoring out socioeconomic differences, there are disparities in rates of CVD between African Americans and Caucasian Americans. Statistics show that nearly 50 percent of all black adults have some form of cardiovascular disease compared to approximately 33 percent of white adults. Researchers suspect that genetic conditioning may have been an influential factor and that genetic differences predispose adult blacks to high blood pressure (Blacher & Safar, 2007). Statistics also show that Black Americans who are descended from Africa developed physiological protection around salt-sensitivity (Kressin et al., 2010). Salt sensitivity assists in helping the body to conserve water in dry hot climates such as in Africa. However, later generations of Black American descendants show that they remain disproportionately salt sensitive which leads to the body being more susceptible in retaining sodium. This becomes a risk factor through increases in blood volume and blood pressure (Kressin et al., 2010). Black Americans have consistently shown a higher prevalence of risk factors such as diabetes mellitus, hypercholesterolemia and hypertension, all of which consequently puts them at a higher risk for diabetes and CVD (Kressin et al., 2010; Narayan et al., 2003; Egan et al., 2010).

Family History A family history of cardiovascular disease can be viewed from both a genetic and a family lifestyle influence. If an individual’s father or brother develops CVD before the age of 55 or if a mother or sister develops CVD before the age of 65 these factors are considered a genetic influence (Miller et al., 1995). Factors such as a family history of diabetes, high blood pressure and high cholesterol are also taken into consideration related to risk for CVD. Having a genetic influence or a family history of CVD does not necessarily lead to CVD, but given the potential of genetic predisposition it is recommended that individuals be proactive in their reduction of modifiable risk factors.

Age Age is considered an independent but non-modifiable risk factor for the development of CVD. The risk associated with aging is considered a causal relationship, which is dependent upon the partnering of other risk factors such as smoking, obesity, and lack of exercise. Independent factors that contribute to aging and CVD include an increase in blood pressure due to consequential structural changes in the aging arteries which includes large arterial stiffness (Blacher & Safar, 2007). Age-related differences can also be attributed to increases in serum total cholesterol levels, body mass indexes, and potential diabetes prevalence. It is known that about 4 out of 5 deaths due to heart disease occur in individuals older than 65 years (Blacher &Safar, 2007).

244

 Yoga and Lifestyle Factors in Heart Disease

Gender Difference Although cardiovascular disease is the leading cause of death worldwide there are important gender differences that exist between men and women with CVD. A 2013 U.S. analysis of CVD showed a higher level of female deaths compared to male deaths due to CVD. Although more women die of stroke and CVD than men, statistics show that women have more hospital discharges for heart failure and stroke (Rapsomaniki et al., 2015). The prevalence of death from CVD for females is at 35 percent whereas the number of deaths for males is at 32 percent (Hammond et al., 2013). Despite the higher percentage of CVD deaths in women, CVD is still widely considered a man’s disease. Coronary artery disease (CAD) statistics show that more men are living with and dying of CAD than women, and that men have more hospital discharges for coronary artery disease (Mosca et al., 2011). Due to sex hormones, women are usually protected from heart disease until menopause, after which, their risk increases in relation to men (Mosca et al., 2011; Rapsomanikiet al., 2015; Leening et al., 2014; Global Burden of Disease, 2013).

Modifiable Risk Factors Cigarette / Tobacco Smoking Smoking has serious detrimental effects on the heart and circulatory system and is considered a high-risk factor for heart disease. According to the Center for Disease Control, individuals who smoke increase their risk of stroke and heart disease by 2-4 times (Centers for Disease Control, Tobacco, 2015). Chemicals in tobacco, such as nicotine cause vasoconstriction which increases cellular demand for oxygen but make cells less receptive to oxygen supply. Nicotine also increases endothelial dysfunction and systemic inflammation (National Academy of Sciences Institute of Medicine, 2009; U.S. Dept. of Health and Human Services, Office on Smoking and Health, 2014). Secondhand smoke also increases the risk for CVD. It is estimated that those exposed to secondhand smoke increase their risk of death due to heart disease by approximately 30 percent (National Academy of Sciences Institute of Medicine, 2009). However, once smokers quit, they are typically able to impact on reducing their risk of heart disease despite the length and frequency of their smoking habit. Efforts to stop smoking can be challenging, and it is estimated that only 6 percent of all smokers each year are successful at smoking cession (CDC, 2011). The negative effects of smoking cessation include increased irritability, negative affect, cravings, and increased stress. These corollary factors have a direct negative impact on smoking relapse and cravings (Rosen et al., 2016). Yoga practices have been shown to have a promising impact on reducing negative symptoms associated with smoking cessation. Rosen et al. (2016) found that yoga interventions such as postures, breathing, and relaxation positively impacted on the reduction of stress and cigarette cravings and helped increase body awareness and feelings of well-being.

Hypertension Hypertension is known as the “silent killer” because it can go undetected for a long time. Under normal conditions the blood vessel walls are flexible and have a degree of elasticity to them, but with constant high pressure on the interior walls of the blood vessels they can lose elasticity and microscopic tears 245

 Yoga and Lifestyle Factors in Heart Disease

can occur (Blacher et al., 2016). The loss of elasticity and microscopic tears lay the groundwork for fat and cholesterol to be deposited in the walls of the arteries. This causes a narrowing of the blood vessels and forces the heart to work harder to push the same amount of blood through narrowed arteries (Rapsomaniki et al., 2014). Contributing factors to hypertension include obesity, high salt intake, excess alcohol, family history, and physical inactivity. In some cases, there are no apparent causative factors that can be detected in individuals with hypertension. It is recommended that blood pressure be monitored and kept at less than 120/80 mmHg (Whelton et al., 2018; Angeli et al., 2014).

Cholesterol Cholesterol is a fat (or lipid) and is a normal substance found in muscles, red blood cells and cell membranes. High levels of low-density lipoprotein cholesterol (LDL - also known as bad cholesterol) are linked to a range of cardiovascular diseases. If too much LDL is present in the system, it can cause fatty substances to build up in the artery walls and lead to complications. LDL cholesterol picks up cholesterol from ingested fats and delivers it to blood vessels and muscles where it is deposited as fat. Although the liver produces some cholesterol, the majority of cholesterol is typically ingested through saturated fats such as egg yolks, animal fats, and oils. The liver metabolizes normal levels of cholesterol, but it does not have the ability to break down excessive levels of ingested fat. As such, some fat is stored, and some is converted to cholesterol and has the potential to be deposited in the arteries. High levels of LDL cholesterol are often caused by factors such as an unhealthy diet, smoking, physical inactivity, high alcohol intake and liver and kidney disease. It is recommended that individuals exercise, eat a balanced diet, and quit smoking in order to reduce LDL levels. HDL is categorized as the “the good cholesterol” and has the potential to remove excess cholesterol from the body. Aerobic exercise, smoking cessation, and eating a heart healthy diet can also positively impact on HDL levels in the body (Wang et al., 2017). Smoking cessation in particular, was found to significantly improve total HDL, HDL-C, and large HDL particles. These findings were specifically strong in women (Gepneret at el., 2011). Yoga practices have also been found to have a significant impact on serum total cholesterol, triglycerides and the reduction of LDL cholesterol (Manchnada et al., 2000). In a study by Shantakumari (2013) they found that the practice of yoga had a statistically significant impact on the reduction on body weight, body mass index, total cholesterol levels, triglycerides and on lower LDL cholesterol and on higher HDL cholesterol levels. Normal cholesterol levels range between 150-200 ml/dl (WebMD, 2018a). A moderate cholesterol risk is considered to be between 200-240 ml/dl. A high risk for the development of coronary artery disease is considered anything over 240 ml/dl (Grover et al., 1995). Total blood cholesterol that is over 200 mg/dl can potentially contribute to the build-up of plaque on artery walls and cause them to narrow. Some experts believe the ratio of HDL to LDL is more important than the total cholesterol, while others believe that as long as the total blood cholesterol is under 200 the ratio is unimportant (WebMD, 2018a). The American Heart Association recommends a total fat intake of less than 25-30 percent per day (American Heart Association, 2018). The Ornish research study on reversing and preventing coronary artery disease recommends fat intake at 10 percent per day (Everyday Health, Ornish Diet, 2018). Ornish also advises a replacement of saturated fat with mono-unsaturated or poly-unsaturated fats and to increase the amount of complex carbohydrates in one’s diet (American Heart Association, Cholesterol, 2018).

246

 Yoga and Lifestyle Factors in Heart Disease

Medication may be recommended if efforts to reduce high total cholesterol levels are not reduced through lifestyle changes. Optimal levels of cholesterol and triglycerides are: Total Cholesterol Less than 200 mg / dL LDL (bad) Cholesterol Less than 100 mg / dL HDL (good) Cholesterol At least 60 mg / dL Triglycerides Less than 150 mg / dL (Harvard Health, Making Sense of Cholesterol Tests, 2018).

Lack of Exercise / Sedentary Lifestyle A sedentary lifestyle is a known risk factor for developing heart disease. Exercise on the other hand has been found to contribute to the reduction of stress, high blood pressure, obesity, and can help increase HDL levels (Thompson et al., 2007). Regular exercise has been found to significantly decrease CVDrelated mortality and to reduce the risk of having an MI (Anderson et al., 2016). Exercise has many potential benefits which include strengthening the heart, helping the heart to work more efficiently, and improving risk reduction for CVD (Platt et al., 2015). Aerobic exercise constitutes the use of large muscle groups and helps the heart to pump faster to meet oxygen demands in the body (Shata et al., 2011). Even modest levels of low-intensity aerobic activity can make a difference in helping to impact on lowering the risk of heart disease (Shata et al., 2011). However, for maximum benefit it is recommended to do aerobic exercise three to four times per week for 30-60 minutes. Examples of aerobic exercise include swimming, biking, running and brisk walking.

Obesity In many countries the incidence and prevalence of childhood and adult obesity are increasing at alarming levels. The World Health Organization (2018) reports that since 1975 worldwide levels of obesity have nearly tripled. Obesity is a risk factor that impacts on a variety of health issues and is correlated with numerous co-morbidities such as type 2 diabetes, hypertension, certain cancer and cardiovascular diseases (Poirier et al., 2006). Studies have shown that when an individual is more than ten percent over their ideal body weight it puts more strain on the heart muscle and can impact on dyslipidemia, glucose intolerance, inflammatory markers, hypertension, and obstructive sleep apnea (Poirier et al., 2006). As a general rule, optimal waist circumference is recommended at less than 40 inches for men. For Asian men the recommendation is 37-39 inches. The optimal waist circumference for women is less than 35 inches and for Asian women it is 31-35 inches. Staying within recommended weight standards can potentially influence the lowering of blood pressure and improve both cholesterol and blood sugar levels (National Institute of Health, National Heart, Lung, & Blood Institute, 1998). The reduction of obesity is critical to the health of heart patients and yoga has been found to positively impact on obesity levels. In one study, Telles, et al. (2018) found that the practice of yoga along with a dietary plan consisting of 1,900-2,000 Kcal/day significantly reduced anthropometric measures associated with central obesity compared to the control group (who received only nutritional advise similar to the experimental group). Differences were found to be higher in the 30-45 years age range. Similarly, Yadav et al., (2016) found that a yoga-based intervention of postures, breathing, and relaxation strategies (along with group support, lectures, individual guidance and nutrition information) was effective in the improvement of the quality of life and overall health. Additionally, after only 10 days of a yoga 247

 Yoga and Lifestyle Factors in Heart Disease

intervention the following factors were reduced: body weight, BMI (body mass index), total body fat, blood pressure, total cholesterol, low-density lipoprotein, triglycerides, fasting glucose, and waist and hip circumference, and waist to hip ratio.

Diabetes Diabetes occurs when the body is not able to make enough of a hormone called insulin or cannot effectively use the insulin that it has. Insulin is a critical component for helping the body digest glucose (sugars) that are ingested from foods. Without the proper balance of insulin, glucose can build up in the blood and can cause damage to the heart, kidneys, nerves, and blood vessels (American Diabetes Association, 2016). When there is damage to the blood vessels (such as to the endothelial lining of the arteries) it provides an opportunity for cholesterol to be deposited in the lining of blood vessel walls leading to the development of atherosclerotic plaque and to the development of coronary artery disease (Mozaffarian et al., 2016; Centers for Disease Control and Prevention, 2015; American Diabetes Association, 2016). The risk of developing coronary artery disease is twice as great in diabetic men and three times as high in diabetic women (Peters et al., 2014). The incidence and prevalence of diabetes in the U.S. is increasing in all age groups and in particular, in racial and ethnic groups such as African Americans and Hispanics (American Diabetes Association, 2016). The seventh leading cause of death in the U.S. is diabetes, and effects 1out of every 11 people. Another 86 million individuals are affected with prediabetes wherein their glucose levels are high but not yet high enough to trigger diabetes (American Diabetes Association, 2016). Diabetes is a serious and potentially life-threatening disease that requires monitoring and management. The impact of diabetes can have serious detrimental effects on the system and co-morbid conditions are a common occurrence. Researchers have found the practice of yoga to have a positive impact on several diabetic symptoms. Some examples include increased insulin sensitivity, improved glucose tolerance, and improved lipid metabolism (Raveendran et al., 2018). Yoga also has been shown to have a significant impact on common co-morbid conditions associated with diabetes which include hypertension, obesity, hyperlipidemia, chronic kidney disease, cardiovascular disease and mood dysregulation (Raveendran et al., 2018; Iglay et al., 2016; Hagins et al., 2013; McDermott, et al., 2014). It is recommended that individuals who have diabetes actively participate in controlling their diet, weight levels, stress, exercise and medications. All efforts in this regard can contribute to helping decrease the risk of kidney disease, stroke and heart attack.

Stress Chronic stress is finally being recognized as an attributable risk factor that is of equal concern as smoking, hypertension, diabetes and lipid concentrations (Tawakolet al., 2017). External stressors such as low socioeconomic status, marital strife, job dissatisfaction and family caregiving has been shown to have an impact on the pathogenesis of cardiovascular disease (Tafetetet al., 2016; Harvard Health, 2018). Additionally, internal stressors related to personality factors and neurobiological wiring have also been implicated in impacting on risk factors and biological processes that affect oxidative stress within the body (Tafet et al., 2016; Kivimäki et al., 2006; Pollitet al., 2006). Both internal and external stressors signal the activation of the HPA axis (hypothalamus, pituitary, adrenal axis) and creates a cascading chemical effect that, when chronic, can involve an inflammatory and vasoactive response in the blood 248

 Yoga and Lifestyle Factors in Heart Disease

vessels leading to a compromising of the endothelial lining in the coronary arteries (Tawakol et al., 2017). Under the chronic stress response cycle other disruptive processes also occur such as an increase in blood pressure, respiration, and heart rate (Bunker et al., 2003). Investigation between psychosocial stress and cardiovascular risks show that resting metabolic activity within the amygdala was significantly associated with CVD events. The relationship between amygdala activity and CVD was substantially mediated by arterial inflammation and in turn was mediated by upregulated bone-marrow activity (Tawakol et al., 2017). These observations target amygdala activity as a key stress mechanism that causes upregulation of haemopoietic tissue activity and increased atherosclerotic inflammation and can be implicated in a neuronal-haemopoietic-arterial axis (Tawakol et al., 2017). Personality factors (both state and trait components) and lifestyle and behavioral choices are all known to have a neurobiological impact on the activation of the stress response. The over-activation of the sympathetic nervous system (rajas quality) can lead to the chronic circulation of elevated stress hormones such as cortisol, epinephrine and norepinephrine throughout the body (Moreno et al., 2017). This cascading effect and over-activation of stress contribute to elevated blood pressure, an increase in heart and respiratory rates, reduced resiliency and an inability to self-regulate (Moreno et al., 2017). Additionally, chronic activation of cortisol (a stress hormone) is disruptive to sleep, and in high levels, is considered a neurotoxin causing a compromising of memory and a dysregulation of the emotions (Moreno et al., 2017; Holzel et al., 2010). In a similar manner, a threatening and high stress event can also cause an over-activation of the parasympathetic nervous system (tamas quality) and arousal of the dorsal vagal complex (DVC). Under severe threat, the DVC sets in motion a metabolic inhibitory process causing decreased output in such systems as respiration, heart function and digestion. Immobilization, fainting, and dissociative states are also reflective of extreme activation of the dorsal vagal complex (Sullivan et al., 2018). When the system is not under extreme trauma or threat, activation of the DVC is associated with such states as clinical depression, lethargy, passivity, and lack of motivation. From a yogic perspective the desirable guna (quality) to strive for is the sattvic state. The sattvic guna is correlated with the ventral vagal complex (VVC) of the parasympathetic nervous system and is associated with psychological and neurobiological states that reflect safety, homeostasis, and social engagement. Yoga has long been known to be a self-regulating and self-healing practice that impacts on neurobiological activation caused by stress. Yoga practices are aimed at restoring the system to homeostasis through increased regulation of the ventral vagal complex (Sullivan, et al., 2018; Porges & Carter, 2017).

Alcohol and CVD The relationship between alcohol and CVD is complicated in large part because of the influence of other lifestyle risk factors that may be present and influential. Continuous and high levels of alcohol have the potential to affect the heart by exposing it to toxic and poisonous agents. This process can cause a compromising of the heart muscle leading to cardiomyopathy (Chen et al., 2008). High levels of alcohol intake can also impact on vitamin deficiencies which can lead to damage to the heart muscle (WebMD, 2018b). More studies are needed to assess the effects of alcohol in correlation to individual risk factors - including factoring in the caloric intake of the alcohol consumed. Despite the complexities, studies have consistently shown that the consumption of high levels of alcohol can contribute to heart-related

249

 Yoga and Lifestyle Factors in Heart Disease

problems such as cardiomyopathy, stroke, irregular heartbeats and high blood pressure (Chen et al., 2008; Djoussé et al., 2009; WebMD, 2018b). The recommended guidelines for alcohol intake include low to moderate levels consisting of 1-2 drinks per day for men and 1 drink for women. One drink is defined as 1½ fluid ounces of 80-proof spirits, 1 fluid ounce of 100-proof spirits, 4 fluid ounces of wine, or 12 fluid ounces of beer (WebMD, 2018b).

THE EFFECTS OF YOGA ON HEART DISEASE AND LIFESTYLE FACTORS The American Heart Association supports the practice of yoga for improving circulation, muscle tone, increasing respiratory function and heart rate, and decreasing blood pressure levels (American Heart Association, 2017). Numerous studies have provided statistically significant evidence for the positive effect of yoga on CVD. Leading the way has been Dr. Dean Ornish who provides decades of research on using a yoga-based intervention for reversing and preventing coronary heart disease (Ornish et al., 1990; Ornish et al.,1998; Pischke et al., 2008; Silberman et al., 2010). Randomized control trials have consistently documented that patients with moderate to severe levels of heart disease who participated in lifestyle changes and utilized a yoga-based intervention were able to reverse their heart disease at statistically significant levels (Ornish et al., 1990; Ornish et al., 1998). In one study, exercise radionuclide ventriculography showed that the experimental group ejection fraction (EF) levels rose after only 24 days of changes in lifestyle and diet, whereas the control group showed significant decline in EF response. Further results indicated a 91 percent reduction in angina in the experimental group versus a 165 percent increase in reported angina in the control group. Most notably, there were indications of a reduction in anatomic severity of disease related to improved endothelial function, reduced microvascular ischemia, and improved collateral flow (Ornish et al., 1990). In another randomized control trial utilizing a yoga-based, lifestyle change intervention, statistically significant outcomes showed that patients in the experimental group (after one year in the program) showed a reversal of diameter stenosis from 16.9 percent to 16.5 percent (Ornish et al., 1998; Pischke et al., 2008; Silberman et al., 2010). The control group outcomes showed a progression of diameter stenosis from 15.5 percent to 18.5 percent. Overall, 82 percent of the patients in the experimental group realized a positive regressive change in diameter stenosis. In a 5 year follow up of the same subjects, the experimental group continued to show regression of diameter stenosis and the control group continued to realize progression of diameter stenosis. During the 5 years the experimental group also showed 2.5 times fewer cardiac events than the control group (Silberman et al., 2010; Ornish et al., 2007; Ornish et al., 1990; Pischke et al., 2008). In another research study, Silberman et al. (2013) investigated a yoga-based lifestyle change program that was conducted through multi-site Cardiac Rehab programs over a one-year period of time. The study looked at 2,974 men and women and used paired T-tests to assess differences in: 1) baseline values and 12-week values, 2) baseline values to 1-year values; and 3) 12-week values to those at 1 year. After 12 weeks 88 percent of the patients remained enrolled in the study, and after one year 78.1 percent were enrolled. The results of this yoga-based and lifestyle change program were impressive. Further analysis revealed that the subjects had statistically significant results in the areas of body mass index (BMI), triglycerides, low density lipoprotein cholesterol, total cholesterol, hemoglobin A1c, systolic blood pressure, diastolic blood pressure, depression, hostility, exercise, and functional capacity. These differences remained consistent after one year (Silberman et al., 2010). 250

 Yoga and Lifestyle Factors in Heart Disease

It is known that risk factors such as smoking, obesity, diabetes, stress, and dyslipidemia can have a direct disruptive effect on the endothelial lining in the coronary arteries (Tawakol et al., 2017). The consequential negative effect of these risk factors and the neurobiological HPA stress cascade can lead to inflammation and vasoactive mediators such as fibrinogen, C-reactive protein, interleukin-6, and tumor necrosis factor-alpha (TNF) contributing to the development of CVD (Tawakol et al., 2017). Studies have shown that yoga has the ability to impact on these inflammatory biomarkers by influencing the parasympathetic nervous system (the ventral vagal pathway) and vagal tone. As such, in the cascading chain of neurobiological events, a yoga-based and mindfulness practice can potentially interrupt the autonomic and chronic HPA stress activation process before it reaches a consequential inflammatory response pattern leading to CVD (Sarvottam et al., 2013; Manchanda, 2014; Dzauet al., 2006; Chu et al., 2016; Khalsa et al., 2016). Yoga is a comprehensive treatment option that embraces the physical, mental, emotional and spiritual dimensions of healing. As a stress and life management intervention, yoga impacts on restoring homeostasis and reducing allostatic load (allostatic load refers to an elevated or prolonged chronic stress response that has negative physiologic and neuroendocrine consequences). It is well documented that heart patients are routinely affected by stress, anxiety and depression throughout their diagnosis and in the management of CVD (Tafet et al., 2016; Kivimäki et al., 2006; Pollitt et al., 2005; Chen, et al., 2008). Yoga provides tools, techniques and strategies for disrupting the stress response which includes management of the HPA axis and the cascade of neurobiological events leading to a dysregulation of homeostasis and an increase in inflammation. According to Peters et al. (2017) allostatic load can permanently alter brain structure and systemic pathophysiology. Further, they posit that “the neuroendocrine, cardiovascular, neuroenergetic, and emotional responses become persistently activated so that blood flow turbulences in the coronary and cerebral arteries, high blood pressure, atherogenesis, cognitive dysfunction and depressed mood accelerate disease progression” (Peters et al., 2017). For obvious reasons, managing stress on all levels becomes crucial for the CVD patient.

CARDIAC MEDICAL YOGA: AN INTERVENTION FOR HEART PATIENTS The Cardiac Medical Yoga® program is a yoga-based intervention developed by Cunningham (1997) in consultation with cardiologists and cardiac rehab staff from the University of Virginia (Cunningham, 2010). The Cardiac Medical Yoga program has been designed as both a lifestyle intervention program for heart patients as well as a teacher training certification program for healthcare workers and others. Hospitals throughout the U.S. and abroad have enrolled their healthcare staff in this certification program thereby enabling heart patients access to the Cardiac Medical Yoga Lifestyle Change program. The Cardiac Medical Yoga program for heart patients is both a health education and lifestyle change program that emphasizes the reduction of risk factors and stress by providing patients with specific tools and strategies for increasing health, resiliency, and self-regulation. The program focuses on a reorganization of negative lifestyle patterns, a reduction of both internal and external stressors (with an emphasis on reducing inflammation and establishing homeostasis through vagal toning), and teaches patients how to manage and impact on risk factors through learning life management skills. The program embraces a Bio-Psycho-Social-Spiritual model of care and includes an initial comprehensive assessment as well as specific recommendations for realizing targeted heart healthy goals.

251

 Yoga and Lifestyle Factors in Heart Disease

The Cardiac Medical Yoga classes, consults, and retreat programs are held in relaxing and healing environments where participants learn gentle and modified yoga postures, deep relaxation, vagal toning, stretching, breathing, guided imagery, healthy dietary tips, lifestyle management strategies and stress reduction techniques. Prior to participating in the Cardiac Medical Yoga program patients are encouraged to obtain their doctors approval and to view the program as a complement to their existing treatment plan.

BREAD™ The acronym BREAD™ is used to encapsulate the Cardiac Medical Yoga intervention program. Patients are exposed to an educational and training program that includes: 1) Breathing Practices, 2) Relaxation Strategies, 3) Exercise / Yoga instruction, 4) Attitudinal Exploration and Psychosocial interventions, and 5) Dietary Tips. Due to space limitations it is not possible to review the entirety of the Cardiac Medical Yoga program. Listed below is a representative sampling of each of the targeted areas that embrace the BREAD concept.

Breathing - Sample Practices Due to stress, surgery, emotional trauma and various other factors, many patients have compromised breathing patterns. In the Cardiac Medical Yoga program patients start with the basics of re-learning diaphragmatic, even, and deep breathing. Once these basic practices are established, patients can then progress to other breathing techniques such as Alternative Nostril Breathing and the Calming Breath Technique. Yogic breathing practices have a powerful impact on assisting individuals in interrupting the neurobiological stress pattern, restoring homeostasis, reducing allopathic load and impacting on vagal health and functioning. A sampling of the breathing practices include: 1. 2. 3. 4. 5.

Diaphragmatic & Even Breathing Deep Breathing / Revitalizing Breath Calming Breath Technique Alternative Nostril Breathing Other breathing practices aimed at disrupting the HPA axis stress response and improving vagal tone

Relaxation Strategies In the beginning stages, some patients may not have full awareness of the need to re-wire their neurobiological patterns and nervous system. Over time, as patients learn to navigate internal cues and stress, they start to recognize where there may be dysregulation and over-activation of the autonomic nervous system. Awareness techniques and relaxation strategies assist patients in developing a deeper connection to proprioception and interoception processes which can lead to a greater capacity for self-regulation and healing. Utilizing relaxation strategies and identifying stress patterns earlier on in the casual chain of events empowers an individual in increasing resiliency and on impacting on risk factors. Here is a sampling of relaxation strategies.

252

 Yoga and Lifestyle Factors in Heart Disease

1. Deep Calming Relaxation (Yoga Nidra) 2. In-Motion Relaxation Strategies (learning to relax while in action) 3. Learning to identify and disrupt cognitive and behavioral patterns that activate the HPA axis stress response 4. Learning to utilize breath, postures, imagery, and cognitive revamping to activate relaxation (the ventral vagal component of the parasympathetic nervous system)

Exercise / Asana’s (postures) Depending on the physical capacity of an individual there are a variety of options for the patient to begin a yoga lifestyle change program related to postures. Patients have the option of starting with a Hospital Bed Yoga™ format and progressing to Limited Mobility Yoga™. Limited Mobility Yoga consists of chair yoga, chair-assisted yoga, wall assisted yoga, and gentle, modified yoga practices. Below is a representative sample of options for doing the Sun Salutation starting in a chair and progressing to chair-assisted and wall-assisted movements. Cardiac Yoga Medical Program has been shown in Figure 1.

Attitudinal Interventions Heart patients who are attempting to make changes in their lifestyle are often challenged by the emotional and psychological correlates of anxiety, depression, stress, and grieving (Bunker et al., 2003). It is not uncommon to find patients seeking emotional comfort through self-medicating behaviors such as smoking, chronic drinking, inactive lifestyle, and indulging in foods that are high in saturated fat, sugar, and carbohydrates. These behaviors often involve issues of denial, grieving and loss, and contributes to a downward cycle of dysregulated risk (Cunningham, 2010). For many patients, the diagnosis of CVD itself sets in motion psychological and emotional correlates that effect an individual’s sense of identity, value, self-esteem, and self-determination (Cunningham, 2010). The psychosocial and attitudinal aspects of the Cardiac Medical Yoga program whether in a group setting or individually, assists patients in: 1) setting heart healthy goals, 2) addressing and reframing negative thoughts and adverse behaviors, 3) understanding co-morbid issues of anxiety, depression, grieving and stress, and 4) improving issues of compliance, motivation and success. In general, the Cardiac Medical Yoga program offers support and education, and provides patients with specific tools, resources, techniques, and strategies for changing and impacting on risk factors and dysregulated behaviors. A sampling of topics that might be discussed in a group setting include: 1. 2. 3. 4.

Setting meaningful and realistic heart healthy goals Navigating emotional distress including anxiety, depression and other mood states Managing loss and grief issues related to the diagnosis and management of CVD Educational opportunities for how to successfully manage stress, risk factors, and negative behaviors that impact on heart disease

253

 Yoga and Lifestyle Factors in Heart Disease

Figure 1. Cardiac Yoga Medical Program

Dietary Tips The Cardiac Medical Yoga program supports dietary recommendations that have been formatted by a patient’s physician or nutritionist, and encourages a diet that is tasty, healthy, and close to nature. The role of the Cardiac Yoga Consultant or Teacher is one of motivation, encouragement, and assistance in adhering to a heart healthy diet. Adhering to a heart healthy diet can be challenging for heart patients and compliance tends to fluctuate quite dramatically. The emotional and psychological support they receive from the Cardiac Medical Yoga group becomes invaluable, as well as motivational, and contributes to higher levels of compliance in managing dietary recommendations and other risk factors.

254

 Yoga and Lifestyle Factors in Heart Disease

CONCLUSION Studies have shown that the diagnosis of CVD can create high levels of emotional and cognitive dissonance in heart patients which can potentially impede any reasonable attempt to impact on CVD risk factors (Tafet et al., 2016; Kivimäki et al., 2006). The Cardiac Medical Yoga program provides both support and educational assistance to heart patients and can impact on increasing emotional resiliency, self-regulation and healing. Given the typical co-morbid issues of depression, anxiety and grieving, patients need support, assistance, and tools as part of their recovery process. Participation in yoga-based programs offer heart patients strategies and techniques to assist them in reducing both risk factors and dysregulated behaviors within an environment that includes emotional support and that starts with small manageable steps to help build a patient’s confidence (Cunningham, 2010). Yoga-based programs are also cost-effective and have been found to complement conventional care through seamless integration into existing Cardiac Rehab programs (Cunningham, 2010). Since CVD is the number one killer for both men and women in many countries and contributes to high, escalating healthcare costs (Hammond et al., 2013) the need for continued implementation of yoga-based programs that are effective and low-cost remains high.

REFERENCES American Diabetes Association. (2017). Statistic about diabetes. Centers for Disease Control and Prevention, U.S. Department of Health and Human Services. Retrieved from https://www.cdc.gov/diabetes/ pdfs/data/statistics/national-diabetes-statistics-report.pdf American Heart Association. (2017). Yoga and heart health. Retrieved from http://www.heart.org/ heartorg/healthyliving/physicalactivity/UCM_434966 American Heart Association. (2018). Health Topics Cholesterol. Retrieved from https://www.heart.org/ en/health-topics/cholesterol/about-cholesterol/what-your-cholesterol-levels-mean Anderson, L., Thompson, D. R., Oldridge, N., Zwisler, A. D., Rees, K., Martin, N., & Taylor, R. (2016). Exercise-based cardiac rehabilitation for coronary heart disease. Cochrane Database of Systematic Reviews, 1, CD001800. PMID:26730878 Angeli, F., Reboldi, G., & Verdecchia, P. (2014). Hypertension, inflammation and atrial fibrillation. Journal of Hypertension, 32(3), 480–483. doi:10.1097/HJH.0000000000000112 PMID:24477094 Blacher, J., Levy, B. I., Mourad, J. J., Safar, M. E., & Bakris, G. (2016). From epidemiological transition to modern cardiovascular epidemiology: Hypertension in the 21st century. Lancet, 388(10043), 530–532. doi:10.1016/S0140-6736(16)00002-7 PMID:26856636 Blacher, J., & Safar, M. E. (2007). Large-artery stiffness, hypertension and cardiovascular risk in older patients. Nature Clinical Practice. Cardiovascular Medicine, 2(9), 450–455. doi:10.1038/ncpcardio0307 PMID:16265585 Brainy Quote. (2018). Osler quotes. Retrieved from https://www.brainyquote.com/authors/

255

 Yoga and Lifestyle Factors in Heart Disease

Bunker, S., Colquhoun, D., Esler, M., Hickie, I., Hunt, D., Jelinek, V. M., Oldenburg, B. F., Peach, H. G., Ruth, D., Tennant, C. C., & Tonkin, A. M. (2003). Stress and coronary heart disease: Psychological risk factors. The Medical Journal of Australia, 178(6), 272–276. doi:10.5694/j.1326-5377.2003. tb05193.x PMID:12633484 Centers for Disease Control. (2015). Tobacco. Retrieved from www.cdc.gov/tobacco/basic information/ health effects/heart_disease/index.htm Centers for Disease Control and Prevention. (2015). Number (in millions) of civilian, non-institutionalized persons with diagnosed diabetes. Author. Chen, L., Smith, G. D., Harbord, R. M., & Lewis, S. J. (2008). Alcohol intake and blood pressure: A systematic review implementing a Mendelian randomization approach. PLoS Medicine, 5(3), e52. doi:10.1371/journal.pmed.0050052 PMID:18318597 Chu, P., Gotink, R. A., Yeh, G. Y., Goldie, S. J., & Hunink, M. M. (2016). The effectiveness of yoga in modifying risk factors for cardiovascular disease and metabolic syndrome: A systematic review and metaanalysis of randomized controlled trials. European Journal of Preventive Cardiology, 23(3), 291–307. doi:10.1177/2047487314562741 PMID:25510863 Chu, P., Pandya, A., Salomon, J., Goldie, S., & Hunink, M. G. (2016). Comparative effectiveness of personalized lifestyle management strategies for cardiovascular disease risk reduction. Journal of the American Heart Association, 5(3), e002737. doi:10.1161/JAHA.115.002737 PMID:27025969 Cramer, H., Lauche, R., Haller, H., Steckhan, N., Michalsen, A., & Dobos, G. (2014). Effects of yoga on cardiovascular disease risk factors: A systematic review and meta-analysis. International Journal of Cardiology, 173(2), 170–183. doi:10.1016/j.ijcard.2014.02.017 PMID:24636547 Cunningham, M. (2010). The cardiac yoga teacher training manual. Djoussé, L., Lee, I. M., Buring, J. E., & Gaziano, J. M. (2009). Alcohol consumption and risk of cardiovascular disease and death in women: Potential mediating mechanisms. Circulation, 120(3), 237–244. doi:10.1161/CIRCULATIONAHA.108.832360 PMID:19597054 Dzau, V. J., Antman, E. M., Black, H., Manson, J. E., Plutzky, J., Popma, J., & Stevenson, W. (2006). The cardiovascular disease continuum validated: Clinical evidence of improved patient outcomes: Part 1: Pathophysiology and clinical trial evidence risk factors through stable coronary artery disease. Circulation, 114(25), 2850–2870. doi:10.1161/CIRCULATIONAHA.106.655688 PMID:17179034 Eckel, R. H., York, D. A., Rossner, S., Hubbard, V., Caterson, I., St Jeor, S. T., … Blair, S. N. (2004). American Heart Association. Prevention Conference VII: Obesity, a worldwide epidemic related to heart disease and stroke: executive summary. Circulation, 110, 2968–2975. Egan, B., Zhao, Y., & Axon, R. (2010). US trends in prevalence, awareness, treatment, and control of hypertension, 1988-2008. Journal of the American Medical Association, 3(20), 2043–2050. doi:10.1001/ jama.2010.650 PMID:20501926 Everyday Health. (2018). Ornish diet. Retrieved from https://www.everydayhealth.com/diet-and-nutrition/ the-ornish-diet.aspx

256

 Yoga and Lifestyle Factors in Heart Disease

Ford, E., Li, C., Pearson, W., Zhao, G., & Mokdad, A. (2010). Trends in hypercholesterolemia, treatment and control among United States adults. International Journal of Cardiology, 140(2), 226–235. doi:10.1016/j.ijcard.2008.11.033 PMID:19081646 GBD 2013 Mortality and Causes of Death Collaborators. (2015). Mortality and causes of death collaborators. Global, regional, and national age-sex specific all-cause and cause-specific mortality for 240 causes of death, 1990–2013: A systematic analysis for the Global Burden of Disease Study 2013. Lancet, 385, 117–171. Gepner, A. D., Piper, M. E., Johnson, H. M., Fiore, M. C., Baker, T. B., & Stein, J. H. (2011). Effects of smoking and smoking cessation on lipids and lipoproteins: Outcomes from a randomized clinical trial. American Heart Journal, 161(1), 145–151. doi:10.1016/j.ahj.2010.09.023 PMID:21167347 Grover, S., Coupal, L., & Xiao-Ping, H. (1995). Identifying adults at increased risk of coronary disease: How well do the current cholesterol guidelines work. Journal of the American Medical Association, 274(10), 801–806. doi:10.1001/jama.1995.03530100041032 PMID:7650803 Hagins, M., States, R., Selfe, T., & Innes, K. (2013). Effectiveness of yoga for hypertension: Systematic review and meta-analysis. Evidence-Based Complementary and Alternative Medicine, 2013, 649836. doi:10.1155/2013/649836 PMID:23781266 Hammond, G., Mochari-Greenberger, H., Twofighi, A., & Albert, M. A. (2013). Fifteen-year trends in awareness of heart disease in women: Results of a 2012 American Heart Association national survey. Circulation, 127(11), 1254–1263. doi:10.1161/CIR.0b013e318287cf2f PMID:23429926 Hartley, L., Dyakova, M., Holmes, J., Clarke, A., Lee, M. S., Ernst, E., & Rees, K. (2014). Yoga for the primary prevention of cardiovascular disease. Cochrane Database of Systematic Reviews, 5, CD010072. doi:10.1002/14651858.CD010072.pub2 PMID:24825181 Harvard Health. (2018). Making sense of cholesterol tests. Retrieved from https://www.health.harvard. edu/heart-health/making-sense-of-cholesterol-tests Harvard Health. (2018). Uncovering the link between emotional stress and heart disease. Retrieved from https://www.health.harvard.edu/heart-dease-overview/uncoveringthelinkbetweenemotionalstressandheart disease Healthy Living (2018). Yoga and heart health. Retrieved from Healthy living/Physical Activity/Yogaand-Heart-Health_UCM_434966_Article.jsp Holzel, B., Carmody, J., Evans, K., Hoge, E., Dusek, J., Morgan, L., ... Lazer, S. W. (2010). Stress reduction correlates with structural changes in the amygdala. Social Cognitive and Affective Neuroscience, 5(1), 11–17. doi:10.1093can/nsp034 PMID:19776221 Iglay, K., Hannachi, H., Howie, P., Xu, J., Li, X., Engel, S. S., ... Rajpathak, S. (2016). Prevalence and co-prevalence of comorbidities among patients with type 2 diabetes mellitus. Current Medical Research and Opinion, 32(7), 1243–1252. doi:10.1185/03007995.2016.1168291 PMID:26986190 Khalsa, S. B., Cohen, L., McCall, T., & Telles, S. (2016). Principles and practices of yoga in health care (1st ed.). Handspring Publishing.

257

 Yoga and Lifestyle Factors in Heart Disease

Kim, T., & Han, P. (2016). Physical exercise counteracts stress-induced upregulation of melaninconcentrating hormone in the brain and stress-induced persisting anxiety-like behaviors. Experimental Neurobiology, 25(4), 163–173. doi:10.5607/en.2016.25.4.163 PMID:27574483 Kivimäki, M., Virtanen, M., Elovainio, M., Kouvonen, A., Väänänen, A., & Vahtera, J. (2006). Work stress in the etiology of coronary heart disease a meta-analysis. Scandinavian Journal of Work, Environment & Health, 32(6), 431–442. doi:10.5271jweh.1049 PMID:17173200 Klainin-Yogas, P., Oo, W. N., Suzanne Yew, P. Y., & Lau, Y. (2015). Effects of relaxation interventions on depression and anxiety among older adults: A systematic review. Aging & Mental Health, 19(12), 1043–1055. doi:10.1080/13607863.2014.997191 PMID:25574576 Kolacz, J., & Porges, S. W. (2018). Chronic diffuse pain and functional gastrointestinal disorders after traumatic stress: Pathophysiology through a polyvagal perspective. Frontiers in Medicine, 5, 145. doi:10.3389/fmed.2018.00145 PMID:29904631 Kressin, N., Orner, N., Manze, M., Glickman, M., & Berlowitz, D. (2010). Understanding contributors to racial disparities in blood pressure control. Circulation, 3, 173–180. PMID:20233981 Leening, M., Ferket, B., Steyerberg, E. W., Kavousi, M., Deckers, J. W., Neiboer, D., ... Roos-Hesselink, J. W. (2014). Sex differences in lifetime risk and first manifestation of cardiovascular disease: Prospective population-based cohort study. British Medical Journal, 11, 349. PMID:25403476 Manchanda, S. C. (2014). Yoga—A promising technique to control cardiovascular disease. Indian Heart Association, 66(5), 487–489. doi:10.1016/j.ihj.2014.08.013 PMID:25443600 Manchanda, S. C., Narang, R., Reddy, K. S., Sachdeva, U., Prabhakaran, D., Dharmanand, S., ... Bijlani, R. (2000). Retardation of coronary atherosclerosis with yoga lifestyle intervention. The Journal of the Association of Physicians of India, 48(7), 687–694. PMID:11273502 McCarthy, M. (2015). Half of US cardiovascular deaths are due to modifiable risk factors. British Medical Journal, 350, h3539. doi:10.1136/bmj.h3539 PMID:26130566 McDermott, K. A., Rao, M. R., Nagarathna, R., Murphy, E. J., Burke, A., Nagendra, R. H., & Hecht, F. M. (2014). A yoga intervention for type 2 diabetes risk reduction: A pilot randomized controlled trial. BMC Complementary and Alternative Medicine, 14(1), 212. doi:10.1186/1472-6882-14-212 PMID:24980650 Miller, J. J., Fletcher, K., & Kabat-Zinn, J. (1995). Three-year follow-up and clinical implications of a mindfulness mediation-based stress reduction intervention in the treatment of anxiety disorders. General Hospital Psychiatry Journal, 17(3), 192–200. doi:10.1016/0163-8343(95)00025-M PMID:7649463 Moreno, G., Bruss, J., & Denburg, N. (2017). Increased perceived stress is related to decreased prefrontal cortex volumes among older adults. Journal of Experimental Neuropsychology, 39(4), 313–325. doi:1 0.1080/13803395.2016.1225006 PMID:27615373 Mosca, L., Barrett-Connor, E., & Wenger, N. K. (2011). Sex/gender differences in cardiovascular disease prevention: What a difference a decade makes. Circulation, 124(19), 2145–2154. doi:10.1161/CIRCULATIONAHA.110.968792 PMID:22064958

258

 Yoga and Lifestyle Factors in Heart Disease

Mozaffarian, D., Benjamin, E. J., Go, A. S., Arnett, D. K., Blaha, M. J., Cushman, M., ... Turner, M. B. (2016). Heart disease and stroke statistics—2016 update. A report from the American Heart Association. Circulation, 133(4), e38–e360. PMID:26673558 Narayan, K. M., Boyle, J., Thompson, T., Sorensen, S., & Williamson, D. (2003). Lifetime risk for diabetes mellitus in the United States. Journal of the American Medical Association, 290(14), 1884–1890. doi:10.1001/jama.290.14.1884 PMID:14532317 National Academy of Sciences, Institute of Medicine. (2009). Secondhand smoke exposure and cardiovascular effects: Making sense of the evidence. National Academies Press. National Institutes of Health, National Heart, Lung, and Blood Institute. (1998). Clinical guidelines on the identification, evaluation, and treatment of overweight and obesity in adults—The evidence report. Obesity Research, 6, 51S–209S. PMID:9813653 Ornish, D., Brown, S., Billing, J., Armstrong, L., Ports, T., McLanahan, D., ... Gould, K. L. (1990). Can lifestyle changes reverse coronary heart disease? The Lifestyle Heart Trial. Lancet, 336(8708), 129–133. doi:10.1016/0140-6736(90)91656-U PMID:1973470 Ornish, D., Scherwitz, L., Billings, J., Gould, L., Merritt, T., Sparler, S., ... Brand, R. J. (1998). Intensive lifestyle changes for reversal of coronary heart disease. Journal of the American Medical Association, 280(23), 2001–2007. doi:10.1001/jama.280.23.2001 PMID:9863851 Peters, A., McEwen, B., & Friston, K. (2017). Uncertainty and stress: Why it causes diseases and how it is mastered by the brain. Progress in Neurobiology, 156, 164–188. doi:10.1016/j.pneurobio.2017.05.004 PMID:28576664 Peters, S. A., Huxley, R. R., & Woodward, M. (2014). Diabetes as risk factor for incident coronary heart disease in women compared with men: A systematic review and meta-analysis of 64 cohorts including 858,507 individuals and 28,203 coronary events. Diabetologia, 57(8), 1542–1551. doi:10.100700125014-3260-6 PMID:24859435 Pischke, C., Scherwitz, L., Weidner, G., & Ornish, D. (2008). Long-term effects of lifestyle changes on well-being and cardiac variables among coronary heart disease patients. Health Psychology, 27(5), 584–592. doi:10.1037/0278-6133.27.5.584 PMID:18823185 Platt, C., Houstis, N., & Rosenzweig, A. (2015). Using exercise to measure and modify cardiac function. Cell Metabolism, 21(2), 227–236. doi:10.1016/j.cmet.2015.01.014 PMID:25651177 Poirier, P., Giles, T., Bray, G., Hong, Y., Stern, J., Pi-Sunyer, X., & Eckel, R. (2006). Council on nutrition, physical activity, and metabolism statement on obesity and heart disease from the obesity committee of the weight loss an update of the 1997 American Heart Association scientific obesity and cardiovascular disease: Pathophysiology. Circulation, 113(6), 898–918. doi:10.1161/CIRCULATIONAHA.106.171016 PMID:16380542 Pollitt, R. A., Rose, K. M., & Kaufman, J. S. (2005). Evaluating the evidence for models of life course socioeconomic factors and cardiovascular outcomes: A systematic review. BioMed Central Public Health, 5(1), 7. doi:10.1186/1471-2458-5-7 PMID:15661071

259

 Yoga and Lifestyle Factors in Heart Disease

Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74(2), 116–143. doi:10.1016/j. biopsycho.2006.06.009 PMID:17049418 Porges, S. W. (2009). The polyvagal theory: New insights into adaptive reactions of the autonomic nervous system. Cleveland Clinic Journal of Medicine, 76(Suppl_2), S86–S90. Porges, S. W. (2011). The polyvagal theory: neurophysiological foundations of emotions, attachment, communication, and self-regulation (1st ed.). W. W. Norton. Porges, S. W. (2017). Vagal pathways: Portals to compassion. In E. M. Seppälä, E. Simon-Thomas, S. L. Brown, M. C. Worline, C. D. Cameron, & J. Doty (Eds.), The Oxford handbook of compassion science (pp. 189–204). Oxford University Press. Porges, S. W., & Carter, C. S. (2017). Polyvagal theory and the social engagement system. In P. L. Gerbarg, P. R. Muskin, & R. P. Brown (Eds.), Complementary and integrative treatments in psychiatric practice (pp. 221–240). American Psychiatric Association Publishing. Rapsomaniki, E., Timmis, A., George, J., Pujades-Rodriguez, M., Shah, A. D., Denaxas, S., White, I. R., Caulfield, M. J., Deanfield, J. E., Smeeth, L., Williams, B., Hingorani, A., & Hemingway, H. (2014). Blood pressure and incidence of twelve cardiovascular diseases:lifetime risks, healthy life-years lost, and age-specific associations in 1.25 million people. Lancet, 383(9932), 1899–1911. doi:10.1016/S01406736(14)60685-1 PMID:24881994 Rapsomaniki, G., Pujades-Rodriguez, M., Shah, A. D., Denaxas, S., Herrett, E., Smeeth, L., ... Hemingway, H. (2015). How does cardiovascular disease first present in women and men? Incidence of 12 cardiovascular diseases in a contemporary cohort of 1,937,360 people. Circulation, 132, 1320–1328. doi:10.1161/CIRCULATIONAHA.114.013797 PMID:26330414 Raveendran, A. V., Deshpandae, A., & Joshi, S. R. (2018). Therapeutic role of yoga in type 2 diabetes. Endocrinology and Metabolism (Seoul, Korea), 33(3), 307–317. doi:10.3803/EnM.2018.33.3.307 PMID:30112866 Rosen, R. K., Thind, H., Jennings, E., Guthrie, K. M., Williams, D. M., & Bock, B. C. (2016). “Smoking does notgo with yoga:” A qualitative study of women’s phenomenological perceptions during yoga and smoking cessation. International Journal of Yoga Therapy, 26(1), 33–41. doi:10.17761/1531-205426.1.33 PMID:27797659 Sarvottam, K., Magan, D., Yadav, R. K., Mehta, N., & Mahapatra, S. (2013). Adiponectin, interleukin-6, and cardiovascular disease risk factors are modified by a short-term yoga-based lifestyle intervention in overweight and obese men. Journal of Alternative and Complementary Medicine (New York, N.Y.), 19(5), 397–402. doi:10.1089/acm.2012.0086 PMID:23210469 Shantakumari, N., Sequeira, S., & El Deeb, R. (2013). Effects of a yoga intervention on lipid profiles of diabetes patients with dyslipidemia. Indian Heart Journal, 65(2), 127–131. doi:10.1016/j.ihj.2013.02.010 PMID:23647889

260

 Yoga and Lifestyle Factors in Heart Disease

Shibata, Y., Hayasaka, S., Yamada, T., Ojima, T., Ishikawa, S., Kayaba, K., Gotoh, T., & Nakamura, Y.The JMS Cohort Study Group. (2011). Physical activity and risk of fatal or non-fatal cardiovascular disease among CVD survivors. Circulation Journal, 75(6), 1368–1372. doi:10.1253/circj.CJ-10-0970 PMID:21498910 Silberman, A., Banthia, R., Estay, I. S., Kemp, C., Studley, J., Hareras, D., & Ornish, D. (2010). The effectiveness and efficacy of an intensive cardiac rehabilitation program in 24 sites. American Journal of Health Promotion, 24(4), 260–266. doi:10.4278/ajhp.24.4.arb PMID:20232608 Steptoe, A., Hamr, M., & Chida, Y. (2007). The effects of acute psychological stress on circulating inflammatory factors in humans: A review and meta-analysis. Brain Behavior and Immunity Journal, 21(7), 901–912. doi:10.1016/j.bbi.2007.03.011 PMID:17475444 Su, S. F., Chang, M., & He, C.-P. (2018). Social support, unstable angina, and stroke as predictors of depression in patients with coronary heart disease. The Journal of Cardiovascular Nursing, 33(2), 179–186. doi:10.1097/JCN.0000000000000419 PMID:28489724 Sullivan, M., Erb, M., Schmalzl, L., Moonaz, S., Taylor, J., & Porges, S. (2018). Yoga therapy and polyvagal theory: The convergence of traditional wisdom and contemporary neuroscience for self-regulation and resilience. Frontiers in Human Neuroscience Journal, 12, 67. doi:10.3389/fnhum.2018.00067 PMID:29535617 Tafet, G., & Nemeroff, C. B. (2016). The links between stress and depression: Psycho neuroendrocrinilogical, genetic, and environmental interactions. The Journal of Neuropsychiatry and Clinical Neurosciences, 28(2), 77–88. doi:10.1176/appi.neuropsych.15030053 PMID:26548654 Tawakol, A., Ishai, A., Takx, R., Figueroa, A., Ali, A., Kaiser, Y., Truong, Q. A., Solomon, C. J. E., Calcagno, C., Mani, V., Tang, C. Y., Mulder, W. J. M., Murrough, J. W., Hoffmann, U., Nahrendorf, M., Shin, L. M., Fayad, Z. A., & Pitman, R. K. (2017). Relation between resting amygdalar activity & cardiovascular events: A longitudinal and cohort study. Lancet, 389(10071), 834–845. doi:10.1016/ S0140-6736(16)31714-7 PMID:28088338 Telles, S., Sharma, S. K., Kala, N., Pal, S., Gupta, R. K., & Balkrishna, A. (2018). Twelve weeks of yoga or nutritional advice for centrally obese adult females. Frontiers in Endocrinology, 9, 466. doi:10.3389/ fendo.2018.00466 PMID:30174651 Thompson, P., Franklin, B., Balady, G., Blair, S. N., Corrado, D., Estes, N. A. III, Fulton, J. E., Gordon, N. F., Haskell, W. L., Link, M. S., Maron, B. J., Mittleman, M. A., Pelliccia, A., Wenger, N. K., Willich, S. N., & Costa, F. (2007). Exercise and acute cardiovascular events placing the risks into perspective: A scientific statement from the American Heart Association council on nutrition, physical activity, and metabolism and the council on clinical cardiology. Circulation, 115(17), 2358–2368. doi:10.1161/ CIRCULATIONAHA.107.181485 PMID:17468391 U.S. Department of Health and Human Services. (2010). A report of the surgeon general. How tobacco smoke causes disease. What is means to you? Author.

261

 Yoga and Lifestyle Factors in Heart Disease

U.S. Department of Health and Human Services. Centers for Disease Control and Prevention. (2014). The health consequences of smoking – 50 years of progress: A report of the Surgeon General. National Center for Chronic Disease Prevention and Health Promotion, Office on Smoking and Health. Wang, Y., & Xu, D. (2017). Effects of aerobic exercise on lipids and lipoproteins. Lipids in Health and Disease, 16(1), 132. doi:10.118612944-017-0515-5 PMID:28679436 Web, M. D. (2018a). Cholesterol Management. Retrieved from https://www.webmd.com/cholesterolmanagement/guide/understanding-numbers Web, M. D. (2018b). Alcohol and CVD. Retrieved from https://www.webmd.com/heart-disease/guide/ heart-disease-alcohol-your-heart Whelton, P. K., Carey, R. M., Aronow, W. S., Casey, D. E. Jr, Collins, K. J., Dennisson Himmelfarb, C., ... Wright, J. T. Jr. (2017). Guidelines for the prevention, detection, evaluation and management of high blood pressure in adults: A report of the American College of Cardiology/American Heart Association Task Force in Clinic Practice Guidelines. Hypertension, 71(6), 1269–1324. doi:10.1161/ HYP.0000000000000066 PMID:29133354 World Health Organization. (2000). Obesity: Preventing and managing the global epidemic. WHO Technical report series No. 894. Geneva: WHO. Yadav, R., Yadav, R. K., Pandey, R. M., & Kochar, K. P. (2016). Effect of a short-term yoga-based lifestyle intervention on health-related quality of life in overweight and obese subjects. Journal of Alternative and Complementary Medicine (New York, N.Y.), 22(6), 443–449. doi:10.1089/acm.2015.0268 PMID:27136198

262

263

Chapter 15

Advances in Understanding the Use of Yoga as Therapy in Lymphedema S. R. Narahari Institute of Applied Dermatology, Kasaragod, India Madhur Guruprasad Aggithaya Institute of Applied Dermatology, Kasaragod, India Terence J. Ryan Department of Dermatology, Churchill Hospital, UK

ABSTRACT Lymphedema may be caused by dysfunction of the lymphatic system due to damage, block, hypoplasia due to genetic causes or by lymph overload. Lymphatic Filariasis is most prevalent and among the leading causes of disability. This chapter describes the improvements in three yoga protocols of integrative treatment, for lower limb, upper limb, and genital lymphedema. There are two sessions of yoga in our treatment protocol. Yoga helps to drain lymph through various mechanisms. Asanas focus on the dermal stretch, joint movement, muscle pumps, and muscle stretch and pranayamas on lung expansion. Joint movements and muscle contractions are designed to mimic nodal drainage. The yoga protocol also provides knee strengthening, gait correction, shoulder joint strengthening. Yoga asanas for comorbidities like hypertension, cardiovascular diseases, and arthritis and joint surgeries with movement restrictions are customized in the revised protocol. Yoga is an effective treatment in lymphedema and considerably improves the patient’s quality of life.

DOI: 10.4018/978-1-7998-3254-6.ch015

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

INTRODUCTION Lymphedema is a disfiguring and disabling, often neglected disease caused by dysfunction of the lymphatic system due to damage/block, hypoplasia due to genetic causes or by lymph overload. Breast Cancer Related Lymphedema (BCRL), elephantiasis due to lymphatic Filariasis (LF) and primary lymphedema of various genetic aberrations are the most prevalent world over. There are two standard yoga protocols for lower limb and upper limb lymphedema described by us previously (Narahari et al., 2007; Narahari et al., 2013). This chapter describes the changes done in yoga protocols incorporating lymphedema patients’ feedback and expert discussions. Commonly seen mild oedema of over three months duration, which pits when compressed with relatively light pressure represents lymphatic overload could be the result of a failing heart or other systemic illness unrelated to lymphatic block. Some people develop persistent oedema which becomes organized, brawny and non-pitting at least in certain parts of affected area termed lymphedema or elephantiasis, usually found in the tropics. It is acquired following mosquito bites in childhood, transmitting the parasites, Wucheraria bancrofti, Brugia malayi or Brugia timori. Also, it appears some years after therapies for cancer such as surgical removal of lymph nodes with supporting fat pads in adults or following radiotherapy. Several genetic disorders are accounting for rarer presentations, especially but not inevitably in childhood. Lymphedema occurs when lymph vessel function gets impaired due to blockade, gross dilatation or loss. Accumulation of protein-rich interstitial fluid (lymph) stimulating tissue overgrowth and swollen limbs is the hallmark of failure of lymph drainage. There is also the phenomenon of lymphatic overload due to an excess of fluid flow from the capillary bed of the upper dermis as a result of inflammation by cytokines and other inflammatory mediators from the epidermis, dermal mast cells and macrophages. A primary underlying mechanism is the failure epidermal barrier function, which also loads the dermis with inflammatory mediators such as cytokines and growth factors (Ryan, 2016) in an attempt to repair itself. Lymphedema caused by impaired lymphatic flow and the effects of dermal inflammation much worse if there is also venous overload due to the gravitational impacts of an immobile and dependent limb. The role of an overloaded venous system in aggravating lymphatic overload is alleviated by anti-gravitational interventions such as elevation, and ankle movements activating the calf muscle contractions and compressing the calf veins. There are many causes for the failure of lymph drainage and all lead to lymphedema. But in India, the major cause of lymphedema, and indeed throughout the tropics is LF. It is a neglected disease of the poor, prevalent in 78 countries. In India, there are at least 20 million people with signs and symptoms of LF. The mosquito inoculates the filarial parasites as a microscopic worm. After a prolonged circulation in the bloodstream, the parasite resides as a much more giant ‘worm’ in the lymphatic collecting ducts near their lymph node termination causing dilatation of the vessel (Young et al., 1976) and failure of flow (Witte et al., 1993). Three drugs given once yearly over five years to remove this cause circulating in the blood and consequent transmission. They are albendazole, ivermectin and diethyl carbamazine. Once the elephantiasis has developed, there is little to no reduction in lymphedema in response to these drugs alone. Similarly, another common cause especially identified in Ethiopia in Africa is Podoconiosis (Davey et al., 2007) due to not wearing footwear in an irritant soil. WHO has classified both diseases as Neglected Tropical Diseases (NTD) of the poor because until recently, there are no specific treatments everywhere available for their morbidity control. Surgical treatments for lymphedema have advanced to include repair of lymphatic function by anastomoses to small veins and other technologies, but these interventions are expensive. Also, sometimes very useful, failure cannot be guaranteed, and 264

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

long term follow-up studies in India are lacking. The only public health intervention widely offered has been advice on how to wash the limbs. The poor in rural areas are the most affected by the disease (Geyer et al., 2008). Reaching them with allopathic surgical or compression therapies is expensive for Governments. Hence the cocktail treatment consisting of phanta (a herbalized infusion) soaking, treatment for bacterial entry point (BEP) using modern dermatology drugs, Indian Manual Lymph Drainage (IMLD) and compression bandaging (Narahari et al., 2007) is now approved for nationwide rollout through AYUSH Health & Wellness Centres established under Ayushman Bharat Cell, the Ministry of AYUSH, Government of India to treat lymphedema in villages. A novel approach using Ayurveda and yoga as a self-care locally available treatment has produced statistically significant results in endemic villages of South India (Narahari et al., 2013) (see Figure 1). Such a combination was initially proposed in the Filarial Journals by Vaqas and Ryan (Vaqas & Ryan, 2003) and was taken up by the Institute of Applied Dermatology in Kasaragod, Kerala (IAD). Figure 1. Volume changes observed in lymphedematous limb after ten months of integrative treatment. Integrative treatment includes two sessions of yoga, Ayurvedic massages and compression bandaging, to be practised daily.

In this chapter, we focus mainly on the use of yoga asana (posture) and its adaptation for various forms of lymphedema. Yoga exercises, including deep breathing techniques, replace physiotherapy, long promoted in Europe. Also, instead of biomedical emollients, traditional ayurvedic local skin treatment measures were used.

265

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

YOGA PROTOCOL FOR ALL LYMPHEDEMAS Two sessions of yoga are incorporated into the integrative treatment protocol, pre-and post IMLD (Indian Manual Lymph Drainage). IMLD is an ayurveda oil massage comprising of two parts. This session includes folding exercises, breathing and stretching exercises. During the Pre-IMLD yoga compression bandages are not applied. Dermal stretch, attained through stretching exercises, evacuates lymph from dermal lymphatics from the interstitium to collecting ducts and then to larger lymphatics. Apart from the dermal stretch, joint movements and muscle pumps are activated during this session and also facilitate lymph drainage. The leg lift in prasruthahastapada asana (Prasrutha = elevated; Hasta = hand; Pada = leg), forced hamstring action in bheka asana (Bheka = frog), strengthens the knee. Post-IMLD yoga done along with compression bandaging wrapped around the lymphedematous leg, at least 3 hours after Pre-IMLD yoga. Asanas during this session focus on ankle exercises and muscle stretch, as in paschimothana asana (back stretching) and bhujanga asana (Bhujanga = snake). The toe touch, attained through forward bending and hamstring stretch in paschimothana asana, quadriceps stretch in bheka asana is beneficial to stretch the muscles after the exercises. Yoga is practiced either on an empty stomach or three hours after a meal. We advised that yoga should be practiced in a quiet room with proper ventilation. Fan or air conditioners are to be switched off during the period. Patients wear loose garments during yoga. Patients are taught prolonged, diaphragmatic breathing during yoga. Each yoga movement is coordinated with breathing; flexion of joints with exhalation and extension of joints with inhalation. The exhalation is more prolonged than inhalation. The yoga and pranayama were learned as in Swami Satyananda Saraswati’s book, Asana Pranayama Mudra Bandha and Swami Vishnu-devananda’s Complete Illustrated Book of Yoga (Swami, 2005; Swami, 2013). A yoga therapist coaches lymphedema patients to perform yoga and decides whether the patient has to practice alternative yoga positions (AYP), or regular practising can be achieved. Each yoga posture takes around 5 minutes to attain samasthithi (initial position) to sthithi (final position) and vice versa. Each session completes within 45 minutes. The room where yoga is practised should be clean and with sufficient ventilation. The patients should be calm minded, with comfortable, loose fit inner garments. The yoga should not be done when the patient has inflammatory episodes or other acute ailments, causing discomfort. Therapists should understand that yoga should be a comfortable intervention to attain its maximum benefits. During the initial days of integrative treatment, yoga exercises were taught singly at supervised phase. However, it is easy to educate the yoga exercises, if it is practised as a small group of 3-5 patients at a stretch. Now we recommend teaching lymphedema yoga to a group of elephantiasis patients of similar grade and physical ability. Two positions are recommended for yoga postures; samasthithi, the starting position and sthithi, final position. Patients start the yoga from samasthithi, and move slowly to sthithi with breathing coordination and then vice versa. We recommend remaining in the final position for 6- 7 breaths (approximately 30 seconds to one minute) depending on the yoga posture. For example, a patient has to remain in a front bending position, with the forehead touching the knee for 10 breaths in thada asana 2 (Thada = mountain). The initial position in standing yoga postures is, standing erect with feet parallel to each other and touching medially, placing the hands facing downwards with stretched fingers. In sitting postures, patients sit with both legs stretched forward and heels together, placing the palms on the floor by the side of the buttocks. The spine, neck and head should be straight and in the same line. The initial position for lying posture is supine, keeping both hands parallel to the body. The neck has to be in a slightly flexed position. The eyesight (drusti) while doing yoga should be straight ahead, or eyes should be closed. The 266

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

final position is dependent on the yoga posture and it depends on the flexibility of the patient’s body. The final position of Prasruthahasthapada asana is straight leg rising to 90 degrees, while the patient is lying in the supine position and both hands should be kept parallel to the body. There are three standard positions of fingers and hands [replica or symbol (Mudra)] used while doing the yoga postures. They are chinmudra (symbol of good), mrugi mudra (symbol of an animal) and namaskara mudra (symbol of salutation). Chinmudra is used while doing swasthika asana (Swasthika = auspicious), ujjayi (the intentional variations in breathing), anuloma-viloma (alternate nostril breathing) (for the left hand). It is performed by keeping the thumb and index finger flexed and joined together while other fingers are straight. Both the palms rest on the folded knee facing upwards. Mrugi mudra is used to close the nostrils for selective inhalation or exhalation during anuloma-viloma, suryabhedana (passing through suryanadi (right nostril)), bhasthrika (bellows breath). Here, the right forefinger and middle finger are clenched, while the thumb and other two fingers are straight. The right forefinger and middle finger are used to hold the left nostril while the right thumb is used to hold the right nostril. Namaskara mudra is performed in thada asana by keeping both the palms and fingers touching one another. The yoga postures were the part of IAD’s integrative treatment protocol for lymphedema and were employed to achieve lymph drainage. Lymph drainage through collecting vessels usually is produced by contraction of the lymphatic vessel wall, active and passive body movements, results in the stretch of skeletal muscles, dermal stretch, pulsation of arteries adjacent to deep lymphatics and compression of tissues for tissue hypertrophy. With the control of breathing, while performing asana and pranayama, both vena cava entering the heart will empty. Along with emptying of the venous system, the contents of the lymphatic system also empty into the blood vascular system but to a lesser extent through natural lymphatic venular anastomoses in the region of lymph nodes. The largest lymph trunk, the thoracic duct, which drains into the systemic (blood) circulation at the junction of the left subclavian and internal jugular veins, at the commencement of the brachiocephalic vein, collects lymph from the lower half of the body, the deep layers of the back and three-body quadrants including the left half of head and neck and left upper limb (known as left drainage area). The right lymphatic duct drains lymph from the right upper quadrant of the body, whereas the thoracic duct drains lymph from the other parts and finally joins with the left subclavian vein. The initial three postures thada asana 1, thada asana 2 and thrikona asana (Thrikona = triangle) allow a dermal stretch of the whole body and movements of all joints, thus activating joint pumps and muscle pumps. Lymph drains into the venous system when intra-thoracic pressure decreases in inspiration, whereas expiration allows the flow of lymph from extremities. Major lymphatics’ failure leads to whole-body lymphedema and local venous or lymphatic system failure results in a local overload of the drainage system and local lymphedema. Involvement of the venous system in lymphedema has long been recognized (Narahari et al., 2007). Yoga replaces central Manual Lymphatic Drainage (MLD) during the treatment of lower and upper extremities lymphedema used in Europe (Foldi et al., 2007). According to the current practice of lymphedema treatment, central MLD is a prerequisite for achieving the peripheral LD because peripheral lymphatics cannot flow efficiently into overfilled central lymphatics. The underlying mechanism may be equivalent to impact therapy (a force such as induced by massage meeting resistance causing lateral pressure at right angles to expand the tissues in that direction) (Robson, 1980). Lymph drainage from lymph stagnated area (lymphedematous segments of body) may be achieved by bypassing the damaged regional lymph nodes. Cancers can recur in distant sites in the absence of lymph node involvement by through circulation without traversing lymph nodes. Lymphatic pathways from the popliteal fossa (efferent vessels of nodes) pass through different routes to reach various lymph 267

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

node groups in inguinal and pelvic areas. Neo-lymphagiogenesis is observed in skin during lymphedema. Also, there is a dermal mesh of lymphatics networking throughout the integument. All these may facilitate lymph drainage in alternate route to decongest the lymphedema through yoga and massages. Apart from yoga asanas, three balancing exercises also advised during post IMLD yoga session. They are standing hip flexion, balancing over the heels and heel raising, side walking. This helps to improve the patient’s balance and steadiness.

Yoga Protocol for Lower Limb Lymphedema Table 1 lists the names and sequence of yoga asanas used in the treatment of lower and upper extremity lymphedema. There is a difference between yoga protocols for lower limb and upper limb lymphedema. Patients practice yoga in two sessions. The focus of both protocols is to achieve a dermal stretch, movements of the joints and muscle stretch coordinated with breathing. Most of the patients initially experience difficulty in yoga movements due to the immobility and massive oedema persists in extremities. So AYPs are introduced during the initial self-care training period. They were done using other supportive measures such as holding a cloth, passive exercises with the support of the unaffected limb or with the support of a therapist or a home caretaker. The passive movement of the affected limb with the help of unaffected limb in thada asana 1, thada asana 2 and thrikona asana are AYPs for upper-limb lymphedema. But if a patient with lower-limb lymphedema has difficulty to do these 3 exercises, we advise them to stand against a wall to provide support. The AYP for gomukha asana (Go = cow; Mukha = face) is, holding the cloth in both hands behind the back initially for a few days, then clasping the cotton cloth by decreasing the gap of the grasped hands to attain the final position by holding both hands together. In gomukha asana, lower limb lymphedema patients also have difficulty to achieve sthithi although upper extremities are normal. In Prasruthahasthapada asana, the support of bystander also recommended (see Figure 2). This may be due to stiff muscles as a result of inactivity and oedema in the muscles themselves. Yoga’s coordinated movements are slow and encourage patients to gradually begin body movement that was previously restricted by lymphedema and which slowly empties the muscle compartment of excess fluid. All the yoga is to be performed in the sequence of group 1 to group 4. In Foldi’s technique, the MLD treatment of an extremity starts proximally. This allows the successful mobilization of fluid from the periphery. Here yoga postures were adopted to achieve movements and pressure over the group of lymph nodes in the same pattern to simulate MLD of Foldi’s technique.

268

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Table 1. The sequence of yoga for lymphedema Sl no. 1.

Upper limb lymphedema Pre-IMLD yogai

Post-IMLD yogai

Loosening exercises

Lower limb lymphedema Pre-IMLD yogai

Post-IMLD yogai

Loosening exercises

2.

Warm-up yogaii Thada asana 1 and 2 (Thada = mountain) Thrikona asana (Thrikona = triangle)

Warm-up yoga Thada asana 1 Thrikona asana

Warm up yoga Thada asana 1 Thrikona asana

Warm up yoga Thada asana 1 Thrikona asana

3.

Pranayamas for central lymph drainage Ujjayi (Ut= up; Jayi= victory, Towards victory) Anuloma-viloma (alternate nostril breathing) Rechaka- kumbhaka (Inspiration and expiration using alternate nostrils) Suryabhedana (Passing through suryanadi (right nostril)) Kapalabhathi (Kapala = forehead; Bhati = shining)

Pranayamas for central lymph drainage Ujjayi Anuloma-viloma Rechaka- kumbhaka Suryabhedana Kapalabhathi

Pranayamas for central lymph drainage Ujjayi Anuloma-viloma Rechaka- kumbhaka Suryabhedana Kapalabhathi

Pranayamas for central lymph drainage Ujjayi Anuloma-viloma Rechaka- kumbhaka Suryabhedana Kapalabhathi

4.

Yoga to improve ROM and facilitate peripheral lymph drainage Swasthika asana (Swasthika = an auspicious sign) Vajra asana (Vajra = diamond) Gomukha asana (Go = cow; Mukha = face) Greevasanchalana (Greeva = neck; Sanchalana = movement) Mustikabandha (Mustika = fist; Bandha = tight holding) Manibandhanamana (Manibandha = wrist; Namana = bending) Manibandha chakra asana (Manibandha = wrist; Chakra = rotatory movement) Kehuni namana (Kehuni = elbow; Namana = bending) Skanda chakra asana (Skandha = shoulder; Chakra = rotatory movement) Rajjukarshana asana (Rajju = rope; Karshana = pulling) Kastatarkshna asana (Kasta = wooden log; Tarkshna = breaking using axe) Bheka asana (Bheka = frog) Makarasana (Makara = crocodile)

Yoga to improve ROM and facilitate peripheral lymph drainage Swasthika asana Vajra asana Greevasanchalana Mustikabandha Manibandhanamana Manibandha chakra asana Rajjukarshana asana Kastatarkshna asana

Asanas to facilitate peripheral lymph drainage Swasthika asana Gomuka asana Bheka asana Makara asana

Asanas for ankle joint exercises Padangulinamana (Padanguli = toes; Namana = bending) Gulphanamana (Gulpha = ankle; Namana = bending) Gulphachakra asana (Gulpha = ankle; Chakra = rotator movements) Paschimothana asana (Paschima = back; Uthana = extreme stretching) Bhujanga asana (Bhujanga = snake) Makara asana

5.

Yoga to achieve foot end elevation Prasrutha Hasthapada asana (Prasrutha = elevated; Hasta = hand; Pada = leg) Vipareethakarani (Vipareetha = opposite; Karani = body)

Yoga to achieve foot end elevation Prasrutha Hasthapada asana

Yoga to achieve foot end elevation Prasruthahasthapadasana (Prasrutha = elevated; Hasta = hand; Pada = leg)

Yoga to achieve foot end elevation Prasruthahasthapadasana (Prasrutha = elevated; Hasta = hand; Pada = leg)

6.

Relaxation Shava asana-1and 2 (Shava = cadaver)

Relaxation Shavasana-1and 2

Relaxation Shavasana-1and 2

Relaxation Shavasana-1and 2

i Pre-IMLD yoga was done prior to compression bandage, and Post-IMLD yoga was practised with compression. The exercises which need flexion of the joints were avoided in post IMLD yoga. Post-IMLD yoga was performed after 3 hours of Pre IMLD yoga.iiAsana means posture. Yoga asana is generally named after an object, animal, or naturally occurring things. The name of the yoga is in italic (Sanskrit) and its meaning in parenthesis.

269

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Figure 2. Patient with bilateral lower limb lymphedema performing Prasrutha hasta pada asana. He had difficulty in attaining final position due to huge sized limb. So, bystander gave support.

Thada asana 1 and 2, and trikona asana (see Figure 3) are used for warm-up in our protocol, since it activates all the major joints of the body, thus stimulating superficial lymph nodes. However due to the necessity of preparedness of the body to do yoga exercises, loosening exercises are later added to the protocol focusing more on the chest and abdominal muscles along with warming up of whole body, kapalabhathi (frontal brain cleansing breath) instead of ujjayi. We incorporated vajra asana (Vajra = diamond) and vipareetha karani in our earlier versions of yoga protocol. There was difficulty to sit in vajra asana due to the swelling and many patients complained of low backache after practising vipareetha karani due to the lifting of large sized limbs. So, both exercises were withdrawn in the new protocol. Yoga asanas for comorbidities like hypertension, cardio vascular diseases and arthritis and joint surgeries with movement restrictions are customized. In patients with cardio vascular diseases we avoid suryabhedana, bhasthrika and bheka asana. Surya bhedana pranayama is also avoided for hypertensive patients. Partial bending of joints while folding exercises is advised for those patients who have undergone joint surgery or suffering from knew joint osteoarthritis. The patients who are unable to sit in floor are advised to do sitting and lying yoga exercises in wooden cot.

270

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Figure 3. Thrikona asana performed by a patient with left lower limb lymphedema, in post-IMLD yoga session. This yoga session is performed with a compression bandage.

The duration of initial supervised care at IAD is 14 or 21 days, depending upon the severity of the disease.

271

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Figure 4. Patient with left upper limb lymphedema doing bheka asana. Patient is using cloth to hold the left lower limb.

Yoga Protocol for Upper-Limb Lymphedema The treatment should aim to improve the range of movements (ROM) of upper-limb joints. Gross changes in ROM occur due to inactivity as a result of oedema, partial loss of mobility and fibrosis. Studies have shown the therapeutic benefit of exercise and breathing in secondary upper limb lymphedema (Narahari et al., 2016; Moseley et al., 2005). In lymphedema, overfilled and dilated truncal collecting lymphatics constitute a significant issue to peripheral lymph flow, which results in swelling. So, the European MLD treatment of an extremity starts proximally (also known as central MLD). This allows the successful mobilization of fluid from the periphery. The stagnated lymph is directed to axillary lymph nodes of the contralateral normal limb and the inguinal lymph nodes of the same side. Thus, before initiating the treatment, these groups of lymph nodes require nodal drainage. Initial three yoga of our treatment protocol (thada asana 1, thada asana 2 and trikona asana) achieve movement of all joints of the body activating all major muscles and yielding pressure over the groups of lymph nodes in the region. The pressure differences might stimulate the lymph capillaries all over the body. Then the central lymph drainage is done through breathing exercises (pranayama). The decrease in intrathoracic pressure, while inspiration leads to drainage of lymph into the venous system and expiration, allows the flow of lymph from extremities (Vaqas & Ryan, 2003). In China, recent studies have emphasized that this can be induced by an external apparatus, a mini iron lung (Brorson, 2011). Controlled breathing along with contraction of rectus abdominus, diaphragm and intercoastal muscles as in bhasthrika creates pressure differences in both abdomen and thoracic region (Narahari et al., 2013). These pressure differences allow lymph to drain towards the thorax. Next series of 13 yoga (see Table 1) help peripheral limb movement activating lymph drainage. Swasthika asana and vajra asana apply pressure over inguinal and popliteal lymph nodes and the next eight yoga drain from axillary and anterior and lateral cervical lymph nodes. Bheka asana allows movements of all major joints and activates all major muscles (see Figure 4). Makara asana (Makara = crocodile) and shava asana 1 and 2 are relaxing yoga postures. The yoga practised in this

272

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

protocol has coordinated slow movements with breathing. The breathing allows chest expansion leading to maximizing lung capacity. So, it plays a similar role as that of central MLD of Foldi’s technique (Foldi et al., 2007; Narahari et al., 2013). Table 2. The joint movements during yoga asana and its benefits in lymphedema patients Yoga

Joint movements

Actions

Benefits

1.

Warm-up asanas Thada asana 1 and 2 Thrikona asana

Spinal column- Flexion, Extension, Lateral flexion, Reduction, Rotation Shoulder joint- Flexion, Extension, Adduction, Abduction, Inward medial rotation, Outward medial rotation Elbow joint- Flexion, Extension, Pronation, Supination Wrist joint- Flexion, Extension Hip joint- Flexion, Extension, Adduction, Abduction, Medial rotation, Lateral rotation Knee joint- Flexion, Extension, Adduction, Medial rotation, Lateral rotation Ankle joint- Plantarflexion, Dorsiflexion Tarso metatarsal joint- Flexion, Extension metatarsophalangeal joint-Flexion, Extension

2.

Pranayama 1. Ujjayi 2. Anuloma-viloma 3. Rechaka- kumbhaka 4. Suryabhedana Kapalabhathi

No importance to joint movement

Intercoastal muscles, diaphragm, rectus muscles

Improves respiration Central lymph drainage Increases oxygen consumption of the body

a) Asanas acting around ankle joint in lower limb lymphedema Padangulinamana Gulphanamana Gulphachakra asana

Ankle joint- Plantarflexion, Dorsiflexion, Inversion, Eversion Tarsometatarsal joint- Flexion, Extension Metatarsophalangeal joint- Flexion, Extension

Peroneus muscles (longus, brevis, tertius), Tibialis anterior, Tibialis posterior, Gastrocnemius, Soleus, Extension digitorumlongus, Extensor hallucislongus, Extensor digitorumbrevis, Plantaris, Flexor hallucis posterior, Flexor digitoriumlongus, Flexor digitorumbrevis, Flexor hallucisbrevis, Flexor hallucislongus, Flexor digitiminimibrevis, Flexor digitorumlongus, Quadratusplantae, Lumbricales, Interossei,

Strengthen Gastrocnemius and Soleus. Activation of smaller joints of foot and muscles especially Flexor digitoriumlongus (Padangulinamana) and Gastrocnemius, Soleus (Gulphanamana and chakra asana)

b) Asanas acting around cervical and axillary lymph node groups in upper limb lymphedema Greevasanchalana Mustikabandha Manibandhanamana Manibandha chakra Kehuni Namana Skanda chakra

Neck: Flexion, Extension/hyperextension, Lateral Flexion (abduction), Reduction, Rotation Shoulder joint: Flexion, Extension, Circumduction Elbow joint: Flexion, Extension, Pronation, Supination Wrist joint: Flexion, Extension, Circumduction Carpometacarpal joint: Flexion, Extension, Metacarpophalangeal joint: Flexion, Extension

Sternocleidomastoid, Obliquuscapitis, Longissimuscapitis, Trapezius, Deltoid, Pectoralis major, Coracobrachialis, Biceps brachii, Teres major, Triceps brachii and other muscles.

Increases joint movement Helps peripheral lymph drainage due to the activation of muscles. Nodal drainage

Asanas to facilitate peripheral lymph drainage Swasthika asana

Hip joint- Flexion, Extension, Adduction, Abduction, Medial rotation, Lateral rotation Knee joint- Flexion, Extension. Ankle joint- Plantarflexion, Dorsiflexion

Gluteus (maximus, medius, minimus), Iliopsoas, Tensor fasciae latae, Rectus femoris, Sartorius, Adductors (longus, brevis, magnus), Pectineus, Semitendiunosus, Semimembranosus, Biceps femoris, Pectineus, Gracilis, Gemellus superior and inferior, Obturatorinternus and externus, Quadratusfemoris, Piriformis

Strengthen Weak hip abductors (Trendenlennberg gait, Waddling gait)

Gomukha asana Bheka asana Paschimothana asana Bhujanga asana Makara asana

Spinal column- Flexion, Extension Neck- Flexion, Hyperextension, Abduction, Reduction Shoulder girdle- Elevation Shoulder joint- Flexion, Extension, Adduction, Inward medial rotation, Outward medial rotation Elbow joint- Flexion, Extension Wrist joint- Flexion Hip joint- Flexion, Extension, Adduction, Abduction, Medial rotation, Lateral rotation Knee joint- Flexion, Extension, Adduction, Medial rotation, Lateral rotation Ankle joint- Plantarflexion, Dorsiflexion, Inversion

All major muscles of the body

Stretching (backwards in bhujanga asana and bheka asana, anterior stretch in paschimothana asana) induces dermal stretch. This improves lymph drainage. Increased abdominal pressure in bhujanga asana and bheka asana activates rectus muscles gomukha asana and bheka asana allow chest expansion. This facilitates pressure change in thoracic regions which helps central lymph drainage. Hip and knee strengthening Corrects outer rotation of hip, anterior and lateral pelvic tilt

Knee strengthening in antalgic gait, improves weak gluteus muscles Strengthening of abdominal muscles. Allows chest expansion

--

3.

4.

All major muscles of the body Body balancing on tiptoe during thada asana 1.

Activation of muscles over the lymph node groups. Body posture correction due to the stretch. Stretching (backwards in thada asana 2, lateral in thrikona asana and upward stretch in thada asana 1) induces dermal stretch. This improves lymph drainage Improves Flat foot, Hyperextension of thumb (Hitchhiker’s Toe) and toes, Hammer toe deformity, Bunions. Knee strengthening

5.

Yoga to achieve foot end elevation Prasruthahasthapadaasana

Hip joint- Flexion, Extension Ankle joint- Plantarflexion, Dorsiflexion

Gluteus maximus, Adductors (longus, brevis, magnus), Iliopsoas, Tensor fasciae latae, Rectus femoris, Sartorius, Semitendiunosus, Semimembranosus, Biceps femoris, Gastrocnemius, Soleus, Plantaris, Tibialis posterior, Flexor hallucis posterior, Flexor digitoriumlongus, Tibialis anterior, Extension digitorumlongus, Extensor hallucislongus, Peroneus tertius

6.

Relaxation asanas 1. Shava asana-1and 2

--

--

273

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Joint movements help to drain lymph and massage of the affected areas achieve drainage from the lymph nodes due to the pressure exerted over nodes. These asanas imitating the nodal drainage (see Table 2). Superficial lymph nodes are situated around major joints. The level of pressure used in joints is different in each type of yoga. The pressure on axillary lymph nodes is not similar while comparing the sudden, jerky movements of kastatharkshna asana (Kasta = wooden log; Tarkshna = breaking using axe) with the slow, anticlockwise movement of the shoulder joint of skandhachakra asana (Skandha = shoulder; Chakra = rotatory movement). The sequences are arranged to give pressure over lymph nodes from extremities to the central area. The initial three and last four yoga postures act on the lymph nodes over both upper and lower extremities. Pain and structural changes due to increased limb size lead to restricted joint movements in upper limb lymphedema and wasting of muscles of the shoulder girdle (Brennan et al., 1998). Slow and consistent joint movements in yoga treatment protocol help to tolerate pain and to attempt free movement of upperlimb joints. The thada asana 1, thada asana 2, thrikona asana and bheka asana require movement of all joints of the body activating all major muscles. Swasthika asana and vajra asana have more action on hip and knee joints with abdominal muscle relaxation. Gomukha asana, greeva sanchalana (Greeva = neck; Sanchalana = movement), mustika bandha (Mustika = fist; Bandha = tight holding), manibandha namana (Manibandha = wrist; Namana = bending), manibandha chakra (Manibandha = wrist; Chakra = rotatory movement), rajjukarshana (Rajju = rope; Karshana = pulling) and kastatarkshna asana focus on movements of shoulder and elbow joints. The deltoid, pectoralis major, biceps brachii, triceps brachii and intercoastal muscles are activated during these asanas. Shoulder stretching is essential to improve breast cancer-related lymphedema, kastatarkshna asana, rajjukarshana helps in shoulder stretching, thus improves the shoulder strength.

Breathing Exercises and Lymphedema During inhalation, lymph from the thoracic duct and right lymphatic duct is drawn into the great veins through the thoracic duct into the blood vascular system. Although no direct clinical studies have been done in patients with lymphedema, there is evidence in animals and healthy humans to support this hypothesis (Narahari et al., 2013; Pilller et al., 2006). Positive pressure inside the lungs impedes the flow of blood into the chest and heart from peripheral veins. Thada asana 1, trikona asana (see Figure 3), gomukha asana results in chest expansion due to deep inhalation. The increased abdominal pressure by flexion of the spinal column in thada asana 2 leads to forceful exhalation. Relaxed abdominal muscles due to the flexion of the hip in swasthika asana, vajra asana which focuses thoracic breathing. Deep, slow inhalation with extension or abduction and exhalation with flexion or adduction of joints in greeva sanchalana, mustika bandha, manibandha namana, manibandha chakra asana, kehuni namana (Kehuni = elbow; Namana = bending), skandhachakra asana. Generally, the length of the inhalation is more compared to exhalation. In rajjukarshana asana, there is deep inhalation due to expansion of chest followed by forceful exhalation by sudden jerky depression of shoulder girdle, whereas in kastatarkshna asana, there is deep, slow inhalation followed by forceful exhalation by sudden jerky adduction of the shoulder joint. The abdominal pressure is high in bheka asana due to extension of the spinal column and passive inward pressure of prone position; increases lung air capacity due to expansion of the chest. The parasympathetic nerve supply depends a great deal on the vagus which is situated in the left of the thorax. Yoga breathing preferentially raising left side thoracic pressure can induce typical cardiovascular parasympathetic stretch reflexes (Fan et al., 2011). Makara asana is relaxed breathing; passive inward 274

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Table 3. The gait abnormalities, its clinical characteristics and yoga responsible for gait correction No

Gait abnormalities

Characteristics of gait pattern*

Yoga which may improve the gait pattern

1.

antalgic gait

This gait pattern has a shortened swing phase of the uninvolved/less involved side due to the pain on the affected limb. So unaffected limb has to bear bodyweight most of the time while walking. This commonly results in degenerative changes in the unaffected knee.

2.

Stiff-knee gait

Difficulty in flexing the leg while stepping forward due to excessive knee contracture

Padangulinamana, Gulphanamana, Gulphachakra asana reduces spasticity

mimics scissors gait

The excessive adduction of the leg in swing, plantar flexion of the ankle and outer rotation of the hip are the characteristics of this pattern of gait. This method was to compensate for the centre of gravity while walking. However, there was no neurological involvement, the patients did not have muscle spasticity, and the arms swing was not altered.

Gomukha asana, Bheka asana, Paschimothana asana and Bhujanga asana, Swasthika asana improves posture..

Lurchgait

Suggesting weak gluteus maximus. The affected side begins to contract at the moment of heel-strike, slowing forward motion of the trunk by arresting flexion of the hip and initiating extension; the trunk lurches backwards at heel-strike on the weakened side to interrupt the forward movement of the trunk

Thada asana, Thrikona asana, Gomukha asana, Bheka asana, Paschimothana asana and Bhujanga asana, Swasthika asana improves weak gluteus muscles.

Trendelenberg gait (unilateral trunk bending)

Weak gluteus medius. This gait allows the opposite side of the pelvis to tilt downward during stance on the weak side causing both compensated and uncompensated gluteus medius weakness

Swasthika asana improves weak gluteus muscles. the posture gets corrected due to the stretching (anterior stretch in paschimothana asana, posterior in thada asana 2, bheka asana, bhujanga asana, lateral in thrikona asana and upward stretch in thada asana 1, gomukha asana)

Waddling gait

Lymphedema patients with obesity, with more oedema over thigh, has waddling gait. There will be lateral trunk movements, sometimes with hip elevation. This gait pattern is often associated with a flat foot, genu valgum. The waddling gait pattern is due to weak adductors of the lower limb.

Body balancing on tiptoe during Thada asana 1 Swasthika asana improves weak gluteus muscles. The posture gets corrected due to the stretching (anterior stretch in Paschimothana asana, posterior in Thada asana 2, Bheka asana, Bhujanga asana, lateral in thrikonaasana and upward stretch in Thada asana 1, Gomukha asana)

3.

4.

5.

6.

Prasruthahasthapada asana

* The alteration in gait is due to inactivity or less joint movements causing minimized muscle action which results in flaccid muscles. The lymphedema patients are inactive due to the discomfort caused by oedema, fear of the increase in oedema when they perform exercises or walk and occasionally due to pain.

275

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

pressure on chest and abdomen due to lying in a prone position with relaxed breathing when shava asana 1 and 2 follows it. In prasruthahasthapada asana, flexion of hip leads to higher abdominal pressure influencing forceful breathing. Bhasthrika, ujjayi, anuloma-viloma, suryabhedana, rechaka-kumbhaka (inspiration and expiration using alternate nostrils) induce physiological pressure changes in the abdomen and thoracic cavity. Deep, slow, methodical breathing, as in ujjayi and rechaka- kumbhaka causes parasympathetic nerve stimulation. Breathing of anuloma-viloma facilitates the activity of the contralateral cerebral hemisphere (Telles et al., 2012). Breathing in suryabhedana stimulates the sympathetic nervous system (Vaqas & Ryan, 2003; Telles et al., 1994). Controlled diaphragmatic breathing as in bhasthrika, creates pressure differences in both abdomen and thoracic region and allow lymph to drain towards the thorax. The strokes of continued exhalation in bhasthrika are likely to achieve maximum lymphatic clearance. The strokes of continued exhalation in bhasthrika are aimed at completely emptying the thoracic cavity. If at the same time forceful movements over the abdomen from diaphragmatic and abdominal muscles cause maximum lymphatic emptying, this might allow unimpeded peripheral drainage. Recent evidence showed that yoga techniques could be used to stimulate the sympathetic nervous system (Pascoe et al., 2015). Another detail, which in cardiovascular studies in advanced scientific laboratories has received increasing attention, is the practice of breathing through the right side of the nose in the belief that this has effects in selecting sympathetic control as opposed to the left side of the nose the vagal parasympathetic effects. The introduction of Foldi’s decongestive therapy and MLD showed how overfilled lymphatics in the limbs could be emptied by body movements (both active and passive) and using massage. It also showed that peripheral lymphatics cannot flow efficiently into overfilled central lymphatics. Thus, attempts to do so by massaging a lymphedematous leg diverts lymph into the lymphatics of the scrotum (Ryan, 2013) and is a well-known complication of intermittent pressure pumps used without first treating the central drainage system (Boris et al., 1998).

Yoga for Gait Correction in Lower Limb Lymphedema Patients The common gait abnormality is antalgic gait, due to the unequal size of the limb in unilaterally affected patients. The difficulty to walk due to the bilaterally affected lymphedema also results in antalgic gait (antalgic = anti- + alge, “against pain”). Stiff-knee gait was common in patients with excessive knee contracture. Patients with edematous swelling in the thighs showed a specific pattern which is similar to scissors gait with excessive adduction of the leg in swing, plantar flexion of the ankle and outer rotation of hip with no muscle spasticity and no alterations during arm swing. The lurch gait, Trendelenberg gait and waddling gait patterns are also common. The inactivity or less joint movements cause minimized muscle action, which results in flaccid muscles. The most affected muscles were gluteus, quadriceps femoris, gastrocnemius, soleus and flexor digitorium longus. Structural and functional abnormalities occurred due to the improper gait, which further impaired the gait pattern (Schrale et al., 2011) (see Table 3). The initial reduction of limb volume corrected the postural dysfunction during the initial two weeks in patients with waddling and lurch gait patterns. Patients reported a reduction in knee pain in two weeks. They also reported reduced frequency of analgesics consumption. The patients were able to climb stairs without difficulty after 6 months of treatment. There were no changes in structural abnormalities of the toe, but the increased range of motion (ROM) in tarso-metatarsal and metatarso-phallengeal joints were recorded. In calf muscle tightness and reduced dorsiflexion of the ankle, calf muscle stretches, attained through gulpha namana (Gulpha = ankle; Namana = bending) and paschimothana asana are recom276

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

mended (Radford et al., 2006). There are new observations on toe nail deformity. It has long been known that mechanical stresses influence nail growth. Rickshaw workers have a characteristic deformity found especially where the toe is affected by uphill exertion. Gross curvature of the greater toenail presenting as in-growing toenail can be due to loss of a flattening effect of toe compression during normal walking but often lost where there is sparing of such pressure due to impaired gait. Also, a factor of gait mechanics can be the typical ‘ski jump’ shape of all toenails so often observed in lymphedema. More gait studies are required to identify the exact pathogenesis of toenail deformity so that yoga posture corrections can be used as interventions (Aggithaya et al., 2015). Table 4. The yoga protocol for Genital lymphedema No

Yoga postures

1

Warm-up yoga Thada asana 1 Thrikona asana

2

Pranayamas Ujjayi Anuloma-viloma Rechaka- kumbhaka Suryabhedana Bhasthrika (bellows breath)

3

Asanas acting around cervical and axillary lymph node groups Greevasanchalana Skandha chakra asana Rajjukarshana asana Kastatakshana asana

4

Asanas to facilitate inguinal lymph node drainage and abdominal drainage Shashanka asana (Shashanka = deer) Manduka asana (Manduka = frog) Koorma asana (Koorma = tortoise) Ardhamatsyendra asana (Ardha = half; Matsyendra = king of fishes) Prasruthahasthapada asana

5

Relaxation Shava asana 1 Shava asana 2

Yoga for Genital Lymphedema The amount of tissue fluid in the genital tissues varies with posture and time of day. Minor degrees of scrotal oedema often resolve at night when lying flat. In genital lymphedema, the lymph is less welldrained through abdominal and inguinal lymphatics. When the lymphatics usually work, the genital swelling completely resolves (Narahari et al., 2013). There are no studies on the benefits of yoga for genital lymphedema. The integrative treatment for genital lymphedema includes two sessions of yoga (see Table 4), IMLD and compression bandaging. Initially thada asana 1and 2, trikona asana which provides flexion of hip joint, thus facilitate deep drainage. They are followed by five pranayamas as in upper and lower limb lymphedema for central lymph drainage. Then greeva sanchalana, skanda chakra asana, rajjukarshana asana and kastatarkshna asana is to be given to facilitate pressure change in the

277

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

thoracic region, a dermal stretch which helps central lymph drainage. This yoga also increases abdominal pressure, shashanka asana (shashanka = deer), manduka asana (manduka = frog), koorma asana (korma = tortoise), ardhamatsyendra asana (ardha = half; matsyendra = king of fishes) is given to attain abdominal drainage, prasruthahasthapada asana for foot end elevation and shava asana 1 and 2 for relaxation. All testicular swellings and scrotal oedema due to filariasis are not hydroceles. A true hydrocele is a clear and non-inflammatory fluid in the testicular sack. In Filariasis hydrocoele fluid is full of white cells, and even microfilaria can be observed (we believe that hydrocoele caused by LF should be renamed as filariocele because hydrocoele occurs in primary lymphedema). Noroes of Recife writes in his PhD thesis (Noroes et al., 2010) that the testicular artery is vulnerable in such inflamed tissue. We come across patients asking questions if surgery to their genital organs will affect their fertility and sexual function. As no measurements of testicular size at one year have been done, this cannot be truthfully answered. It is an argument Ryan uses to persuade further investigation of conservative (non-surgical management and especially to further research into the effects of yoga in the genital region. Noroes and many urologists believe ultrasound studies before surgery to confirm the pathogenesis of swelling. As this is not always available outside urosurgical units, the habit in India and elsewhere of encouraging peripheral community health units to risk surgery should not be regarded as a good practice. There has been no encouragement for publication of the few studies showing postoperative bleeding, infection, or failure to recognize malignant causes of genital swelling. Funding of well-equipped units able to show benefits of yoga is urgently required. The IAD approach does not aim to provide the benefits of meditation, but some mind-body effects are likely to follow the plentiful benefits of lengthy periods of skilled counselling along with the protracted and holistic questioning by nurses and therapists. IAD is not teaching spiritual enlightenment nor supreme consciousness but seeks to achieve a comfortable form of better posture together with the manipulation mentioned above of the respiratory tract and to breathe to influence the autonomic nervous system. “Any position of the body even slightly uncomfortable will result in the preoccupation of the mind with the body” (Micozzi, 2006). Thus, by using the massage and tissue mobility effects of yoga and its employment of breathing, we may permit a little discomfort, but when seeking to influence the contractility of the lymphatics through autonomic effects the question of discomfort has to be addressed. This brings in to question how does it work? The policy of the IAD is to observe the effects of integration. It is not yet able to interpret its results in terms of mechanism of action. As part of patient-centred care, some tenets of yoga practice which is inexplicable to modern science are adopted in the IAD’s integrated medicine protocol. One of these is the hand postures or mudras, developed form of posture believed to “regulate the elements” and the coordinating position of the fingers which as stated by Ramdev (Ramdev, 2010) “controls the internal glands, body parts, and their function” and “the dormant powers of the body are awakened”. We do not question this, but the patients when at home may cease to bother with these details or give up yoga altogether and careful measurement of the patients’ actual practice may in time give us the means to analyze practice and shed light on mechanisms. The Journal of Alternative and Complementary Medicine is accumulating publications on yoga which encompasses evidence-based and aspects of science’s reductionist approach. There is a range of results from ‘no significant differences’ (Ikai et al., 2014) to ‘a safe and effective means of managing symptoms’ (Rogers et al., 2015) or ‘a remarkably effective attenuation of oxidative stress’ (Lim et al., 2015) or ‘development of self-motivation’ (Martin et al., 2015) or ‘almost certainly helpful’ (Jindani et al., 2015). While yoga is part of a constellation of interventions at the IAD, the complex is useful but we do not know the specific effects of yoga. 278

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Yoga was also taught in lymphedema camps in endemic villages (Aggithaya et al., 2013). The components of camp can be patient education class using audiovisual aids and distributing Information Education Communication (IEC) materials, demonstration of skin care measures along with a session of simple yoga exercises. The patients have to actively participate in these sessions under the supervision of experts. The IEC materials should contain a stepwise description of each yoga with photos of yoga postures. Vajra asana, padanguli namana (Padanguli = toes; Namana = bending), gulphanamana, gulphachakra asana (Gulpha = ankle; Chakra = rotator movements) with breathing coordination and five types of pranayama (breathing exercises) can be demonstrated in one-day camp (see Figure 5). Yoga exercises are particularly important in this protocol because of the breathing and ankle movements. Studies have shown that the small movements of the ankle can increase the rate of lymphatic clearance (Mortimer et al., 1990). Community based one-day camps conducted at endemic districts, trained LF patients on skincare and daily yoga and breathing practices improved QoL. Figure 5. The yoga instructor teaching Gulpha namana to lymphatic filariasis patients in one day LF camp at Pallikkere, Kasaragod, Kerala, India.

In conclusion, when managing lymphedema, flow into and along the lymphatics must be encouraged. Yoga is a particular way of doing this. It also helps with reducing venous load. The use of herbs in the management of lymphedema is insufficiently understood and awaits more thorough investigation and funding. Studies of the maintenance of an epidermal barrier were first shown to be relevant in Podoconiosis in studies in Ethiopia (Ferguson et al., 2013) and have given rise to the Matts’ hypothesis that much of the inflammation observed in lymphedema is due to a repair response by the epidermis for which washing, herbal soaking, and oiling is an effective intervention. As these are diseases of the poor, fortunately, all that is most effective is low-cost self-help. Agreeing to self-help for life requires much encouragement. Therefore, time given to counselling, including yoga training, must be generous.

279

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

REFERENCES Aggithaya, M. G., Narahari, S. R., & Ryan, T. J. (2015). Yoga for correction of lymphedemas impairment of gait as an adjunct to lymphatic drainage: A pilot observational study. International Journal of Yoga, 8(1), 54–61. doi:10.4103/0973-6131.146063 PMID:25558134 Aggithaya, M. G., Narahari, S. R., Sudha, V., Shefuvan, M., Neethu, K. J., & Sushma, K. V. (2013). Self care integrative treatment demonstrated in rural community setting improves health related quality of life of lymphatic filariasis patients in endemic villages. Acta Tropica, 126(3), 198–204. doi:10.1016/j. actatropica.2013.02.022 PMID:23499714 Boris, M., Weindorf, S., & Lasinski, B. B. (1998). The risk of genital edema after external pump compression for lower limb lymphedema. Lymphology, 31, 15–20. PMID:9561507 Brennan, M. J., & Miller, L. T. (1998). Overview of treatment options and review of the current role and use of compression garments, intermittent pumps, and exercise in the management of lymphedema. Cancer, 83(S12B), 2821–2827. doi:10.1002/(SICI)1097-0142(19981215)83:12B+3.0.CO;2-G PMID:9874405 Brorson, H. (2012). Progress in Lymphology XXIII: proceedings of the 23rd international congress of lymphology. International Society of Lymphology. Davey, G., Tekola, F., & Newport, M. J. (2007). Podoconiosis: Non-infectious geochemical elephantiasis. Transactions of the Royal Society of Tropical Medicine and Hygiene, 101(12), 1175–1180. doi:10.1016/j. trstmh.2007.08.013 PMID:17976670 Fan, W. H., Ko, J. H., Lee, M. J., Xu, G., & Lee, G. S. (2011). Response of nasal airway and heart rate variability to controlled nasal breathing. European Archives of Oto-Rhino-Laryngology, 268(4), 547–553. doi:10.100700405-010-1402-0 PMID:20978779 Ferguson, J. S., Yeshanehe, W., Matts, P., Davey, G., Mortimer, P., & Fuller, C. (2013). Assessment of skin barrier function in podoconiosis: Measurement of stratum corneum hydration and transepidermal water loss. British Journal of Dermatology, 168(3), 550–554. doi:10.1111/bjd.12134 PMID:23445314 Foldi, M., Foldi, E., & Kubic, S. (2007). Textbook of Lymphology. Elsevier. Geyer, M. J., Brantus, P., & Macdonald, J. (2008). 2nd meeting of the working group on integration of wound - lymphoedema management across diseases in rural poor settings: working group report Handicap International. http://www.handicap-international.org Ikai, S., Suzuki, T., Uchida, H., Saruta, J., Tsukinoki, K., Fujii, Y., & Mimura, M. (2014). Effects of weekly one-hour Hatha yoga therapy on resilience and stress levels in patients with schizophrenia-spectrum disorders: An eight-week randomized controlled trial. Journal of Alternative and Complementary Medicine (New York, N.Y.), 20(11), 823–830. doi:10.1089/acm.2014.0205 PMID:25364946 Jindani, F. A., & Khalsa, G. F. S. (2015). A yoga intervention program for patients suffering from symptoms of posttraumatic stress disorder: A qualitative descriptive study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 21(7), 401–408. doi:10.1089/acm.2014.0262 PMID:26133204

280

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Lim, S. A., & Cheong, K. J. (2015). Regular yoga practice improves antioxidant status, immune function, and stress hormone releases in young healthy people: A randomized, double-blind, controlled pilot study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 21(9), 530–538. doi:10.1089/ acm.2014.0044 PMID:26181573 Martin, E. C., Dick, A. M., Scioli-Salter, E. R., & Mitchell, K. S. (2015). Impact of yoga interventions on physical activity, self efficiency and motivation in women with PTSD symptoms. Journal of Alternative and Complementary Medicine (New York, N.Y.), 21(6), 327–332. doi:10.1089/acm.2014.0389 PMID:25973554 Micozzi, M. S. (2006). Fundamentals of complimentary and integrative medicine. Saunders Elsevier. Mortimer, P. S., Simmons, R., & Rezvani, M. (1990). The measurement of skin lymph flow by isotope clearance –reliability, reproducibility, injection dynamics and the effect of massage. The Journal of Investigative Dermatology, 95(6), 677–682. doi:10.1111/1523-1747.ep12514347 PMID:2250109 Moseley, A. L., Piller, N. B., & Carati, C. J. (2005). The effect of gentle arm exercise and deep breathing on secondary arm lymphedema. Lymphology, 38, 136–145. PMID:16353491 Narahari, S. R., Aggithaya, M. G., Thernoe, L., Bose, K. S., & Ryan, T. J. (2016). Yoga protocol for treatment of breast cancer-related lymphedema. International Journal of Yoga, 9(2), 145–155. doi:10.4103/0973-6131.183713 PMID:27512322 Narahari, S. R., Bose, K. S., Aggithaya, M. G., Swamy, G. K., Ryan, T. J., Unnikrishnan, B., Washington, R. G., Rao, B. P. S., Rajagopala, S., Manjula, K., Vandana, U., Sreemol, T. A., Rojith, M., Salimani, S. Y., & Shefuvan, M. (2013). Community level morbidity control of lymphoedema using self care and integrative treatment in two lymphatic filariasis endemic districts of South India: A non randomized interventional study. Transactions of the Royal Society of Tropical Medicine and Hygiene, 107(9), 566–577. doi:10.1093/trstmh/trt054 PMID:23832181 Narahari, S. R., Ryan, T. J., & Aggithaya, M. G. (2013). How does yoga work in lymphedema? Journal of Yoga & Physical Therapy, 3, 135. Narahari, S. R., Ryan, T. J., Bose, K. S., Prasanna, K. S., & Aggithaya, G. M. (2011). Integrating modern dermatology and Ayurveda in the treatment of vitiligo and lymphedema in India. International Journal of Dermatology, 50(3), 310–334. doi:10.1111/j.1365-4632.2010.04744.x PMID:21342165 Narahari, S. R., Ryan, T. J., Mahadevan, P. E., Bose, K. S., & Prasanna, K. S. (2007). Integrated management of filarial lymphedema for rural communities. Lymphology, 40, 3–13. PMID:17539459 Norões, J., & Dreyer, G. (2010). A mechanism for chronic filarial hydrocele with implications for its surgical repair. PLoS Neglected Tropical Diseases, 4(6), e695. doi:10.1371/journal.pntd.0000695 PMID:20532225 Pascoe, M. C., & Bauer, I. E. (2015). A systematic review of randomized control trials on the effects of yoga on stress measures and mood. Journal of Psychiatric Research, 68, 270–282. doi:10.1016/j. jpsychires.2015.07.013 PMID:26228429

281

 Advances in Understanding the Use of Yoga as Therapy in Lymphedema

Piller, N. B., Craig, G., Leduc, A., & Ryan, T. J. (2006). Does breathing have an influence on lymphatic drainage? Journal of Lymphoedema, 1, 86–88. Radford, J. A., Burns, J., Buchbinder, R., Landorf, K. B., Cook, C., & Rome, K. (2006). Does stretching increase ankle dorsiflexion range of motion? A systematic review. British Journal of Sports Medicine, 40(10), 870–875. doi:10.1136/bjsm.2006.029348 PMID:16926259 Ramdev, S. (2006). Yog-its philosophy and practice. Diamond Pocket Books. Robson, H. E. (1980). Book review - J. B. Tracey impact therapy. British Journal of Sports Medicine, 14(4), 224. doi:10.1136/bjsm.14.4.224 Rogers, K. A., & MacDonald, M. (2015). Therapeutic yoga: Symptom management for multiple sclerosis. Journal of Alternative and Complementary Medicine (New York, N.Y.), 21(11), 655–659. doi:10.1089/ acm.2015.0015 PMID:26270955 Ryan, T. J. (2013). The skin as a barrier: What does it mean when it fails when lymphoedema is present. Journal of Lymphoedema, 8, 6. Ryan, T. J. (2016). Matts’ hypothesis: How simple strategies can lead to better outcomes. Journal of Lymphoedema, 11, 46–48. Schrale, G. H., & Ryan, T. J. (2011). Managing lymphoedema and fibrosis with coordinated movement. Journal of Lymphoedema, 6, 82–86. Swami, S. S. (2013). Asana pranayama mudra bandha. Yoga Publications Trust. Swami, V. (2003). Complete illustrated book of yoga. Crown Publications. Telles, S., Joshi, M., & Somvanshi, P. (2012). Yoga breathing through a particular nostril is associated with contralateral event-related potential changes. International Journal of Yoga, 5(2), 102–107. doi:10.4103/0973-6131.98220 PMID:22869992 Telles, S., Nagarathna, R., & Nagendra, H. R. (1994). Breathing through a particular nostril can alter metabolism and autonomic activities. Indian Journal of Physiology and Pharmacology, 38, 133–137. PMID:8063359 Vaqas, B., & Ryan, T. J. (2003). Lymphoedema: Pathophysiology and management in resource-poor settings-relevance for lymphatic filariasis control programmes. Filaria Journal, 2(1), 4. doi:10.1186/14752883-2-4 PMID:12685942 Witte, M. H., Jamal, S., Williams, W. H., Witte, C. L., Kumaraswami, V., McNeill, G. C., ... Panicker, T. M. (1993). Lymphatic abnormalities in human Filariasis as depicted by lymphangioscintigraphy. Archives of Internal Medicine, 153(6), 737–744. Young, A. E., & Kinmonth, J. B. (1976). Filariasis. Proceedings of the Royal Society of Medicine, 69(9), 708–709. doi:10.1177/003591577606900927 PMID:981278

282

283

Chapter 16

Yoga and Weight Management: A Narrative Review Sachin Kumar Sharma Patanjali Research Foundation, India Sushma Pal Patanjali Research Foundation, India Shirley Telles Patanjali Research Foundation, India

ABSTRACT A comprehensive lifestyle intervention that is effective, yet safe, for weight loss is recommended for weight management. Yoga is one such intervention that includes (1) increased physical activity and (2) suggestions about healthy eating and other behavioral changes. With this, there have been attempts to assess the effectiveness of yoga for weight management. The present review aimed at (1) evaluating studies assessing the effects of yoga on obesity and (2) grade them according to standard grading methods. Twenty-three studies, of which nine were RCTs, were included. The rating of RCTs was between 70.6 and 94.2, which can be considered fairly good. However the present review shows that there are fewer studies with (1) adequate sample sizes, (2) study designs, (3) long term follow up, and (4) adequate reporting of adverse events to conclude that yoga is an effective and safe intervention for weight loss. Hence, further studies with the points described above are required to conclude the safety and efficacy of yoga for weight loss.

INTRODUCTION In 2016, the World Health Organization estimated that there are more than 1.9 billions overweight adults globally (39 percent of the world’s adult population); of whom more than 650 million (13 percent of the world’s adult population) are obese (“World Health Organization”, 2018). It is known that overweight and obesity are associated with an increased risk of the diseases (viz. cardiovascular problems, type 2 DOI: 10.4018/978-1-7998-3254-6.ch016

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga and Weight Management

diabetes mellitus, stroke, asthma, arthritis and certain cancers) (Telles et al., 2018). Apart from this, obesity is also known to impair the several aspects of quality of life (Telles et al., 2019). Overweight and obesity places an enormous burden on economy through (i) direct cost (e.g., spending on diagnosis and treatment of medical conditions associated with obesity) and (ii) indirect costs (e.g., being overweight/or obese reduces productivity and negatively influences physical and mental well-being) (Dee et al., 2014). It is estimated that relative medical expenses of obese adults are 100 times higher than normal weight adults. Hence it is apparent that if the prevalence of obesity continues to rise, it will place a huge burden on health care sector as well as on economy. Lifestyle interventions consisting of physical activity, dietary modifications and changes in thinking or perception are considered effective for the prevention and management of obesity (Galani & Schneider, 2007). Lifestyle interventions of different durations (8-24 weeks) have been reported to reduce 5-10 percent of initial body weight (Jensen et al., 2014). These changes in body weight through lifestyle interventions are considered of clinical importance as these changes in initially body weight are associated with a decrease in metabolic and cardiovascular risk factors which have been linked with obesity (Jensen et al., 2014). However it has been reported that overweight and obese persons experience physical and psychological challenges, due to compromised physical and psychological health, which make it difficult for them to adhere to a lifestyle intervention that includes an increased level of physical activity and dietary modifications. Hence tailored behavioral interventions which are safe, yet effective, are recommended for overweight and obese persons. Yoga is a comprehensive lifestyle intervention which includes an increased level of physical activity, dietary modifications, guided relaxation and changes in thinking and perception (Telles et al., 2014). The popularity of yoga, as a weight loss intervention, is increasing rapidly. The increase in popularity is primarily due to the fact that the practices of yoga can be modified according to the need of the persons with varied conditions associated with obesity; hence the practices are considered relatively safe for overweight and obese persons. With this background the practice of yoga has been assessed for its effectiveness in the management of obesity with studies suggesting long term adherence (Yu et al., 2018). The aims of the present review were to (i) review the studies assessing the effects of yoga practices on obesity and (ii) grade them according to the level of evidence using standard grading method.

Methods We searched PubMed and Google Scholar using the search words ‘yoga, obesity’. Two authors performed the screening of the articles by title and abstract. After screening, potentially eligible articles were accessed. Twenty articles from PubMed and 3 from Google Scholar were found relevant to the present review. Original articles assessing the effects of yoga on overweight and obese adults were included in the review. Anecdotal or reported case histories and articles which were not written in English were excluded.

Methods of Grading Grading of all the studies included in the review. Two individuals independently graded all the studies which are included in the review according to the levels of evidence recommended by the United States Department of Health and Human services (https://www.ahrq.gov/). The guideline states that randomized 284

 Yoga and Weight Management

controlled trials which includes quasi randomized-process such as alternate allocation are recommended for Level-1 evidence, while a non-randomized controlled trials with predetermined eligibility criteria are considered of Level-2 evidence, observational studies with controls including case controlled studies represents the evidence of Level-3, and observational studies without control represents the evidence of Level-4. Grading of randomized controlled trials included in the review. Out of 23 studies included in the review, there were 9 randomized controlled trials. The validity of the randomized controlled trials was further assessed by two individuals using a checklist rating system based on the ‘Strength of Evidence Report from the Evidence-based Practice Center Program of the Agency for Healthcare Research and Quality (AHRQ)’ (https://www.ahrq.gov/). The checklist for RCT’s is given as Table 2 where the individual score for each parameter denotes the presence or absence of the item and not its quality. The items were (i) study question, (ii) study population, (iii) method of randomization, (iv) blinding, (v) interventions, (vi) outcomes, (vii) statistical analysis, (viii) results, (ix) discussion (including limitations and biases), and (x) source of funding.

Results Overview This review presents 23 articles which met the criteria. The studies are presented below in chronological order (from 2006 to 2019). All the studies were evaluated for their definition of the aim, the characteristics of the participants (such as age, gender and BMI), the presence of co-morbid conditions, the setting, the variables studied, the study design, blinding and masking, interventions and diet, adverse events associated with the yoga program, and limitations of each study.

Review of the Studies A 2006 study examined was carried out to assess the effect of a program on abdominal obesity in females (Ahn et al., 2006). The sample size was small (n = 20), the participants were aged 33 to 62 years, with abdominal obesity based on a waist circumference > 80 cm. The variables were relevant to the aim. The program consisted of aerobic dance and yoga for an hour, twice a week. There was no control group or any alternate intervention. There was no mention of masking or blinding. Adverse events were not mentioned. Despite the limitations the authors interpreted the program as partially effective in reducing abdominal obesity. Another study evaluated the effects of a seven day intensive residential yoga and diet program on anthropometric and biochemical variables in obese persons (Telles et al., 2010). There was a single group of 47 persons who were assessed on the first and last day with 6 days of the intervention. Participants’ ages ranged from 17 to 68 years, which was a wide range. There were 16 males in the group. All of them had a BMI more than 30 kg/m2. Obesity was not secondary to any other condition nor were there any complications in the participants selected. The variables included anthropometric assessments, body composition, postural stability, grip strength, and biochemical measures including leptin. Hence the variables were selected to evaluate the outcome of the program as well as possible mechanisms involved in any change observed. The study was a single group longitudinal trial. The absence of a control group or an alternate intervention was an obvious disadvantage of the study. There is no mention of 285

 Yoga and Weight Management

blinding or masking. The interventions consisted of both yoga and diet. The yoga program consisted of 5 hours a day as 2 sessions, for 6 days. The program emphasized voluntarily regulated yoga breathing techniques as they could be done while seated and in participants whose mobility was limited. The diet was plant-based, approximately 1965 Kcal/day. However this was based on the retrospective diet recall of 8 participants after the yoga program and may not be accurate. There is no mention of adverse event reporting. Data taken at the beginning and end of the program were compared with a t-test for paired data, which may not be considered adequately rigorous. There were beneficial changes in most of the weight related variables measured. The findings were limited by the absence of a control group, the fact that the participants had self selected to enroll for the program, the diet was not properly noted, neither was the fluid intake and there was no attempt to report adverse events or difficulties accounted by the participants in carrying out the program. Breast cancer survivors who were obese were assessed in a randomized controlled trial (Littman et al., 2012). The study aimed at evaluating yoga in this group with special emphasis on its efficacy on fatigue, quality of life and change in weight. The participants were 63 post treatment breast cancer survivors with ages between 21 and 75 years; a wide range which would include pre and post menopausal women, which is not ideal. Their BMI was ≥ 24 kg/m2 (or ≥ 23 kg/m2 if they were of Asian descent). The sample size was calculated prior to the trial and n = 60 was considered adequate to demonstrate any changes. Participants were block randomized to a yoga or to a wait-list control group, based on (i) age (strata 21- 49 years, 50-69 years, 70-75 years), (ii) stage (2 strata: 0/I grade and II/III grade) and BMI (2 strata: 24-29.9 kg/m2 and ≥ 30 kg/m2). This detailed stratified sampling helped to ensure comparability of the two groups. After stratification participants were randomized to the two interventions. The variables assessed were intended to understand the effect of the intervention on cancer as well as on obesity; and included Functional Assessment of Cancer Therapy General (FACT-G), a quality of life measure (QOL); physical activity, anthropometric measures, class attendance, home practice, knowledge of yoga and feasibility (based on time for recruitment, frequency of attendance, and participant’s satisfaction, among other factors). Questionnaires were not blind scored. Baseline data were compared with Chi-Square and t-tests. An intent-to-treat approach was used. Linear regression was used to assess changes in outcomes between groups, adjusted for baseline values. The yoga intervention was based on Vini yoga a Hatha yoga type of therapeutic yoga which includes physical stretches, postures, breath control and meditation. Yoga was to be practiced at least 5 times in a week with one 75 minutes class at the facility. There was no dietary advice. The wait list control group was asked not to practice yoga. After the 6 months follow up assessments they were given the option and facilities to practice yoga. The study was a stratified randomized controlled trial, with assessments at baseline and after 6 months. Adverse events were looked for, but none were reported. The yoga group showed better quality of life, reduced fatigue, and a decrease in waist circumference. However none of the findings were statistically significant. It is a limitation of the study that the sample size was not adequate for potentially clinically important effects to achieve statistical significance. Other limitations were that the control group did not receive additional attention which the yoga group received, hence it is difficult to understand the exact contribution of yoga practice per se, and of additional attention along with yoga. The lack of blind scoring and of objective quantitative measures are other limitations of the study. Though this study was partially conducted in a facility for cancer treatment, with the remaining sessions as home practice, there have been studies which have tried residential yoga programs in India (Telles, et al., 2010). Similarly a 5-day residential program for weight loss was conducted in the U.S. (Barun et al., 2012). There were 37 participants; the ages were between 32 and 65 years, predominantly females with the 286

 Yoga and Weight Management

BMI ≥ 25 kg/m2. The setting was residential. The aim was to assess whether yoga would influence psychological wellbeing, health behaviors and weight loss. Participants were assessed using the 52 item Health Promoting Lifestyle Profile II (HPLP), the 26 item Self Compassion Scale, a 39 item Five Facet Mindfulness Questionnaire and the Profile of Mood States, which rates 65 adjectives on a 5-point Likert scale. These scales were rather extensive. It is known that the longer the scale the less likely it is that the responses will be accurate. The only objective measure was self reported body weight at a one year follow-up. Data were assessed with Students’ t-tests, which are fairly, but not very, rigorous. The intervention was the Kripalu yoga program, which included ninety minutes of Kripalu yoga in addition to opportunities for reflection and mindfulness practice, including mindful eating. The diet provided was chiefly plant based and consisted of whole foods. The quantity was not regulated. After 5 days there were improvements in all variables assessed. After 3 months there were some improvements, though fewer were statistically significant as the number who responded at 3 months was less. The self reported body weight at 1 year was significant lower. However self-reported body weight is extremely unreliable. There would be no way of ensuring that the scale used or the method of assessment was kept constant. Other limitations of the trial include the short duration (5 days), the absence of a control or alternate intervention group, no attempt at blind scoring the data, use of lengthy questionnaires and no reliable quantitative measures to assess changes in BMI and body composition. Several studies cited above have described how yoga influences psychological, anthropometric and biochemical measures in obese persons after varying durations of weight loss (Telles et al., 2010; Littman et al., 2012; Barun et al., 2012). However sustained weight loss requires treating the causes. The causes could be medical conditions associated with weight gain (such as hypothyroidism, polycystic ovarian syndrome) or medications which alter fat metabolism and deposition (e.g., the chronic intake of steroids). If neither of these reasons are the basis for abnormal increase in weight, weight gain could be due to excessive intake of calories related to unhealthy food habits and/or reduced physical activity. The factors which influence the quality of food eaten are varied. They include ‘comfort eating’ in people who are depressed, boredom and lack of self control. A study in India attempted to evaluate whether yoga practice would increase self control in obese women (Doshi & Jogsan, 2012). The participants were a small sample of 13 females whose average age was 39.0 years and who had self elected to join a 10 day yoga program. The average weight at baseline was 76.3 kg, though the BMI was not mentioned. Self control was measured by a scale developed in India. The reliability and validity are not mentioned. There is no mention whether questionnaires were blind scored. The other measure recorded was the body weight (in kg). There is no mention of how the body weight was measured, the degree of accuracy, the machine error, and the steps taken to ensure that measurements were accurate. The yoga program was 90 minutes each day for 10 days and included stretches, sun salutations (5 rounds), postures and voluntarily regulated breathing techniques, as well as chanting. There was no attempt to report adverse events. Data were analyzed using t-tests. There was no statistically significant change in self control. The study has several limitations such as a small sample size (n = 13), absence of a control or alternate intervention group, use of a scale for which validity/reliability was not clarified, no mention of the BMI, and overall poor reporting. Several of the studies cited above have described the effect of yoga on different variables in adults, ranging from young adults to older adults in a single report (Telles et al., 2010; Barun et al., 2012; Doshi et al., 2012). One study examined obese older adults alone (Hunter et al., 2013). Fifteen older (average age 46.0±12.0 years) adults of whom three were male, all with BMI ≥ 30 kg/m2 were recruited for the study. The comparison was with fourteen young (group average age 32.0 ± 10.0 years) adults, with 287

 Yoga and Weight Management

3 males, who were lean (BMI between 18.1 and 24.9 kg/m2). The aim of the study was to determine whether yoga practice would improve glucose tolerance in the older obese persons. Glucose tolerance was assessed using a standard glucose tolerance test and by measuring the area under the curve. Twohour oral glucose tolerance tests were administered after 12 hours of fasting. Participants ingested a 15 gm glucose solution and blood samples were obtained at 30 minute intervals and analyzed. Glucose area under the curve was determined using the trapezoid method. All assessments were carried out 48 hours after the last yoga session. The design was a two group comparative longitudinal trial. It has not been mentioned how the ‘obese’ or the lean participants were recruited. Assessments included body composition via dual-energy X-ray absorptiometry. The intervention was given to both groups, that is older obese (considered high risk) and young lean (considered low risk). The yoga program was Bikram yoga, given as 90 minute. Bikram yoga classes performed 3 times per week for 8 weeks at Bikram yoga studios which were heated to 40.5oC with 40-60 percent relative humidity. The actual program consists of 26 yoga postures which follow standardized instructions. Dietary intake was not monitored. There is no mention if data were blind scored. The data were presented as means and standard error of the mean (SEM) (SEM is less desirable as measures of dispersion compared to standard deviation or confidence interval). One-way ANOVA was used to compare the baseline data of the two groups. A repeated measure ANOVA was used to compare the two groups and data at baseline and after 8 weeks. There was a significant reduction in the area under the curve in high risk (older obese persons), but did not in low risk (younger lean) persons. The findings are limited by the small sample size, the absence of a non-yoga control group, the comparison between groups which differed so much in their age and BMI. Obesity impacts different functions. A study was conducted to compare the effects of yoga and aerobics on reducing the BMI in obese persons and also to determine the effect of yoga and aerobics on pulmonary functions (Shinde et al., 2013). The participants were 60 adults between 30 and 50 years of age, both males and females who were selected randomly by the block method and assigned to two groups. To be included in the trial participants had to have grade I and grade II obesity based on the BMI by WHO criteria. At baseline their BMI and pulmonary functions were assessed. One group was given aerobic exercise (walking) for 45 minutes to an hour every day for five days in a week; the other group had a yoga program which included voluntarily regulated breathing techniques, sun salutations and yoga postures, also for 45 minutes to an hour every day for five days in a week. Data were analyzed using paired t-tests. This may not be considered adequately rigorous. Both groups were followed up monthly and re assessed after a year. The author interpreted the findings as suggestive of improved pulmonary function and weight reduction in the yoga group. The findings were limited by the facts that the two groups differed significantly at baseline with respect to their BMI as well as various pulmonary function measures (such as FVC, FEV1/FVC). Several studies cited above have assessed psychological variables in obese persons (Littman et al., 2012; Barun et al., 2012). In most cases the sample size was small. A study was conducted on 272 participants with ages between 20 and 45 years, waist circumference > 90 cm (males) or > 80 cm (females) (Dhananjai et al., 2013). Of 272 persons, 205 participants practiced yoga while the control group consisted of 67 participants who practiced aerobic exercise. Hence the number in the two groups was not equal. Also it is not clear but appears that participants self selected the interventions. The two groups were assessed for anxiety (Hamilton Anxiety Rating Scale) and depression (Hamilton rating scale for depression) at baseline and after six months. Comparisons between groups at baseline were carried out using the Chi –square and t-tests. The final comparisons were with repeated measures ANOVA. The yoga practices included stretches, yoga postures, voluntarily regulated breathing techniques and guided 288

 Yoga and Weight Management

relaxation. The aerobic exercise intervention is not detailed. The yoga program (and most probably, though not specified the aerobic exercise program) was for 60 minutes a day five days a week. There was no mention of reporting adverse events. The yoga group showed beneficial changes in BMI, waist circumference, waist-hip ratio, depression and anxiety scores. The aerobic exercise group also showed improvements. However changes in the yoga group were noted at 30 days, while the aerobic exercise group showed changes after 90 days. The findings of the present study are limited by the differences in the number in the two groups, and inadequate reporting of the interventions. However the results suggest the importance of evaluating obese persons for anxiety and depression. In most of the studies cited above overweight and/or obese persons of both genders were studied, while some studies were exclusively carried out on female participants (Littman et al., 2012). However it is known that males (and post menopausal females) are at higher risk of coronary heart disease (Franklin & Cushman, 2011). A study was conducted to assess the effect of a short term yoga-based lifestyle intervention on risk factors for cardiovascular disease in men who were overweight and obese (Sarvottam et al., 2013). Fifty one males, with an average BMI of 26.2 ± 2.4 kg/m2 were enrolled and underwent a yoga based lifestyle program for 10 days. Thirty participants completed the study. The study was a single group, prospective, lifestyle intervention study with a pre-post design. Participants were assessed at baseline and on Day 10 for changes in BMI, systolic blood pressure, interleukin-6 (IL-6), adiponectin and endothelin-1 (ET-1). The program consisted of yoga postures, voluntarily regulated breathing techniques, lectures, group discussions and individual advice. There is no mention of dietary advice. Adverse events were not looked for or reported. After 10 days there was a significant decrease in BMI, systolic BP, IL-6, and an increase in adiponectin. While the results suggest possible benefits of a short term yoga based lifestyle intervention in reducing the risk for cardiovascular disease the study has several limitations. An obvious limitation is the absence of a control group or a group given alternate intervention, the criteria for selecting subjects was not specific for high risk for cardiovascular disease, apart from the higher than normal BMI (e.g., inclusion of regular smokers). Also other factors (change in diet or in physical activity) could have contributed to the changes in the markers of inflammation and endothelial function (that is adiponectin, interleukin-6) were noted or regulated. A few of the studies cited above randomized participants to two alternate interventions (such as yoga or aerobic activity) (Littman et al., 2012; Shinde et al., 2013). Sixty eight participants with group average age ± SD, 36.4 ± 11.2 years of whom 35 were female, with BMI ≥ 25 kg/m2 (5 persons) and BMI ≥ 30kg/m2 (63 persons) were assessed after two interventions yoga and walking (Telles et al., 2014). Participants received the interventions in a residential setting. Participants were assessed for anthropometric variables, body composition, postural stability and balance, bilateral hand grip strength, serum lipid profile and serum leptin and adiponectin. The sample size was calculated based on a previous effect size and the variables measured were consistent with the aims. Participants were randomly allocated to the two interventions (i.e., yoga and walking). The yoga intervention consisted of 45 minutes of practice repeated twice a day. This included yoga postures, voluntarily regulated breathing techniques (which formed the main part of the program and was 33 minutes) and guided relaxation. At the same time of the day there were two sessions of walking. The pace was not regulated but the walking track was fixed, with a perimeter of approximately 1200 meter. On an average participants walked 2600 meters/session. Both groups had comparable activities between the sessions. The diet was regulated for both groups and served to them. It consisted of approximately 1650 Kcal/day. Data were analyzed using a repeated measures ANOVA with Bonferroni adjusted post-hoc analyses. Where required missing value analyses were performed. The authors mention that adverse events were specifically checked for, but none were 289

 Yoga and Weight Management

reported. The findings suggested beneficial effects of both programs, though there was a difference: yoga increased serum leptin whereas walking decreased adiponectin. The limitations of the findings include (i) the absence of a control group, (ii) the high drop-out rate (33 percent) attributed to difficulties in retaining participants in a residential setup with just 90 minutes of interventions every day, (iii) the intervention was for 15 days and a longer follow-up was not possible, and (iv) the adipokines, leptin and adiponectin were evaluated in a small sub-sample as these tests are expensive and the grant allocation did not provide for their estimation in all participants. In the study cited above, a yoga program which included voluntarily regulated breathing techniques, yoga postures and relaxation was compared to walking in obese persons (Telles et al., 2014). Another study compared a single yoga practice (sun salutations) to walking in females who were obese (Jakhotia et al., 2015). There were 119 participants recruited and 87 completed the study. Their ages were between 20 and 40 years of age, with a BMI between 25.1 and 34.9 kg/m2. The study was a randomized controlled study, with participants randomly allocated to four groups, that is circuit training, treadmill walking, yoga (sun salutations or surya-namaskar) and a control group. Interventions were given in a community setting. The assessments included body composition, cardio-respiratory fitness (VO2 Max), flexibility and upper and lower limb endurance. Assessments were made at baseline and after 8 weeks. Baseline and post-intervention data were compared using t-test for paired data. Between groups comparisons were carried out using baseline values as the covariate. The interventions are described in adequate detail. There is no mention of diet. Circuit training involved push ups, pull down, knee extension, leg press and other specified positions, performed as 10 repetitions. Treadmill walking was at a speed of 4 miles/hour with 0 percent grade till RPE of 7 was noted; and surya namaskar (sun salutations) was in 12 steps; this was a modified version which used a chair. The BMI reduced in all three groups {circuit training, treadmill walking and surya namaskar (sun salutations)}. There were differences in some of the other assessments. There are two main limitations of this study (i) the methods used to assess flexibility, endurance and visceral fat are provided with a reference but are not the standard methods used, and (ii) surya-namaskar (sun salutations) practice with a chair is not a conventional method. Obesity is associated with an increased risk of falls (Jeon, 2013). It was noted that obese individuals had increased postural sway while standing compared with individuals whose weights were normal (Singh et al., 2009). Two of the studies reviewed above examined postural stability in persons who were obese using a stability platform and noted improvement (Telles et al., 2014). A study was conducted specifically to investigate the effect of yoga training on static and dynamic standing balance in individuals who were obese (Jorrakate et al., 2015). There were 16 participants which is a small sample size. Those persons with a BMI of 25kg/m2 or more were and who had poor standing balance were included. Their average age was 22.2 years and there were 8 females. Participants were randomly assigned to yoga and control groups (8 in each). Assessments included standard tests for static and dynamic balance. Static balance was tested while standing on one leg with eyes closed and eyes open. Dynamic balance was based on the longest distance participants were able to reach forward. There is no mention in the manuscript, whether the assessments were carried out by an individual who was blinded to the assessments to which participants were allocated. Assessments were at baseline and after 4 weeks. The yoga program consisted of 8 consecutive yoga postures, specifically intended to improve balance and postural stability. The authors did not mention whether they observed or looked for possible adverse events during the program. The control group who were also obese, were asked to maintain their usual physical activity during the four weeks and not undertake any exercise programs. There is no mention of diet regulation. Differences between groups characteristics was determined with a t-test for unpaired 290

 Yoga and Weight Management

data. Two way repeated measures ANOVA with post-hoc analyses based on Tukey’s honestly significant difference (HSD) analysis was used. While the assessments were at baseline and at the end of the program, at 4 weeks, interim assessments were carried out of ‘one leg standing (OLS)’, and ‘functional reach test (FRT)’ with both eyes closed and eyes open. OLS improved after 2, 3 and 4 weeks of yoga; whereas FRT improved after 4 weeks of yoga. The results suggested improvement in static balance after 4 weeks of yoga. These findings were limited by factors such as the small sample size (16), the fact that participants had grade I obesity, and also that there are more extensive standardized batteries used to assess static and dynamic balance. Metabolic syndrome (MetS) refers to a collection of cardiovascular risk factors, which includes high blood pressure, central obesity, insulin resistance, and dyslipidemia (Alberti et al., 2005). A study was conducted to evaluate the cardiovascular risk factors in middle-aged and older adults with MetS following a one year yoga intervention (Siu et al., 2015). The sample size was 182 and consisted of adults with ages between 30 and 80 years (a wide range), with average age 56.0 ± 9.1 years and a diagnosis of MetS based on National Cholesterol Program (NCEP) criteria. The sample size was derived from a power analysis. The central obesity was assessed based on waist circumference alone. The yoga intervention consisted of 3 sessions weekly for year. A session was 60 minute: 10 minutes for warm-up, 40 minutes for Hatha yoga and 10 minute for regulated breathing techniques and relaxation. Data were analyzed with mixed designs. Repeated measures analysis of co-variance for quantitative data. Interim assessments were compared to pre with t-tests. The main limitation of the study was that central obesity was assessed based on waist circumference alone, and there was no active control group. The main advantage was that the follow-up was after 1 year. Diet was recorded but not regulated. Just as metabolic syndrome is associated with increased morbidity, type-2 diabetes mellitus, also is associated with several complications. Twenty four middle aged persons, with an average age of 55.4 ± 8.0 years and diagnosis of type-2 diabetes mellitus were assessed at beginning and end of a 7 day intensive residential yoga program (Tikhe et al., 2015). The sample size was derived from a power analysis. Participants were assessed using the Karada Scan body composition monitor to assess the weight, fat, (possibly from body composition), visceral fat, resting metabolism, body mass index, body age, subcutaneous whole body (this term is unclear), subcutaneous arms and subcutaneous legs (presumably this refers to the subcutaneous fat in these regions), skeletal whole body, skeletal trunk, skeletal arms and skeletal legs (presumably this refers to the lean body mass in these parts of the body). The authors have cited references mentioning that the equipment is useful and accurate for measurement of body composition. However it should be noted that most of the other measurements (e.g., resting metabolism, visceral fat) were derived based on formulae. The data were analyzed using Wilcoxon paired signed ranks test. Most assessments such as weight, fat, visceral fat, BMI, subcutaneous fat and skeletal fat reduced significantly. The yoga program included postures, breath regulation, cleansing practices, guided relaxation, and meditation. It is mentioned that participants were given a high-fiber, low fat, plant-based balanced diet. The authors do not report adverse events. Limitations of the study include the absence of a control group or an alternate intervention, the fact that all assessments were based on a single measurement method intended to assess levels of fat (subcutaneous, skeletal, visceral and total). The assessments were not specific to type 2 diabetes mellitus which was the main inclusion criteria rather than the BMI or percent of body fat. It is also worth emphasizing that the method used is questionable for its accuracy in determining body fat in different locations (total, subcutaneous, visceral), despite reports which emphasize that the bioelectric impedance (BI) method is useful. In all probability BI methods are better to assess body composition compared with complex analysis of subdivisions of body fat or lean mass. 291

 Yoga and Weight Management

The previous study (Tikhe et al., 2015) described in this review evaluated body composition in adults who were obese and had type-2 diabetes mellitus (T2DM). A study was undertaken to determine if a short term yoga based lifestyle intervention would alter the levels of interleukin (IL-6), Vitamin D, neopterin, vaspin and diabetes risk factors (Netam et al., 2015). The participants were 34 overweight/ obese persons with average age 36.8 ± 9.9 years and 13 females in the group. All participants had BMI ≥ 23kg/m2. The assessments included interleukin-6 (IL-6), 25-OH- Vitamin D and diabetes risk factors (such as obesity, fasting blood glucose, fasting insulin, insulin resistance, waist hip ratio, BP, heart rate and lipid profile). The levels of vaspin and neopterin were assessed as well. Assessments were taken at baseline, Day 10 and Day 30. Data were analyzed using repeated measures ANOVA or Friedman test as appropriate. After 10 days of intensive practice at the hospital-based center participants had a significant decrease in weight, BMI, waist-hip ratio, blood glucose and a significant improvement in lipid profile. Median fasting insulin, insulin resistance, and IL-6 decreased as well. The yoga intervention had two phases: 10 days in a non-residential center and 20 days of home practice. The program included postures, voluntarily regulated breathing techniques, lectures, dietary advice and discussions. After home practice for 20 days at day 30, weight loss remained, also the decrease in systolic BP. Maximum benefits were obtained within 10 days. The study was interesting as it assessed factors which increase the risk of type-2 diabetes mellitus, which were not assessed in other studies reviewed here, which reported the role of yoga for obesity management. The limitations of the study include the absence of a control group, inability to regulate home practice and inability to check whether participants followed the dietary suggestions given to them. Apart from this, this single group, longitudinal study does add new information to this area of study. While most of the studies above reported the effects of yoga on generalized obesity a randomized controlled trial assessed the effects of 12 weeks of yoga on women with abdominal obesity (Cramer et al., 2016). The sample size was adequate (n = 60); all the participants were females with aged between 18 and 64 years. The participants were assessed for measures of abdominal obesity (waist circumference, waist-hip ratio), BMI, body composition, blood pressure and health related quality of life; hence the assessments were relevant to the aim of the study. The participants were randomized to the two groups (i.e., yoga group and a wait list control group) using randomized block design sampling techniques. The assessors were blinded to the groups to which a participant belonged. The yoga intervention was consisted of yoga postures, breathing techniques, guided relaxation and meditation. None of the participants had low calorie diet prior to and during the study period. Adverse events were described in details. The authors concluded that the yoga had strong moderate positive effects on anthropometric and self reported variables. While the previous study described above was carried out to assess the effects of yoga on females with abdominal obesity, a subsequent randomized controlled trial was carried out to determine the effects of yoga on adult males with obesity (Rshikesan & Subramanya, 2016). The sample size was adequate (n = 72) and was calculated using a previous study. The participants were aged between 18 and 60 years. The participants were randomly allocated to the two groups using an online software. The assessors were not blinded to the group allocation. The assessments were body weight, mid arm circumference, hip circumference, skin fold thickness and psychological measures; hence the assessments were relevant to the aim of the study. The details about adverse events related to the intervention were not provided. The total duration of the yoga intervention was 14 weeks (5 days/week for 90 minutes). The yoga intervention was comprised of counseling, yoga postures, breathing techniques and meditation. Despite the

292

 Yoga and Weight Management

limitations (large age variations and blinding) the authors reported that the yoga intervention is useful to manage weight related outcomes in obese persons. A single group study was carried out to assess the effect of 10 days of yoga intervention on health related quality of life in overweight and obese persons of both sexes (Yadav et al., 2016). The study is considered of importance as overweight and obesity is known to impair several aspects of quality of life which in turn acts as a barrier to weight loss. There were 279 overweight and obese participants (210 females) of both sexes aged between 20 and 60 years; hence the sample size was adequate. All the participants were assessed for BMI, WC, HC, body composition (percentage body fat and lean mass), lipid profile and fasting blood glucose, systolic and diastolic blood pressure along with health related quality life; therefore the variables measured were relevant to the aim of the study. The study followed a single group pre-post study design. There was no information about the blinding of the assessors. The participants received an intensive yoga intervention (2 hrs /day) for 10 days. The intervention was consisted of yoga postures, breathing techniques and theory of yoga. Though the nutritional education was given to the participants; no information about the contents of the nutritional education was given. The information about the adverse events related to the intervention was not provided. The study concluded that a short-term intensive yoga intervention had a positive effect on health related quality of life. Another study was carried out to assess the effects of yoga on arterial stiffness and quality of life (Hunter et al., 2016). There were 43 participants of whom 20 were overweight and obese. The participants were aged between 18 and 70 years which is wide age range. Assessments were body composition, arterial stiffness and health related quality of life. The study used a pre-post study design where the participants were assessed before and after 8 weeks of yoga intervention. The intervention was consisted of 3 weekly classes of vikram yoga; the duration of each class was 90 minutes. Adverse events related to the intervention were not mentioned. The study was limited by the factors such as absence of a control group, poor study design and a small sample size. Despite these limitations the authors concluded that the yoga intervention is useful to reduce arterial stiffness and improves quality of life in overweight and obese persons. A randomized controlled trial was carried out to see the effects of 12 months of yoga intervention on Beta cell functions and insulin resistance in centrally obese persons with metabolic syndrome (Yu et al., 2018). Seventy nine centrally obese adults (group average age ± SD; 58 ± 8 years) with metabolic syndrome were included in the trial. The participants were assessed for (i) metabolic parameters (WC, blood pressure, fasting blood glucose, HDL-c and triglycerides), (ii) physical performance (heart rate, chair stand test, chair sit and reach test and back stretch test), and (iii) beta cell functions and insulin resistance using HOMA. The yoga intervention was for 60 minutes/day for 3 days/week during the 12 months period. The details about the blinding were not given. There was no information about the adverse event. With these limitations the authors concluded that yoga improves physical functions and reduce central obesity by alteration in Ghrelin gene and growth hormone. A comparative controlled trial was carried out to see whether 12 week of yoga or nutritional advice would positively influence the anthropometric indices which are linked to central obesity (Telles et al., 2018). Fifty two centrally obese females aged between 30 to 59 years were divided into two groups i.e., yoga and nutritional advice. All the participants were assessed for (i) anthropometry (BMI, WC, HC, SAD), (ii) lipid profile, (iii) details about their diet and physical activity and (iv) quality of life. In addition to this, the anthropometric indices (viz, ABSI, CI, BRI, VAI AVI) which have been linked to diseases associated with central obesity were derived. The yoga intervention was consisted of yoga postures, breathing techniques and guided relaxation. Participants received three yoga session/week (where each 293

 Yoga and Weight Management

was session of 75 minutes) during the 12 week period. The assessors were unaware about the details of the groups to which participants belonged to. The nutritional advice group received lectures on nutrition (one lecture/week) for 12 week period. The participants did not experience any adverse event related to the interventions. The authors concluded that both yoga and nutritional advice are useful to improve the anthropometry indices linked to disease associated with central obesity. Another study was carried out to determine the psycho-physiological changes in obese female students aged between 18 to 25 years (Kasturi & Deo, 2018). The variables studied were (i) general health questionnaire which is used to assess minor psychiatric disorder in general population, (ii) life satisfaction which is used to assess subjective well being, (iii) perceived stress scale for perception of stress, (iv) self esteem using Rosenberg self esteem scale, (v) the way the participants perceive their body shape and (vi) anthropometry variables. The study was a single group pre post study. The details about blinding were not mentioned. The intervention was consisted of a specific yoga breathing technique and yoga postures. The participants practiced yoga intervention for 80 minutes/day for 6 days/week during thirty days period under supervision of a yoga teacher. Participants were not asked for whether they experienced any adverse event related to the intervention. Despite limitations the authors concluded that the yoga practices are useful to overcome complications arises due to obesity and induces positive psychological changes in obese individuals. Psychological challenges experienced by obese person due to poor quality of life in certain domains (i.e., physical, social, etc) act as a barrier to weight loss in overweight and obese persons. A study compared the quality of life of obese persons who had experience in yoga with those who were naïve to yoga with comparable degree of obesity (Telles et al., 2019). 596 (298 yoga and 298 yoga naïve) obese persons aged between 20 and 59 years took part in the study. The participants were assessed for the six different aspects of quality (i.e, self esteem, enjoying physical activities, social satisfaction, ability to work, sexual pleasure and approach towards food) of life using the Moorehead-Ardelt Quality of Life Questionnaire. The study was a single blind comparative controlled trial where assessors were blinded to the groups. The yoga group showed a significant better quality of life in four (i.e., enjoyment in physical activities, ability to work, self-esteem, and social satisfaction) out of six aspects of quality of life based on t values of the least squares linear regression analyses, adjusted for age, gender, and BMI as covariates. The authors concluded that obese adults with yoga experience appear to have better quality of life in specific aspects, compared to yoga naïve persons with a comparable degree of obesity.

Results of the Grading Grading of all the studies included in the review. Out of twenty three studies included in the review, 9 were graded as Level-1, 13 were graded as Level-2, and 1 was graded as Level-3. No study was graded as Level-4. The level of evidence for each study is given in Table 1. Grading of randomized controlled trials included in the review. The score for each randomized controlled trial is given in Table 3.

CONCLUSION The present review evaluated 23 experimental studies which assessed the effects of yoga for obesity. Of these twenty three studies 9 were considered Level-1, 13 were considered Level-2, 1 was Level-3. Nine 294

 Yoga and Weight Management

out of the twenty three studies were randomized controlled trials (RCTs) with level of rigor ranging from 70.6 to 94.12 which is an adequate rating. However despite this rating there are inadequate studies that have adequate sample sizes, control and long term follow-up to conclude that yoga program is effective for management of obesity. Though this encouraging trends form the basis for recommending further, rigorously designed studies to evaluate yoga for long term weight regulation in obese persons.

REFERENCES Agency for Healthcare Research and Quality. (2018, 16 February). Retrieved from https://www.thecre. com/pdf/ahrq-system-strength.pdf Ahn, S. O., Lee, S. O., Kim, M. O., Lee, K. J., & Park, H. S. (2006). Effects of an abdominal obesity management program on body composition and physiological indicators in women with abdominal obesity. Korean Journal of Women Health Nursing, 12(4), 348–354. doi:10.4069/kjwhn.2006.12.4.348 Alberti, K. G., Zimmet, P., & Shaw, J. (2005). The metabolic syndrome–a new worldwide definition. Lancet, 366(9491), 1059–1062. doi:10.1016/S0140-6736(05)67402-8 PMID:16182882 Barun, T. D., Park, C. L., & Conboy, L. A. (2012). Psychological well-being, health behaviors, and weight loss among participants in a residential, kripalu yoga-based weight loss program. International Journal of Yoga Therapy, 22(1), 9–22. doi:10.17761/ijyt.22.1.y47k2658674t1212 PMID:23070668 Cramer, H., Thoms, M. S., Anheyer, D., Lauche, R., & Dobos, G. (2016). Yoga in women with abdominal obesity a randomized controlled trial. Deutsches Ärzteblatt International, 113(39), 645–652. PMID:27776622 Dee, A., Kearns, K., O’ Neill, C., Sharp, L., Staines, A., O’ Dwyer, V., Fitzgerald, S., & Perry, I. J. (2014). The direct and indirect costs of both overweight and obesity: A systematic review. BMC Research Notes, 7(1), 242. doi:10.1186/1756-0500-7-242 PMID:24739239 Dhananjai, S. (2013). Reducing psychological distress and obesity through Yoga practice. International Journal of Yoga, 6(1), 66–70. doi:10.4103/0973-6131.105949 PMID:23439736 Doshi, D. R., & Jogsan, Y. A. (2012). A study on impact of yoga on self-control in obese women. International Journal of Scientific Research Publications, 2(8), 1–3. Franklin, B. A., & Cushman, M. (2011). Recent advances in preventive cardiology and lifestyle medicine: A themed series. Circulation, 113(20), 2274–2283. doi:10.1161/CIRCULATIONAHA.110.981613 PMID:21606406 Galani, C., & Schneider, H. (2007). Prevention and treatment of obesity with lifestyle interventions: Review and meta-analysis. International Journal of Public Health, 52(6), 348–359. doi:10.100700038007-7015-8 PMID:18368998 Hunter, S. D., Dhindsa, M., Cunningham, E., Tarumi, T., Alkatan, M., & Tanaka, H. (2013). Improvements in glucose tolerance with Bikram yoga in older obese adults: A pilot study. Journal of Bodywork and Movement Therapies, 17(4), 404–407. doi:10.1016/j.jbmt.2013.01.002 PMID:24138995 295

 Yoga and Weight Management

Hunter, S. D., Dhindsa, M. S., Cunningham, E., Tarumi, T., Alkatan, M., Nualnim, N., & Tanaka, H. (2016). Impact of hot yoga on arterial stiffness and quality of life in overweight/obese adults. Journal of Physical Activity & Health, 13(12), 1360–1363. doi:10.1123/jpah.2016-0170 PMID:27633625 Jakhotia, K. A., Shimpi, A. P., Rairikar, S. A., Mhendale, P., Hatekar, R., Shyam, A., & Sancheti, P. K. (2015). Suryanamaskar: An equivalent approach towards management of physical fitness in obese females. International Journal of Yoga, 8(1), 27–36. doi:10.4103/0973-6131.146053 PMID:25558131 Jensen, M. D., Ryan, D. H., Apovian, C. M., Ard, J. D., Comuzzie, A. G., Donato, K. A., Hu, F. B., Hubbard, V. S., Jakicic, J. M., Kushner, R. F., Loria, C. M., Millen, B. E., Nonas, C. A., Pi-Sunyer, F. X., Stevens, J., Stevens, V. J., Wadden, T. A., Wolfe, B. M., & Yanovski, S. Z. (2014). 2013 AHA/ACC/TOS guideline for the management of overweight and obesity in adults: A report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines and The Obesity Society. Journal of the American College of Cardiology, 63(25), 2985–3023. doi:10.1016/j.jacc.2013.11.004 PMID:24239920 Jeon, B. J. (2013). The effects of obesity on fall efficacy in elderly people. Journal of Physical Therapy Science, 25(11), 1485–1489. doi:10.1589/jpts.25.1485 PMID:24396217 Jorrakate, C., Kongsuk, J., Pongduang, C., Sadsee, B., & Chanthorn, P. (2015). Effect of yoga training on one leg standing and functional reach tests in obese individuals with poor postural control. Journal of Physical Therapy Science, 27(1), 59–62. doi:10.1589/jpts.27.59 PMID:25642038 Kasturi, B. K., & Deo, G. (2018). Efficacy of forced right nostril breathing and selected yogasanas on female obese college students. Journal of Complementary & Integrative Medicine, 15(3). Advance online publication. doi:10.1515/jcim-2017-0070 PMID:29742064 Littman, A. J., Bertram, L. C., Ceballos, R., Ulrich, C. M., Ramaprasad, J., McGregor, B., & McTiernan, A. (2012). Randomized controlled pilot trial of yoga in overweight and obese breast cancer survivors: Effects on quality of life and anthropometric measures. Supportive Care in Cancer, 20(2), 267–277. doi:10.100700520-010-1066-8 PMID:21207071 Netam, R., Yadav, R. K., Khadgawat, R., Sarvottam, K., & Yadav, R. (2015). Interleukin-6, vitamin D & diabetes risk-factors modified by a short-term yoga-based lifestyle intervention in overweight/obese individuals. The Indian Journal of Medical Research, 141(6), 775–782. doi:10.4103/0971-5916.160698 PMID:26205020 Rshikesan, P. B., & Pailoor Subramanya, R. N. (2016). Yoga practice for reducing the male obesity and weight related psychological difficulties-A randomized controlled trial. Journal of Clinical and Diagnostic Research: JCDR, 10(11), OC22–OC28. doi:10.7860/JCDR/2016/22720.8940 PMID:28050422 Sarvottam, K., Magan, D., Yadav, R. K., Mehta, N., & Mahapatra, S. C. (2013). Adiponectin, interleukin-6, and cardiovascular disease risk factors are modified by a short-term yoga-based lifestyle intervention in overweight and obese men. Journal of Alternative and Complementary Medicine (New York, N.Y.), 19(5), 397–402. doi:10.1089/acm.2012.0086 PMID:23210469

296

 Yoga and Weight Management

Shinde, N., Shinde, K. J., Khatri, S. M., & Hande, D. (2013). A Comparative Study of Yoga and Aerobic Exercises in Obesity and its Effect on Pulmonary Function. Journal of Diabetes & Metabolism, 4(4), 1–4. doi:10.4172/2155-6156.1000257 Singh, D., Park, W., Levy, M. S., & Junq, E. S. (2009). The effects of obesity and standing time on postural sway during prolonged quiet standing. Ergonomics, 52(8), 977–986. doi:10.1080/00140130902777636 PMID:19629812 Siu, P. M., Yu, A. P., Benzie, I. F., & Woo, J. (2015). Effects of 1-year yoga on cardiovascular risk factors in middle-aged and older adults with metabolic syndrome: A randomized trial. Diabetology & Metabolic Syndrome, 7(1), 40. doi:10.118613098-015-0034-3 PMID:26000038 Telles, S., Naveen, V. K., Balkrishna, A., & Kumar, S. (2010). Short term health impact of a yoga and diet change program on obesity. Medical Science Monitor, 16(1), CR35–CR40. PMID:20037492 Telles, S., Sharma, S. K., Kala, N., Pal, S., Gupta, R. K., & Balkrishna, A. (2018). Twelve Weeks of Yoga or Nutritional Advice for Centrally Obese Adult Females. Frontiers in Endocrinology, 9, 466. doi:10.3389/fendo.2018.00466 PMID:30174651 Telles, S., Sharma, S. K., Singh, A., Kala, N., Upadhyay, V., Arya, J., & Balkrishna, A. (2019). Quality of life in yoga experienced and yoga naïve Asian Indian adults with obesity. Journal of Obesity, 2019, 9895074. doi:10.1155/2019/9895074 PMID:31183215 Telles, S., Sharma, S. K., Yadav, A., Singh, N., & Balkrsihna, A. (2014). A comparative controlled trial comparing the effects of yoga and walking for overweight and obese adults. Medical Science Monitor, 20, 894–904. doi:10.12659/MSM.889805 PMID:24878827 Tikhe, A. S., Pailoor, S., Metri, K., Ganpat, T. S., & Ramarao, N. H. (2015). Yoga: Managing overweight in mid-life T2DM. Journal of Midlife Health, 6(2), 81–84. doi:10.4103/0976-7800.158959 PMID:26167059 World Health Organization. (2018, 16 February). Obesity and overweight. Retrieved from https://www. who.int/news-room/fact-sheets/detail/obesity-and-overweight Yadav, R., Yadav, R. K., Pandey, R. M., & Kochar, K. P. (2016). Effect of a short-term yoga-based lifestyle intervention on health-related quality of life in overweight and obese subjects. Journal of Alternative and Complementary Medicine (New York, N.Y.), 22(6), 443–449. doi:10.1089/acm.2015.0268 PMID:27136198 Yu, A. P., Ugwu, F. N., Tam, B. T., Lee, P. H., Lai, C. W., Wong, C., Lam, W. W., Sheridan, S., & Siu, P. M. (2018). One year of yoga training alters ghrelin axis in centrally obese adults with metabolic syndrome. Frontiers in Physiology, 9, 1321. doi:10.3389/fphys.2018.01321 PMID:30294284

297

Yoga and Weight Management

APPENDIX 1 Table 1. Studies assessing the effects of yoga for obesity Sl. no.

1

2

3

4

5

Citation

Sample

n = 20 females Ahn et al., 2006 Age range: 33-62 years

Design Single group pre-post study design where the participants were assessed at baseline and after 12 weeks

Diet information

Variables Assessed

Results

No information available

BMI, body Fat percent, waist circumference, and waist/hip ratio, systolic blood pressure, blood glucose, total cholesterol, and triglycerides were assessed

Retrospectively 8 participants’ were checked for diet after the camp and their calorie consumption was approximately 1965 Kcal/day.

Following a 6 day residential yoga program      1. BMI decreased,      2. waist circumference decreased,      3. hip circumference Body mass index decreased, (BMI), waist and hip      4. fat-free mass decreased, circumferences, mid     5. body water decreased arm circumference, body composition,      6. total cholesterol hand grip strength, decreased, postural stability, serum      7. high density lipid profile and fasting lipoprotein (HDL) cholesterol decreased serum leptin levels.      8. fasting serum leptin levels decreased and      9. postural stability increased and      10. hand grip strength increased

Telles at al., 2010

n = 47 obese participants (16 males) BMI > 30kg/m2 Age range: 17-68 years

Single group pre-post study design where the participants were assessed at baseline and at day 6

Littman et al., 2012

n = 63 posttreatment stage 0–III borderline overweight and obese (body mass index ≥ 24 kg/ m2) breast cancer survivors Age range: 21-75 years

Random allocation of participants to two groups i.e. yoga and wait-list control No dietary advise where they were before and after 6 months of yoga (5 times/week)

Barun et al., 2012

n = 37 overweight and obese participants BMI > 25 kg/m2 Age range: 32-65 years

Participants were assessed at baseline and after five days of intensive residential No program and at follow-up after 3 months and reported weight after 1 year

Doshi et al., 2012

n = 13 obese (female) Group average age 39.0 years BMI: NOT GIVEN

Single group pre-post study design where the participants were assessed at baseline and at day 10

No dietary restriction

From abstract

Cohen’s d

Findings

The study reported that a yoga and aerobic program From abstract is helpful in the management of abdominal obesity

Level of evidence

Level 2

0.097 0.30 0.32 0.20 0.20 0.33 0.61 0.83 1.98 0.21

A 6 days residential yoga program favorably influence anthropometric Level 2 and biochemical parameters related to obesity.

BMI, waist circumference, hip circumference, quality of life, and assessment of physical activity

Those who attended at least 24 yoga classes showed      1. Better quality of life

0.83

An intensive and long-term yoga intervention in a population of breast cancer survivors improves quality of life in patients.

BMI, Health-Promoting Lifestyle Profile, The Self-Compassion Scale, A nutrition subscale, The Five Facet Mindfulness Questionnaire (FFMQ)

After 1 year of the program      1. Self reported body weight decreased After 6 days of program      1. Health promoting life-style profile scores increased      2. Level of physical activity increased      3. Spiritual growth improved      4. Stress management improved      5. Profile of mood states improved      6. Self compassion scale improved      7. Five-Facet Mindfulness Questionnaire scores increased

0.99 0.83 0.56 0.68 0.65 1.05 0.71 0.85

These findings suggest a yogabased, residential weight loss program improves Level 2 psychological well-being, alters nutrition behaviors, and helps in weight management

Body weight, and self control using a questionnaire

          After 10 days      1. Body weight decreased but not significantly      2. Improved self control but           not significantly

Level 1

A trend of reduction in body weight and improvement in Level 2 self control in obese females was observed following yoga

continued on following page 298

Yoga and Weight Management

Table 1. Continued Sl. no.

6

7

8

9

Citation

Hunter et al., 2013

Shinde et al., 2013

Sample

Design

n = 15 obese (3 males) BMI >30 kg/m2 mean age ± SD; 46.0 ± 12.0

Single group pre-post study design where the participants were assessed at baseline and after 8 weeks

n = 60 obese participants Age range: 30-50 years BMI range: 30-40 kg/ m2

n = 272 (148 males) obese participants Age range: 20-45 years Yoga: n = 207 (113 males; 92 females) Dhananjai et al., mean BMI 2013 ± SD; 31.30 ± 5.8 kg/m2 Aerobic: n = 67 (35 males; 32 females) mean BMI ± SD: 31.37 ± 6.6

n = 51 overweight and obese Sarvottam et al., men 2013 group mean BMI; 26.26 ± 2.4 kg/m2

Diet information

Participants were instructed for not to alter their usual diet

Block random allocation of participants to Not mentioned two groups i.e. aerobic exercise and yoga

A case control trial for obese patients with psychological problems where the participants No diet restriction. were assessed before and after standardized six-month protocol of Yoga classes

Single group pre-post study design where the participants were assessed at baseline and day 10

Information not available

Variables Assessed

Results

Cohen’s d

Findings

Level of evidence

Body composition, body weight, glucose tolerance

Following 8 weeks (3 times/week) of yoga      1. Body weight decreased      2. BMI decreased

0.07 0.13

The study suggested that yoga Level 2 helps to reduce body weight

BMI and Pulmonary functions

          Following yoga for 1 year (5 day/week)      1. BMI decreased           Yoga           Aerobic      2. Forced vital capacity           increased           Yoga           Aerobic      3. FEV1/FVC increased           Yoga           Aerobic      4. MVV increased           Yoga           Aerobic

0.81 0.34 0.61 0.37 0.82 0.61 0.81 0.34

Yoga reduces body weight and improves pulmonary functions in obese participants

Body weight (BW), BMI, waist circumference, hipcircumference, waisthip ratio

Following yoga and aerobic for 6 months      1. BW decreased           Aerobic           Yoga      2. BMI decreased           Aerobic           Yoga      3. WC decreased           Aerobic           Yoga      4. Hip circumference decreased           Aerobic           Yoga      5. Waist-hip ratio decreased           Aerobic           Yoga      6. Depression reduced           Aerobic           Yoga      7. Anxiety decreased           Aerobic           Yoga

0.45 0.99 0.47 0.86 0.38 1.27 0.22 0.76 0.24 0.70 1.11 2.74 0.46 0.83

Incorporating asanas (yoga postures) in the treatment protocol of obese patients suffering from Level 2 anxiety and depression is beneficial to reduce anxiety, depression and parameters related to obesity.

Body weight, BMI, Systolic BP, Adiponectin and IL-6

     1. Body weight decreased      2. BMI decreased      3. Systolic BP decreased      4. Adiponectin increased      5. IL- decreased

A short-term yogabased lifestyle intervention may be an useful to From abstract Level 2 reduce the risk for cardiovascular diseases in obese men

Level 1

continued on following page

299

Yoga and Weight Management

Table 1. Continued Sl. no.

10

11

Citation

Telles et al., 2014

Jakhotia et al., 2015

Sample

n = 68 participants (35 females) BMI ≥25 kg/m2 Age range: 20-55 years

n = 119 females Age range: 20-40 years BMI range: 25.1-34.9 kg/m2

Design

Random allocation of participants to two groups i.e. yoga group and walking group

Random allocation of participants to four groups i.e. circuit training (CT), walking on treadmill (TM), suryanamaskar (SN) and control

Diet information

Variables Assessed

Cohen’s d

Findings

Level of evidence

Following 15 days of yoga and walking      1. BMI decreased           Yoga           Walking      2. Waist circumference decreased           Yoga           Walking      3. Hip circumference decreased           Yoga           Walking      4. Lean mass decreased           Yoga           Walking      5. Body water decreased           Yoga           Walking      6. Total cholesterol decreased           Yoga           Walking      7. Triglycerides decreased           Walking      8. LDL-C decreased           Yoga      9. Leptin increased           Yoga      10. Serum adiponectin decreased           Walking

0.24 0.31 0.19 0.19 0.22 0.36 0.26 0.28 0.27 0.29 0.74 0.51 0.64 0.66 0.76 0.94

Both yoga and walking are useful to improve anthropometric and biochemical variables related to obesity. However yoga increased leptin and walking reduced adiponectin

     1. BMI decreased in following groups           CT           TM           SN      2. Body fat percentage decrease in following groups           CT           TM           SN      3. Muscle Mass increased in following groups           CT           TM           SN      4. Bone Mass increased in following groups Body composition,           TM body weight, cardio           SM respiratory fitness      5. Body water using shuttle walk and percentage increased in run test, sit & reach test following groups and muscle endurance           CT           TM           SN      6. Visceral fat rating decreased in following groups           CT           SN      7. VO2 max improved in following groups           CT           TM           SN      8. Muscle endurance improved in following groups           CT           TM           SN

0.18 0.16 0.18 0.41 0.14 0.32 0.25 0.09 0.32 0.26 0.25 0.54 0.43 0.50 0.69 0.46 1.18 0.49 0.92 1.33 1.16 1.03

All the three interventions (CN, TM, and SN) were useful for weight Level 1 management and to improve physical fitness.

Body composition, BMI, waist Participants of circumference, hip both groups were circumference, waistgiven a plant-based hip ratio, mid-arm diet regulated at circumference, hand approximately grip strength, lipid 1650 kcal/day profile, leptin and adiponectin

All participants continued their present level of dietary habits which was not assessed and reported.

Results

Level 1

continued on following page

300

Yoga and Weight Management

Table 1. Continued Sl. no.

12

13

14

15

Citation

Sample

Jorrakate et al., 2015

n = 16 (2 males 6 females in yoga and 6 males and 2 females in control) BMI of yoga group ± SD; 28.6 ± 3.5 kg/m2 BMI of control ± SD: 29.6 ± 2.8 kg/m2 Age of yoga group ± SD: 21.9 ± 1.2 years Age of control ± SD; 22.5 ± 1.5 years

Siu et al., 2015

Tikhe et al., 2015

Netam et al., 2015

n = 182 (135 females) Mean age ± SD; 56.0 ± 9.1 BMI: NOT GIVEN

n = 24 patients (6 females) Mean age ± SD; 55.38 ± 7.9 years, Mean BMI ± SD; 26.70 ± 4.7 kg/m2

n = 34 (21 males and 13 females) Mean age age ± SD; 36.79 ± 9.9 years Group mean BMI ± SD; 27.52 ± 2.7

Design

Variables Assessed

Results

Not mentioned

Static and dynamic balance using one leg standing (OLS) under the eyes open and eyes closed conditions and The functional reach test14) was used to investigate dynamic standing balance by determining the farthest distance subjects could possibly reach forward

     1. OLS with eyes open increased after           2nd week,           3rd week           and 4th week of yoga      2. OLS with eyes closed increased after           2nd week,           3rd week           and 4th week of yoga      3. Dynamic standing balance increased after 4 week of yoga

Participants were asked to adhere their daily dietary intake

Waist Circumference (WC), fasting glucose, triglycerides, and HDL-C, a battery of questionnaires including Center for Epidemiologic Studies Depression Scale (CES-D), the 12-item Short-form health survey (SF-12), and the International Physical Activity Questionnaire (IPAQ)

Following yoga for 1 year      1. WC decreased      2. Resting heart rate decreased

Weight, fat, visceral fat, resting metabolism, BMI, body age, subcutaneous whole body, subcutaneous trunk, subcutaneous arms, subcutaneous legs, skeletal whole body, skeletal trunk, skeletal arms and skeletal legs

Diet information

Cohen’s d

Findings

Level of evidence

The study reported that yoga training is beneficial to improve standing balance in obese participants who were having poor standing balance.

Level 1

0.42 0.38

These results suggest that yoga exercise improves the cardiovascular risk factors including central obesity and blood pressure in middle-aged and older adults with metabolic syndrome.

Level 1

Single group pre-post study design where the participants were assessed at baseline and after 7 days

High-fiber low-fat vegetarian balanced diet was suggested but not regulated

After 7 days      1. Body Weight decreased      2. Body Fat decreased      3. Visceral Fat decreased      4. Body mass index decreased      5. Subcutaneous trunk decrease      6. Subcutaneous legs      7. Skeletal whole body      8. Skeletal trunk      9. Skeletal arms      10. Skeletal legs

0.06 0.20 0.06 0.06 0.08 0.09 0.11 0.18 0.06 0.10

The present study suggests that 7 days of yoga intervention is useful for the management of overweight in midlife patients with T2DM.

Level 2

Single group Pre-Post study design where the participants were assessed at baseline, day 10 and day 30

At day 10      1. Body weight decreased      2. BMI decreased Lipid profile (total      3. Waist-hip ratio cholesterol, LDL, HDL decreased      4. Total cholesterol and triglycerides), decreased waist/hip ratio, IL-6, 25-OH-vitamin      5. Triglycerides decreased Supervised diet but D,      6. Heart rate decreased total Kilo-calories fasting blood glucose,      7. Fasting blood glucose is not given fasting insulin, insulin resistance,           At day 30 blood pressure,      1. Body weight decreased heart rate, and levels of neopterin      2. BMI decreased      3. Systolic blood and vaspin pressure decreased      4. Heart rate decreased      5. Waist-hip ratio      6. Fasting blood glucose

0.20 0.12 0.04 0.24 0.31 0.70 0.42 0.15 0.51 0.72 0.95 0.25 0.23

The findings of the present study showed that IL-6, vitamin D, and diabetes risk factors were favorably Level 2 influence by a short-term yogabased lifestyle intervention carried out on obese participants.

Random allocation of participants to two groups i.e. yoga and control

Random allocation of participants to two groups i.e. yoga and control

1.93 2.45 2.76 1.47 1.61 1.92 1.13

continued on following page

301

Yoga and Weight Management

Table 1. Continued Sl. no.

16

17

18.

19

302

Diet information

Variables Assessed

Results

Cramer et al., 2016

n = 60 women with abdominal obesity (WC ≥ 88 cm; BMI ≥ 25kg/m2) Age range: 18-64 years

Participants of yoga group were recommended plant-based diet, fewer industrially processed foods and a more conscious approach towards eating without reduction in their usual calories

Waist circumference, waist-hip ratio, BMI, percentage of body fat percent, percentage of body muscle mass percent, BP, health related quality of life (SF-36), self-esteem (SES), perceived stress (PSS), body awareness (BAQ) and body responsiveness (BRS) were assessed

Following 12 weeks of yoga 1. Waist-hip ratio reduced 2. Body weight reduced 3. WC reduced 4. BMI reduced 5. Percentage of body fat reduced 6. Percentage of body muscle increased 7. Physical and mental quality of life increased 8. Self-esteem increased 9. Perceived stress reduced 10. BAQ increased 11. BRS increased

0.401 0.442 0.104 0.120 0.322 0.307 0.475 0.726 0.256 0.736 0.443 0.943

Yoga can be recommended to women for the purpose of reducing abdominal obesity.

Level 1

Rishikesan et al., 2016

Following 14 weeks (5 days/week) of IAYT and physical activity 1. Body weight decreased Body weight, MAC   Yoga for both right and left           2. MACs of both upper arms, WC, HC, arms were decreased WHR, a body shape   Yoga index (ABSI), skin fold Random           3. BMI decreased thickness (SKF) at 4 allocation of   Yoga n = 72 body points applicable males (Yoga participants           4. WC and HC Both groups for male, right upper decreased n = 37, to two groups received a food log arm biceps (SKFraf)   Yoga and control Control n i.e. yoga and right upper arm triceps format and basic = 35) control           5. Percentage sample meal plan (SKFrab), sub scapular BMI: 23-35 The body fat based on guidelines (SKFshob), suprailiac 2 Kg/m   Yoga assessments of NIN (SKFstof), percentage           6. PSS decreased Age range: were taken body fat, perceived   Yoga 18-60 years before and after stress scale (PSS) intervention           7. AAQW decreased and acceptance and   Yoga action questionnaire           8. Skin fold (AAQW) in weight cumulative decreased related difficulties were   Yoga assessed           9. Suprailliac SKF at stomach from froknt side decreased   Yoga

0.112 0.284 0.265 0.155 0.184 0.458 0.414 0.455 0.268 0.613 0.605 0.453 0.654

IAYT for obese male in urban setting will be effective for obesity treatment and for reducing the obesity related problems

Level 1

Yadav et al., 2016

n = 279 male and 210 females both overweight (BMI, 23–24.9 kg/m2) and obese (BMI > 25kg/m2) Age range: 20-60 years

Following 10 days of yogabased lifestyle intervention 1. Body weight decreased 2. BMI decreased, 3. Total body fat decreased 4. Waist and hip circumference decreased 5. Systolic and diastolic blood pressure decreased 6. Total cholesterol decreased 7. Low-density lipoprotein decreased Triglycerides, and fasting glucose decreased

0.044 0.039 0.018 0.091 0.068 0.359 0.244 0.195 0.161 0.225 0.146

A short-term yogabased lifestyle intervention had a positive effect Level 2 on HRQOL in overweight and obese persons

Hunter et al., 2016

n = 43 participants (23 were normal adults BMI = 18.5-24.9 kg/m2 and 20 were overweight/ obese adults BMI ≥ 25 kg/m2) Age range: 18 -70 years

0.648 0.649 1.36

Bikram yoga ameliorates arterial stiffness in overweight/obese adults and can positively impact quality of life

Citation

Sample

Design

Block random allocation of participants to two groups i.e. a 12-week yoga (n = 40) and a waiting list (n = 20) were assessed baseline and after 12 weeks

Single group pre-post study design where participants were assessed at baseline and at day 10

Pre-post study design where participants were assessed baseline and after 8 weeks

World health organization quality Nutrition education of life (WHOQOLwas given to BREF) questionnaire, participants, no anthropometric further details were variables, BP, PR, lipid provided profile and fasting glucose were assessed

Participants were instructed not to make any dietary modifications

Following 8 weeks of Bikram yoga 1. Arterial stiffness Body composition, BP, decreased in obese adults arterial stiffness and 2. Emotional well-being health related quality of improved life were assessed Normal adults Obese adults 3. General health improved Normal adults

Cohen’s d

Findings

Level of evidence

Level 2

Yoga and Weight Management

Sl. no.

20

21

22

23

Citation

Sample

Yu et al., 2018

n = 79 centrally obese adults with MetS (n = 39 received 1-year yoga training and n = 40 received no training Age range: 58 ± 8 years

Telles et al., 2018

n = 52 centrally obese females Yoga group (YOG): n = 26, Nutritional advice group (NAG): n = 26 WC ≥ 80 cm, BMI ≥ 25 kg/m2 Age range: 30- 59 years

Kasturi et al., 2018

n = 32 obese females students Age range: 18-25 years

Telles at al., 2019

n = 596 participants BMI ≥ 25 kg/m2 yoga experience group n = 298 (154 females; group mean age ± SD; 44.0 ± 9.8 years) yoga naïve group n = 298 (154 females; group mean age ± SD; 43.8 ± 10.0 years)

Design

Random allocation of participants to two groups i.e. yoga and control

Diet information

Results

Cohen’s d

Findings

Level of evidence

Participants were asked to adhere their daily dietary intake

Following 1-year of yoga 1. WC decreased Yoga 2. HR decreased    Yoga 3. CS increased    Yoga 4. CSR left improved    Yoga Resting heart rate 5. CSR improved right (HR), chair stand test    Yoga (CS), chair sit and 6. Right side of BS reach test (CSR), back improved scratch test (BS), and    Control yoga single leg stand tests 7. SLS left increased (SLS)], serum samples    Yoga (HOMA), growth 8. SLSright increased hormone (GH), ghrelin,    Yoga UnAG, obestatin, 9. GH increased acylated ghelin (AG)    Yoga and insulin 10. Total circulating ghrelin increased    Yoga 11. UnAG increased    Yoga 12. AG decreased    Yoga 13. Obestatin decreased    Yoga

0.544 0.455 0.312 0.462 0.123 0.029 0.343 1.137 0.854 0.203 0.573 0.215 0.384 0.512

One-year yoga intervention modulated the Level 1 circulating level of ghrelin gene products and GH in adults with MetS

A diet plan for 1,900–2,000 Kcal/ day was given to both groups

Following yoga and nutritional advice for 12 weeks 1. WC decreased    YOG NAG WC, SAD, HC, 2. SAD decreased BMI, derived    YOG anthropometric indices, 3. HC decreased lipid profile, 24h diet    YOG NAG recall questionnaire, 4. BMI decreased International Physical    YOG Activity Questionnaire 5. Total cholesterol short form (IPAQ) increased and Moorehead   NAG Ardelt Quality of Life 6. LDL cholesterol questionnaire (QoL) increased were assessed    NAG 7. HDL cholesterol decreased    YOG 8. QoL improved YOG

0.30 0.81 0.39 0.36 0.46 0.22 0.54 0.41 0.50 0.70

Both yoga and nutritional advice with a diet plan can reduce anthropometric measures associated with diseases related to central obesity in females

Weight, WC, HC, MAC, WHR, BMI, general health questionnaire(GHQ12), satisfaction with life sacale (SWLS), perceived stress scale (PSS), Rosenberg self-esteem scale (RSES) and body shape questionnaire16(A)

Following 30 days of yoga 1. BMI decreased 2. WC decreased 3. HC decreased 4. WHR decreased 5. MAC decreased 6. PSS decreased 7. BSQ decreased

0.241 0.532 0.462 0.036 0.489 1.222 0.525

Yogic practices are effective to overcome complications arise due to obesity Level 2 and are helpful to induce positive psychological changes in obese individual

Moorehead–Ardelt quality of life questionnaire was used for assessment of six component of QoL

Enjoyment in physical activities, ability to work, self-esteem, and social satisfaction these four aspects were higher in experience group in compare to yoga naïve

0.09 0.08 0.06 0.06

Obese adults with yoga experience appear to have better quality of life in specific aspects, Level 3 compared to yoga naïve with a comparable degree of obesity

Single group pre-post study and participants No information were assessed available baseline and at day 30

Comparative trial

Variables Assessed

A diet plan for 1,900–2,000 Kcal/ day was given to both groups

Level 2

303

Section 3

Applications and Mechanisms of Yoga

305

Chapter 17

Overview of Yoga for Teenagers in the UK:

The Rationale, Evidence Base, and the Application Charlotta Martinus Teen Yoga Foundation, UK Nicholas A. Kearney Teen Yoga Foundation, UK

ABSTRACT This chapter outlines the current state of affairs of yoga provision for teenagers in the UK. It looks at the history of yoga in the UK among young people, the research and the application in school contexts, as well as prison and Pupil Referral Units. It outlines the results from the Hippocampus project, an EU-funded 2 year project evaluating the impact of yoga on disadvantaged youth in five countries. The chapter also takes a look at the current financial support for yoga in schools and the possibilities of taking it forward.

BACKGROUND OF YOGA FOR TEENAGERS IN THE UK The history of yoga for young people in the UK is relatively short. Although there have been pockets of yoga being taught as after-school clubs dotted around the UK, yoga had never been taken seriously as something that might actually change health, performance or behaviour until in the last few years. In 2011 Sport England started to recognise the importance and relevance of yoga for young people, contacting yoga deliverers to become more engaged in schools. Some forward -looking schools saw it as an alternative to Physical Education for those who needed something a bit gentler and less competitive. They continued to invest in yoga for young people within the school framework, developing programmes to lower the barrier to entry and include less advantaged young people in yoga as they found that 75 percent of 14 year old girls in the UK were no longer exercising (Sport England, 2016), primarily due DOI: 10.4018/978-1-7998-3254-6.ch017

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Overview of Yoga for Teenagers in the UK

to lack of desire for competition. By 2016 25 percent of schools across the UK offered yoga in some form or another (Teen Yoga Foundation, n.d.), either funded by Sport England or funded by the parents or the school itself. Furthermore, in 2016 Sport England invested £50,000 in an innovation project called Yoga Girls Can (Sport England, 2016, Martinus, 2016), which actively reached 700 young women between the ages of 14 and 21 through delivering a large “Yoga Yurt” to spaces, together with a yoga teacher and an assistant who gave impromptu 20 minute classes to suit the young people in their daily lives. This was considered a huge success with great follow on results, where the students were given free USBs if they completed 6 sessions. The USB had a video on it with yoga sessions they could continue with at home. They also received a yoga mat after 3 sessions to incentivise them to carry on. However, in 2018 the funding from Sport England was redirected to Primary Schools and Secondary Schools were no longer supported in the provision of yoga. Nonetheless, many schools continued with yoga, finding funding in their budgets. This was surprising as all secondary schools have seen a budget cut of approximately 30 percent in the last 10 years (Young Minds, 2018a). Schools voted with their feet, they were starting to see the relevance and utility of yoga in a school environment. Yoga has not only been available in schools, but also within the Pupil Referral Units (PRU) across the UK, these are institutions for pupils who cannot manage regular school, either for mental health issues or academic issues. The PRU often has much smaller classes due to the behaviour management issues, ranging from 6 – 12 students per class, which is a perfect number for yoga. Moreover, Child and Adolescent Mental Health (CAMHs) units across the country have also been offering yoga within various contexts, such as in outpatient and inpatient centres. Psychologists and psychiatrists take an add on course to train to deliver yoga within their professional capacity as an addon to allopathic care. Another area of interest which has grown in recent years is within the area of youth justice. Here is an outline of one individual who is implementing it and the rationale of why she is trained to deliver yoga within this context: At the moment, I am working within a Youth Justice and Preventions Service (YJPS), in mid-west Wales (From June 2014 until present). The YJPS is a multi-agency team including social workers, probation officers, substance misuse workers, project support workers and volunteers. The service accepts referrals for children and young people aged from 8 to 17/18 years old. These children and young people are assessed and classified as being at risk to themselves or other people, as well at risk of anti-social or offending behaviour. My role within the team is to recruit, train, manage and supervise the project support workers and volunteers. Project support worker and volunteer opportunities arise within the YJPS, Team around the Family, the Looked after Children’s Team, the Pupil Referral Team and Schools. At times, project support workers and volunteers work with/volunteer with, children and young people who are perceived as having challenging behaviour(s). Taking a focused approach, I am proactively in engaging with learning and development opportunities that could facilitate evidence-based, effective practice. Recent examples of this include ‘Non-violence resistance. A new approach to violent and self-destructive children’ by Haim Omer and ‘Connective Parenting’ by Sarah Fisher (Omer, 2004; Fisher, 2017).

306

 Overview of Yoga for Teenagers in the UK

Why The head teacher overseeing the Pupil Referral Units in the area, has identified an opportunity for me to teach mindfulness and yoga, to children and young people; the head teacher in high school has invited me to teach pupils there, plus the local wheelchair basket-ball team have shown an interest in my teaching them yoga. I have the support of the YJPS service manager to action these initiatives therefore I feel positive and confident to move forward in this direction. My core beliefs align with this opportunity: I value and believe in children and young people. Young people in the UK are accessing yoga in various contexts, even in youth centres, yoga centres and studios who now also offer tailored “teen yoga’ classes after school. The Teen Yoga Foundation estimates that approx. 30 percent of schools now offer yoga, either as part of the curriculum (either Physical education or Personal, Social, Health and Economic learning) or within the counselling service or as a voluntary club during lunch time, after school or before school. It is sometimes also offered before and after exams or specifically targeted students who have been identified as having social, emotional or mental health issues that could be helped with yoga, such as anxiety, low mood, self-harm or are experiencing bullying (Teen Yoga Foundation, https://teenyogafoundation.com).

RESEARCH TO DATE IN THE UK AND THE USA Thus far it is important to underline that the research in yoga for young people in the UK is scant and only consists of some minor research projects by MA students and the Teen Yoga Foundation teachers. However, the EU Hippocampus Project will change the profile of research in the UK in this topic, with a cohort of 750 students across Europe in different contexts, all from varied disadvantaged backgrounds, following the same yoga programme, rigorously researched both qualitatively and quantitively. The results for this project are coming together as we write this article. Khalsa and Butzer (2016) conducted a bibliometric analysis of research on school-based yoga programs and identified 47 studies, most of which reported beneficial effects of school-based yoga on student outcomes such as psychological well-being, academic performance and health-related measures. In 2013, Leeds University Psychology Department MA student Amy Morgan did a short qualitative analysis for her dissertation on what young people think of yoga and came up with these 6 themes: The interviews were conducted at a school in the city of Bath, with 12, 11 and 12 year old girls and boys. The classes included postures (usually a Sun salutation, some standing postures, some arm balances depending on what they came with that day), breathing exercises, and relaxation. An important part of the class was a checking in process, where each individual had the opportunity to share their day with the group. Any issues coming up in this discussion were then dealt with in the class (Morgan, 2013). Although this study is relatively insignificant, it is interesting to see the themes that emerge that would be interesting to pursue in further, larger scale research. Up to 2016 there hadn’t been any substantial research in yoga for young people in the UK. Sport England did a short piece of research in relation to the Yoga Girls Can project (www.sportengland.org, 2016), finding that the barrier to entry was often the cost, where classes can cost anywhere between £5 and £20 per lesson but also the perception of yoga which in the UK is often the pursuit of something that middle age and middle class women. They also felt that the presence of other young men there 307

 Overview of Yoga for Teenagers in the UK

would put them off coming and make them feel self-conscious. This research was done by interviewing a randomly selected group of young people from the 700 who had taken part in the innovation project. Figure 1.Number of participants

In 2017 the BBC children’s medical programme Ouch! conducted a research study with a small yoga group (12 teenagers between the ages of 14-17) in search for the ultimate de-stress technique before exams. They took a cortisol sample from each participant before and after the class, as they did with two other groups, a control group and also a choir. Here are the results: What is fascinating about these results is that not only is the cortisol lower than others after the event but also it is lower before even joining in the yoga. This seems to indicate that the stress levels of the yoga practitioners are continuously lowering through the week, even though they did not practise at home (BBC, 2016).

308

 Overview of Yoga for Teenagers in the UK

Figure 2.Changes in the cortisol level

Figure 3.

Most recently in 2019 a psychology student at the University College London Anna Freud Institute looked into what the barriers to entry were for young people coming to yoga and what their perceived benefits then were (Jackson, 2019). Her conclusion:

309

 Overview of Yoga for Teenagers in the UK

In the interviews, adolescents spoke of barriers to engagement: feeling judged, finding yoga “weird”, and the physical and mental challenge of yoga. Facilitators were: physical and mental benefits, getting a break from life, non-competitiveness, having a powerful group experience and having their needs met. Despite the physical challenge of yoga initially being perceived as difficult, it later became the motivating factor as participants saw their progression, developing resilience and a growth mindset. The focus on self-care in the yoga classes resulted in improvements in self-esteem, which helps to strengthen inner emotional stability in the face of future challenges, thereby making yoga a useful preventative intervention for adolescents. Participants describe yoga helping them to escape from overwhelming feelings using the breath and present moment awareness (Jackson, 2019). The most recent research and by far the most comprehensive is the Hippocampus Project, funded by the EU Erasmus Plus fund, ending in October 2019 which aimed (and managed) to reach 750 disadvantaged students and 100 staff across 5 countries with yoga. As far as we are aware this is the biggest research sample in Europe and the US of this kind. We were interested in how the yoga supported their wellbeing and measured this with the WEMWEBS scale, sleep quality and the perceived stress scale. What may be different about this research is it sets out to train staff to share some small part of yoga in their daily interactions with students. So, we measured the wellbeing of the staff as well as the students, believing that we need to support the entire structure of an organisation not just the students. The first quantitative results’ conclusion is as follows: The post yoga intervention results show an improvement in wellbeing, reduction in perceived stress and sleep related problems for staff and young people compared to the baseline. The changes have also been measured in terms of effect sizes which are found to be moderate overall. The effect sizes in Spain are the largest suggesting a more pronounced impact of the yoga intervention. The results for both the YP suggest that the effects on wellbeing, perceived stress and sleep related problems are likely to be smaller for those who have had previous experience of yoga compared those who have never practised yoga. There are several limitations to this study. This is a before and after study design without a control. We have therefore not been able to detect any time trend. Second, at follow up only 69 percent and 82 percent of staff and YP completed the measures. Using date of birth (even with country and gender together) resulted in being unable to match 12 staff and 41 YP. There were an additional 52 observations lost as participants (both staff and YP) entered the date they completed the survey instead of their birth date. This could be because of the way the data is collected online with the default being the date they completed the survey. Another limitation is that no process data was collected which means that we could not tell the number of sessions participants attended and how engaged they were as this is likely to impact on outcomes. The date of completion of the measures was not accurately recorded as the dates recorded sometimes referred to when the participant completed the questionnaire and sometimes to when the data was inputted. Therefore, it was not possible to report the time between completions of the measures and study the effects of time on wellbeing, stress and sleep. Although data for each country has been presented separately, the sample size in some is very small; hence any results need to be interpreted with caution. In conclusion, in this before and after study, there is a strong indication that yoga improves wellbeing, reduces perceived stress and improves sleep hygiene for both staff and YP (Hippocampus Project, 2019).

310

 Overview of Yoga for Teenagers in the UK

RATIONALE OF BRINGING YOGA TO TEENAGERS IN THE UK AGAINST THE BACKGROUND OF MENTAL HEALTH STATISTICS For a multitude of reasons, mental health among young people is at an all-time low, some call -25 percent of young people self-harming (Dixon, 2019) ◦◦ ◦◦ ◦◦

20 percent 14-18 year olds experiencing some kind of mental health illness (NHS, 2017) in 10 years 40 percent of 10-11 year olds will be obese (Sir Sam Everington) (Everington, 2019) waiting lists to see a Psychiatrist are up to 18 months long for young people (Bush, 2018; Young Minds, 2018b)

If and when the patients do get medical attention the drugs administered are often not licensed for under 18-year olds. What is more, there has been an increase in illegal selling of prescription drugs in among young people as well as an increase in illegal drug taking such as ketamine and marijuana (Lankenau, 2008). All of which have the potential to lead to increased instability of the mind. Against the backdrop of budget cuts both in CAMHS and secondary schools, yoga is increasingly responding to the mental health needs of schools and students. The Dept of Education reported that 25 percent of schools offered yoga in 2016 (Martinus, 2016). Eighty percent of the 1000 graduates of the teen yoga course to date are currently working in schools or other institutions with young people according to the Teen Yoga Foundation (https://teenyogafoundation.com). Yoga seems to support weight loss (Telles et al., 2014) thereby addressing the burgeoning obesity issue and also, by all accounts, supports optimal mental health, thereby addressing the tidal rise in these issues among young people in the UK currently.

APPLICATION OF YOGA TO TEENAGERS IN THE UK As yoga is taken more seriously and increasingly used as a methodology to prevent as well as cure mental health issues, it is essential that yoga teachers are properly educated and trained to a high standard that ensures safety and accountability in their delivery of yoga. Not only do teachers need to be highly skilled in yoga therapy but they also need to be skilled in delivery to particular age groups. This demand is becoming all the more clear as yoga teachers and therapists are invited into the realm of doctors’ surgeries and into schools, where the training for teachers is 4 years (3 year degree and 1 year PGCE) and doctors, a 7 year training. When public money is directed towards paying for yoga staff, the yoga community needs to ensure that we are providing high quality, properly trained staff. To this end, the Iyengar Yoga Institute (1190 members), Independent Yoga Network (3000 members), British Wheel of Yoga (7000 members), Yoga Alliance Professionals (4000 members) have come together to create a united front in the UK and furnish the government with accurate current information of the state of play of yoga in the country. We hope that this collaboration will go forward towards helping promote yoga in its entirety and ensure safety and quality for the student.

311

 Overview of Yoga for Teenagers in the UK

Furthermore, the Teen Yoga Foundation has created a course for professionals to ensure high quality training for those wishing to work with teenagers in different settings. Over the years and in collaboration with schools and professionals from the fields of medicine, education and psychology, the foundation has outlined the best practise of yoga with teenagers which includes – 1. Being flexible in our approach, so as to meet the needs of the students in any given situation rather than delivering a pre-determined course. The therapeutic element of the course is a very important aspect with this age group, helping them to feel safe and secure in the room with the teacher, where trust can easily be fostered. Therefore, the teacher needs to have certain therapeutic skills. 2. Honouring yoga in its fullest expression and as far as possible, including chanting (potentially in English, if Sanskrit is not acceptable) if possible, breathing exercises and pranayama (breathing exercises), asana (physical exercises), relaxation and some introduction of yamas (restraints) and niyamas (observances) as well as conversation about the concept of karma (action and reaction) and dharma (right living). 3. Where there is trauma or abuse, we suggest that the student lies on the front in savasana and keeps their eyes open. There is a complete protocol that is applied beyond these basic two measures. 4. Boys and girls are taught separately between the ages of 13 and 16 where possible 5. Classes contain no more than 15 students where possible and if there are more, then assistance is required 6. Enabling the students to learn certain postures and sequences that they can practise at home. We feel that one of the most important and a supportive element of yoga is how it allows us to take care of ourselves without help from anyone else. To this end, self-practise is an important element. Teenagers are ready to take on this challenge and often end up teaching others, within the family or friendship groups, thereby spreading yoga wider. As part of the class we encourage, therefore, students to take part in teaching each other. To this end the Hippocampus project also funded an app that is available worldwide especially designed for young people with help from their teachers. The app is currently called Yuva Yoga App. 7. The teachers themselves have access to mentors and supervisors who can support them through their development. 8 That the teachers are continuously updated and trained according to recent medical research and guidelines at the annual conference for yoga in schools, IN still, in London. There are many other best practise points that are specific to certain areas, such as anxiety, obesity, trauma, low mood, self-harm, eating disorders, stress and sleeping difficulties which we cannot go into in this paper.

PSYCHOPHYSIOLOGY OF YOGA FOR TEENS WITH STRESS AND HOW YOGA CAN HELP The over-riding issue that seems to be afflicting most young people, however, is stress. What we are finding with stressed young people is they are often fidgety, finding it hard to focus and be still. To counter this, we offer a fast paced, strong class, to use up the rajasic energy. Once they have burnt off the energy left over from the cortisol flood, they can then start to slow down and learn to hold the 312

 Overview of Yoga for Teenagers in the UK

postures for longer. We then often offer some arm balances as the young men are developing this area of their anatomy at this time and it helps them engage in the class. Very quickly, we find that they are keen to relax and recuperate, at which time we introduce the relaxation aspect. Often in the beginning, there is a lot of giggling and messing about as they are still unfocussed due to long periods of heightened cortisol activity. We keep the relaxation short and light and then come back to the breath, which is the aspect that really changes the mood and shifts the nervous system back into parasympathetic, so allowing the student to come back to focus and concentration with ease. By about the third time of practising yoga, many have come to understand how to relax and the relaxation quickly becomes the most popular part of the class. In wintertime, when they are particularly stressed and sleep deprived, they can have an entire class filled with yoga nidra. At this stage, we can see in their bodies, that they have understood how to engage their relaxation response at will, and this will then furnish them with the necessary toolkit for the rest of their lives to regulate their emotions and attention at will. The more academic of the group often are then interested in taking things further and want to investigate meditation and more advanced pranayama techniques. Teachers can judge whether the students are ready for this by how they can hold the asana “shtira sukha”. When this is achieved we can move on. This move from asana to pranayama, then to pratyahara (withdrawal of the senses, interoception) dharana (focus and concentration) and dhyana (meditation) can move very quickly for young people with a burning interest. It is important at this stage for teachers to be present and guiding the student as much as possible, to help them establish a home practise which they can take with them into their lives. In this way, each individual learns to “avoid future suffering” (Patanjali 2;16) and they are guided by a well-trained and compassionate yoga teacher, through their own Dark night of the soul, like Arjuna by Krishna (in the Bhagavad Gita Chapter 2-6), into determination, strong will and clear mind, with a clear idea of what their particular dharma might be. The teacher, if well trained can even teach the student to navigate their ego, so as to be the best they can be for the good of humanity and society. In this way yoga is an essentially pro-social tool and creates social cohesion within the school, the family and society at large (https://teenyogafoundation.com).

CASE STUDIES We have gathered together a few case studies from our working teachers here, to get a taste of how yoga is helping in the community in various situations. The names have been codified to adhere to GDPR (General Data Protection Regulation).

LY Lucy started coming to yoga aged 11 as she was a worrier and suffering from extreme anxiety exacerbated by her peer group at school. She was also having trouble getting to sleep. It took her mum a lot to convince her to come to the first class but once she had experienced that first session she kept coming back. Her concentration and eagerness to learn the asana were focused and she really enjoyed the visualisations I guided the class through at the end for relaxation.

313

 Overview of Yoga for Teenagers in the UK

She’s been attending classes for over 2 years and although she still calls herself a worrier she now uses the techniques she has learnt in yoga classes to ground and calm herself when needed. These include pranayama, asana and simple relaxation techniques including tapping.

MR Molly was admitted to hospital at the age of 15 with severe Anorexia Nervosa. While at hospital, she was introduced to yoga, which helped her to start to love her body again, through gentle movements and breathing exercises and through visualisations and support from the yoga therapist. She now manages her anxiety through yoga and continues to practise in order to support her wellbeing. She is now thriving and in full time study and is completely clear of Anorexia Nervosa.

RW This 14 year old girl was not getting on with her mother after a divorce, arguing and failing to thrive at school. She was bullied and suffered from anxiety and low mood after her parents divorced and her father moved away. With the practice of yoga, she became calmer and more forgiving. She also managed to feel greater compassion for her mother and father and find her own coping mechanisms and joy which would help build harmony in the home again. At 16 she became the youngest yoga teacher in the UK and is now helping others that have been in her situation.

CW Thirteen year old Charlie says: “Yoga to me is a sense of opportunity to feel relax and it helps you look into your mind and find the positive in your life. It has helped me to real myself and reduce my stress. It gives me a chance to open up and I would recommend it to other young people when things pile up and you cannot cope, it can help you let go of it all and it’s a chance to meet new friends too.”

Girl who Self-harms Yoga is freeing. It taught me self-love. It should be compulsory in all schools.

12-Year-old boy Yoga helps me with family issues, stress, for example when your brothers and sisters stress you out you can use it for a five minute break, using the breathing. You can also stretch yourself out and feel better. (Source: https://teenyogafoundation.com)

314

 Overview of Yoga for Teenagers in the UK

CONCLUSION We live in a time when the National Health Service is spending more money on support and medication for mental health issues for under 18s than ever before. £40 million per annum on addiction alone (Guardian, 2019). Eighty five percent of boroughs across the country are having their budgets cut for mental health services as the government pledges to invest in preventative health care (Guardian, 2019). We can only imagine how these mental health issues might impact society at large in the future. With 1 in 4 young people between the ages of 14 and 18 accessing medication to support their mental health, it looks like we are heading for the perfect storm in terms of continued high cost for the tax payer, absenteeism in the work place and unsatisfactory relationships. As our teenagers continue to navigate their lives in relation to this backdrop of misery and confusion, we need to step in and support them to find a more optimistic path, which includes self-empowerment, where they are no longer dependent on carers and medics but instead able to take their lives into their own hands, identify with clarity their needs and support themselves and each other. It seems from the emerging research that yoga is well placed to deliver this intervention and support our young in development towards a healthier future for them and their children. Things are looking positive for the world of yoga in the UK, with the recent investment from AYUSH into the propagation of ayurveda and yoga into British society (Dixon, 2019), we have good reason to hope that the medical community will become increasingly accepting and embracing of the health and optimal wellbeing that yoga can offer. As we speak, two prominent GP surgeries in the UK are trialling yoga for young people as a preventative measure, with a view to rolling the programme out nationally.

REFERENCES BBC. (2016). BBC Operation Ouch! CBBC. Retrieved from bbc.co.uk/cbbc/shows/operation-ouch Bush, M. (2018). New figures on Child and Adolescent Mental Health Service [CAMHS] waiting times. https://youngminds.org.uk/blog/new-figures-on-camhs-waiting-times/ Dixon, M. (2019). Body image is focus for mental health awareness week. Retrieved from https://collegeofmedicine.org.uk/body-image-is-focus-for-mental-health-awareness-week/ Everington, S. S. (2019). War begins to tackle the chicken shop obesity crisis. Retrieved from https:// www.eastlondonlines.co.uk/2019/01/war-begins-to-tackle-the-chicken-shop-obesity-crisis/ Fisher, S. (2017). Connective Parenting: A guide to connecting with your child using the NVR Approach. Create Space Independent Publishing Platform. Guardian. (2019). Retrieved from Theguardian.com/uk-news/2019/sep/10/addictive-nhs-opoid-crisisgovernment-study-england. Hippocampus Project. (2019). Hippocampus project: promoting mental health and wellbeing among young people through yoga. Retrieved from https://hippocampusproject.eu/about/ Jackson, P. (2019). Barriers and facilitators to engaging with a school-based yoga program: A qualitative exploration of adolescent girls’ views [Unpublished doctoral dissertation]. Anna Freud Institute, London.

315

 Overview of Yoga for Teenagers in the UK

Khalsa, S. B., & Butzer, B. (2016). Yoga in school settings: A research review. Annals of the New York Academy of Sciences, 1373(1), 45–55. doi:10.1111/nyas.13025 PMID:26919395 Lankenau, S. E., Sanders, B., Bloom, J. J., Hathazi, D. S., Alarcon, E., Tortu, S., & Clatts, M. (2007). Prevalence and patterns of prescription drug misuse among young ketamine injectors. Journal of Drug Issues, 37(3), 717–736. doi:10.1177/002204260703700311 PMID:18612374 Martinus, C. C. (2016). Interview with the Teen Yoga Foundation. Retrieved from https://www.soulseedmedia.com/2016/02/09/interview-teen-yoga-foundation/ Morgan, A. (2013). You can be the person you like in yoga, How young people explain the benefits of yoga [Master dissertation]. Institute of Psychological Sciences, University of Leeds, UK. National Health Service (NHS). (2017). Many teenagers reporting symptoms of depression. Retrieved from https://www.nhs.uk/news/mental-health/many-teenagers-reporting-symptoms-depression/ Omer, H. (2004). Non-violent resistance: a new approach to violent and self-destructive children. Cambridge University Press. Sport England. (2016). This girl can deliver results one year on. Retrieved from https://www.sportengland.org/news/thisgirlcanbirthday Sport England. (n.d.). This Girl Can: Inspiring women to get active. Retrieved from https://www.sportengland.org/campaigns-and-our-work/this-girl-can Teen Yoga Foundation. (n.d.). Yoga in Schools. Retrieved from https://teenyogafoundation.com/schools/ Telles, S., Sharma, S. K., Yadav, A., Singh, N., & Balkrishna, A. (2014). A comparative controlled trial comparing the effects of yoga and walking for overweight and obese adults. Medical Science Monitor, 20, 894–904. doi:10.12659/MSM.889805 PMID:24878827 Young Minds. (2018a). Children’s mental health funding: Where is it going? Retrieved from https:// youngminds.org.uk/blog/childrens-mental-health-funding-where-is-it-going/ Young Minds. (2018b). New figures on child and adolescent mental health services (CAMHS) waiting times. Retrieved from https://youngminds.org.uk/blog/new-figures-on-camhs-waiting-times/

316

317

Chapter 18

Yoga and Aging:

Neurobiological Benefits Rui F. Afonso Hospital Israelita Albert Einstein, Brazil Danilo F. Santaella Hospital Israelita Albert Einstein, Brazil Elisa Harumi Kozasa Hospital Israelita Albert Einstein, Brazil

ABSTRACT Governments and societies need to be prepared to confront population aging. Such preparation includes policies that can improve quality of life, functional capacity, and health of the general population, encouraging a more active and healthier lifestyle. Normal aging is associated with changes in brain structure and function, which may cause behavioral and cognitive impairments. It is important to understand which changes make some individuals healthier than others. Yoga has been associated with improved quality of life, cognition, and physical health as well as brain functional and structural changes.

AGING As longevity increases in the population, the incidence of aging-related diseases also increases, generating higher medical costs (Grootjans-van Kampen et al., 2014). Advances in prevention and treatment, in addition to improved public health policies, have successfully increased life expectancy. At present, the greatest number of deaths is due to diseases related to lifestyle (Rouquayrol & Almeida Filho, 1999). Lifestyle has components, such as such nutrition habits, physical activity, and leisure, which constitute important factors to promote health and quality of life. Both, Government and society need to be prepared to confront this state of affairs. Such a preparation includes policies that can improve quality of life, functional capacity, and health of the general population, notably the elder, encouraging a more active and healthier lifestyle. DOI: 10.4018/978-1-7998-3254-6.ch018

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga and Aging

As life expectancy increases, it is necessary to face this reality by tackling the problems of what has been described as unsuccessful aging. Young et al. (2009) define successful aging as: “A state wherein an individual is able to invoke adaptive psychological and social mechanisms to compensate for physiological limitations to achieve a sense of well-being, high self-assessed quality of life, and a sense of personal fulfillment even in the context of illness and disability”. This multidimensional approach takes into consideration three domains: physiological, psychological, and social. From this perspective, lifestyle is of fundamental importance for a healthy aging and also to promote a resilient brain (Taket et al., 2013) which is associated to reduced mental disorders in life, better mental health and greater psychological well-being (Smithet et al., 2014). Whilst sedentary elderly are more likely to develop health problems, a healthy lifestyle may help reverse the situation, potentially modifying risk profiles for frailty syndrome (Cesari et al., 2015), mortality, and overall disability (Seeman et al., 1995), contributing to a more successful aging (Kanning & Schlicht, 2008). In addition, normal aging, even before occurrence of Alzheimer’s disease or cerebrovascular disease, is associated with changes in brain structure and brain function, which may cause behavioral and cognitive impairments (Lockhart & DeCarli, 2015; Persson et al., 2006). Although aging causes changes in brain and cognition, some individuals, unlike others, despite advanced aging, have more preserved brain structure and more stable cognition over the course of aging. Thus, it is important to understand what kind of changes are associated with normal aging and, on the other hand, which makes some individuals healthier than others. Regarding the brain structure, accessed by MRI studies, white matter, cortical and subcortical gray matter may present decreased volume across the lifespan. Although most studies are gray matter ones, some studies show declines in white matter throughout aging, which also affect cognitive functions. Gray matter volume decline was found in subcortical areas such as the caudate, hippocampus and in the cerebellum (Raz et al., 2005). According to Persson and colleagues (2006) bilateral hippocampus was smaller in older adults with declining memory compared to older adults with stable memory. In this study, Persson and colleagues divided 40 older adults into two groups based on longitudinal episodic memory performance: 20 participants classified into stable memory and 20 participants, declining memory. Changes in hippocampal structure and differences in diffusion tensor imaging (DTI) measures of anterior white matter were associated with cognitive decline and it seems that a compensatory mechanism would be increased recruitment of frontal regions, once the brain and cognition are closely linked. De Carli et al. (2005) analyzed the brain MRIs from more than 2200 individuals. According to them, neurodegeneration related to age is associated to reduced cortical thickness in prefrontal cortex (PFC) areas, which presents the largest decline. Frontal subcortical systems are also affected by the aging process. Differences in brain volume in the frontal lobes were even greater in men compared to women. PFC is a source pole of many neuronal circuits. Degeneration of this region can compromise many cognitive or executive function and activities, such as working memory, attention, impulse control and many others. On the other hand, the relationship between PFC thickness and better executive performance was demonstrated in a meta-analysis conducted by Yuana and Raz (2014). Temporal lobe volumes also showed decline in older adults, resulting in declines in cognitive functions such as episodic memory, attention, working memory and executive control (Kirovaet et al., 2015). Thus activities with neuroprotective effect should be considered in a program for the elderly, contributing to cognitive preservation, better mental health and quality of life.

318

 Yoga and Aging

Brain function is also impaired in normal aging. fMRI accesses the brain activity through the Blood Oxygenation Level-Dependent (BOLD) contrast. Studies conducted with fMRI show differences in task-induced activity according to aging. Persson et al. (2006) observed that older adults with memory decline showed greater activity in the right ventral prefrontal cortex. Changes in brain connectivity and in the Default Mode Network (DMN) have been reported throughout life. DMN is a group of a brain areas including: medial prefrontal cortex, the precuneus/posterior cingulate cortex, the inferior parietal lobule, the lateral temporal cortex, and the hippocampal formation. DMN is more active when individuals are not engaged in any task. In other words, DMN is active when individuals are mind wandering. In a review, Damoiseaux (2017) described lower functional connectivity between regions of the Default Mode Network in older adults. Similar result was found by Ferreira et al. in a group of elderly without cognitive and psychiatric disorders (Ferreira et al., 2016). In addition to decreased resting-state functional connectivity, relationship between the DMN and the attentional networks is also altered in normal aging, as well as poorer cognitive performance in the elderly. In another study, Fjell and colleagues (2017) scanned two groups of volunteers: young and middle-aged, n = 63 and older adults, n = 56. In this longitudinal study, followed for 3.3 years assessing structural and functional brain connectivity and executive functions, decline in executive performance was related to reduced connectivity in aging. According to these authors, 82.5 percent of the age-related differences in executive functions could be explained by brain connectivity. Even though there are cognitive declines due to age, Li et al. (2016) reported brain activity correlated with improvements in cognitive functions such as memory and attention in groups of elders (age 65–75 years) undergoing cognitive training compared to older adults with no training. It seems that stress could accelerate the aging-related degenerative process. Besides that, stress also influences neural emotion processing. Based on this assumption, Everaerd compared a group of 25 young men with a group of 25 old healthy men in acute stress test during a fMRI (Everaerdet et al., 2017). Older adults showed a stress-related increased activity in the lingual gyrus extending into the fusiform gyrus compared to the younger adults. Since the impact of stress is different on the brain throughout aging, it is important to consider ways to alleviate stress and increase resilience in health programs for the elderly. Based on these considerations, yoga may be a promising intervention to develop mental health in the elderly. Yoga is a low impact activity, easily adaptable to people with special needs and limitations such as the elderly, positively impacting lifestyle. Yoga is defined as a meditative activity and its characteristics such as sustained attention and introspective vigilance distinguish yoga from the typical forms of physical exercise, although with similar metabolic intensity (Larson-Meyer, 2016; Mody, 2011; Sherman et al., 2017; Taimni, 2010).

BROAD BENEFITS OF YOGA Since yoga aims at establishing the utmost possible state of physical, mental and spiritual health, named Samadhi, in the process of pursuing this aim, many health stabilizing effects and even therapeutic processes happen, and thus, modern investigation aims at those, proving yoga to be effective, necessary, and useful for the contemporary society. Even though its effects are supposed to influence behavior, to do so it benefits both physiology and psychology. Nowadays, there are papers published in the medical journals showing different positive effects of yoga practices: asanas (postures) are related to reducing falls and chronic pain, improving balance, quality of life, flexibility, endurance and also decreasing depression and stress-related illnesses, such as cardiovascular and respiratory diseases. 319

 Yoga and Aging

As a matter of fact, although similar in energy expense, the effects of yoga and exercise on the brain and also on neurotransmitters seem to be different. Streeter et al., (2010 and 2018) have addressed the similarities or differences between asanas (postures) and physical exercise during activities conducted at the same metabolic rate, investigating levels of gamma amino butyric acid (GABA) before and after asanas or walking through magnetic spectroscopy, and found that the practice of asanas lead to a greater increase in GABA levels than walking (Streeter et al., 2010, 2018). Actually, this kind of effect was expected, since yoga practices differ from physical western activity for their attentional component. The same stretching practice, such as the sit and reach, which is called in yogic terms pascimotanasa (intense dorsal stretch pose), may lead to a different cognitive result, since the attentional process during the asanas must be to comfort, stability, effort relaxation, and unattachment from the self (Taimni, 2010), while when performed as a physical exercise, the attention may be free for wandering when and wherever the subject desires. Since yoga may be termed as a practical philosophy, its effects are broad and to try and make them a bit clearer, a didactical strategy of splitting the psychophysiological effects into psychological and physiological will be used in this chapter. Starting with the psychological variables, a recent study by Galantino et al. (2019) has addressed the effects of yoga practice on the risk of fall and quality of life of patients with chemotherapy-induced peripheral neuropathy syndrome. Individuals diagnosed were elder adults and attended at least 61 percent of the classes offered, which happened twice a week for 8 weeks for 1.5 hour. Authors found improved quality of life, flexibility, and balance. Also related to orthopedic variables, however investigating the relationship with psychological ones, a recent systematic review (Sieczkowska et al., 2019) has addressed the effects of yoga on the quality of life of patients with rheumatic disease, retrieving 23 eligible studies for their quality, and concluded that the practice of yoga is beneficial for patients with rheumatic disease, since it improves overall quality of life, and its main domains such as functional capacity, physical aspects, emotional aspects, general health, and social aspects. The authors conclude that there is evidence, although more research still needs to be done, to recommend yoga in the management of rheumatic disease patients. In another study, Telles and colleagues (2018) have investigated the effects of a 15-day immersion training in yoga for school teachers and have found increased well-being and decreased state anxiety only in the intervention group. Interestingly enough, the study was a randomized controlled trial, with a good sample size (n = 236, divided into yoga and control groups). In an additive manner, studies regarding the effectiveness of yoga for managing eating disorders have also been conducted. In a recent systematic review, Ostermann et al. (2019) have located only 8 randomized controlled trials eligible for the review, which indicate there is still great work to be done, since the quality of the available other papers is questionable. However, the authors conclude that, although there is limited evidence about the effectiveness of yoga for patients with eating disorders, it can be considered as an additional treatment option for multi professional psychiatric programs. Regarding other psychological conditions, such as post-traumatic stress disorder, another recent systematic review done by Nguyen-Feng and colleagues (2019) has addressed 12 articles, which evaluated a total of 791 volunteers with a range of interventions from 2 days to 16 weeks.The authors, however, didn’t find strong evidence for the effectiveness of yoga as intervention for post-traumatic stress disorder, or depression and anxiety symptoms, in patients who have had traumatic life experiences, and they state that this lack of strong evidence is due to the poor quality of the available papers and biased data collection. Such a recurrent trend of statements in the conclusions of the systematic reviews lead us to the comprehension that much still is to be done in improving the methodological quality and control of data collection in yoga-related research studies in order to improve the level of confidence of the evidence in this field. 320

 Yoga and Aging

Nevertheless, there are high-quality studies – mainly those addressing biological markers. There are studies involving pranayama training and/or its acute practice, in which variables and techniques were properly controlled and described. The effects of respiratory yoga practices seem to impact the basic cardiorespiratory interaction, figured by baroreflexes and chemoreflexes. Mason and colleagues (2013) have tested the acute effect of introductory ujjayi pranayama without apnea (breathing technique called “the ocean breath”) on baroreflex sensitivity through alpha index of heart rate variability using inspiration/expiration at 6 breaths per minute, with/without ujjayi’s characteristic glottis contraction for producing a peculiar noise due to a narrower air passage. In the group of 17 yoga-naive young healthy participants, baroreflex sensitivity increased with slow breathing regardless of the expiratory noise; the biggest effect on baroreflex with decrease in blood pressure was with slow breathing with equal inspiration and expiration. It is important to state that better baroreflex sensitivity may help the cardiovascular blood pressure control healthy through aging. Besides, Spicuzza and others (2000) have also found effects of slow breath on chemoreflex. Authors investigated chemoreflex through rebreathing maneuvers to isocapnic hypoxia and normoxic hypercapnia at 6 or 15 cycles per minute (the slower pacing aimed at mimicking the usual frequency adopted in many yogic pranayama breathing exercises) in ten healthy adult yoga trainees (average years of practice 7.9) and 12 healthy controls (yoga naïve) with similar body surface and age. Results showed that yoga breathing decreased chemoreflex hypoxic and hypercapnic responses. Authors argue that this may be due to the slower breathing frequency that yoga trainees exhibited spontaneously. They further state that the practice of yoga respiratory exercises named pranayama may have clinical relevance to medical conditions such as chronic heart failure, since it is strongly related to overactivation of the chemoreflex. Interestingly enough, yoga practice has positive and well documented effects on heart function. In a recent meta-analysis, Wang and others (2019) addressed 31 randomized controlled trials, with 1499 subjects, comparing the effects of different respiratory trainings on health status of heart-failure patients as expressed by heart rate, peak oxygen uptake (VO2 peak), 6-min walking distance test (6MWT), and Minnesota Living with Heart Failure QoL. Yoga respiratory training led to the second-best results in improving 6MWT in such frail patients. The effects of such training on chemoreflex may also be interesting for patients with chronic obstructive pulmonary disease (COPD). There are some evidences pointing to beneficial effects of pranayama for COPD patients. In a recent meta-analysis, Cramer and colleagues (2019) have investigated randomized controlled trials which have assessed the effects of yoga on quality of life, dyspnea, exercise capacity, and pulmonary function (FEV1) of COPD patients. Authors have selected 11 papers, which enrolled 586 patients. Positive results were robust for exercise capacity and pulmonary function, as a result of breathing yogapractices, however, those papers which have addressed interventions including yoga postures did not render such results. When investigating neurobiological variables, which may offer a bridge between psychological and physiological ones, some studies deserve specific consideration. For instance, in a controlled study, Telles and colleagues (2013) have shown that the acute practice of alternate nostrils breathing led to increased P300 amplitude and decrease of its latency, facts that may account for the understanding of how yoga practice may contribute with good effects on cognition. The training effects of bhastrika pranayama (“bellows” breathing) on heart rate variability have also been addressed in a randomized controlled trial in which the authors have investigated the effects of 4 months of half an hour pranayama training on sympathovagal balance and also pulmonary function of elderly subjects against a paired control group (same age and physical activity routine) (Santaella et al., 2011). Training has led the volunteers to significantly increase both their inspiratory and expiratory powers when compared to the control group, as well as to a decreased sympathovagal balance, with lower low frequency component and greater 321

 Yoga and Aging

high frequency component of the heart rate variability in the frequency domain. To corroborate these statements, some studies have shown different effects of yoga on the brain of long-term practitioners as compared to matched controls, paired by years of formal education, age and also physical activity. In fact, Afonso et al. (2017) verified a thicker superior and middle prefrontal gyrus in such elders who practice at least 8 years of Hatha Yoga, which may indicate a cognitive preservation, since these areas are related to working memory. In another study, Santaella and colleagues (2019) verified in the same population that the default mode network of the yoga practitioners maintains its anteroposterior connectivity, also indicating a cognitive preservation during the senescence process. These findings corroborate those found previously by Gard et al. (2015) which indicate an increased functional connectivity of the caudate nucleus with the left para hippocampal gyrus in older adults entrained in Kripalu Yoga or meditation (average 13,900 hours of practice in the yoga group, and 7,500 hours of practice in the meditation group) against a control group. Not only connectivity was greater in both groups, but also resilience was significantly greater in the yoga group than both meditation and control groups, indicating a greater potentiality for neural adaptation and cognitive preservation. When dharana (concentration) or dhyana (meditation) are focused, effects also show great potential on mood, attention and behavioral effects, as seen in studies which have found increased cortical thickness and functional connectivity in meditation and/or yoga practitioners. For example, in a study by Sevinc et al. (2018), two types of meditation were compared relaxation response and mindfulness-based stress reduction, after an 8-week training period. Both programs led to a stress reduction and had both similar as well as discrepant brain results. In an analysis that has considered both programs as a single group a greater functional coupling was found between ventromedial prefrontal regions and supplementary motor areas in the pre vs. post comparison. However, the separate-group analysis showed that the relaxation response group was associated with greater functional connectivity of the right inferior frontal gyrus with supplementary motor areas, while the stress reduction program was associated significantly increased functional connectivity of the right anterior insula with pregenual anterior cingulate when compared with rest. Such a finding may in part corroborate the previous findings of decreased sympathovagal balance in yoga practitioners, since the insular cortex is an important component of the autonomic regulation. There is also some evidence of the effects of yoga on multiple sclerosis. A study by de Oliveira et al. (2016) has shown that a 6-month training based on postures, breathing exercises, meditation, and relaxation for mere 60 minutes a week has led to significant improvements on Berg Balance Scale, Expanded Disability Status Scale, and selfreported quality of postural balance, besides a very interesting improvement in the positive influence of postural balance on activities of daily living. Put together, these results may indicate yoga to be a tool of enormous utility for gaining and maintaining physical, psychological and neural health.

NEUROBIOLOGICAL BENEFITS OF YOGA ON AGING Population is ageing; however, nobody wants just to live longer. If people can experience these extra years of life in good health and in a supportive environment, they may also keep the ability to do things they value. On the other hand, if these added years are characterized by declines in physical and mental capacity, serious implications may fall upon public health. According to the World Health Organization: “At the biological level, ageing results from the impact of the accumulation of a wide variety of molecular and cellular damage over time. This leads to a

322

 Yoga and Aging

gradual decrease in physical and mental capacity, a growing risk of disease, and ultimately, death. But these changes are neither linear nor consistent, and they are only loosely associated with a person’s age in years. While some 70-year-olds enjoy extremely good health and functioning, other 70-year-olds are frail and require significant help from others. Beyond biological changes, ageing is also associated with other life transitions such as retirement, relocation to more appropriate housing, and the death of friends and partners. In developing a public-health response to ageing, it is important not just to consider approaches that ameliorate the losses associated with older age, but also those that may reinforce recovery, adaptation and psychosocial growth.” There are many lifestyle changes that one might implement in life-routine to have a better ageing. Why should yoga be the chosen one? Having a strong sense of purpose in life leads to improvements in physical and mental health as well as to increased overall quality of life. In a prospective cohort study, 6985 individuals of the Health and Retirement Study (USA) had their life purpose assessed for the 2006 interview period with a 7-item questionnaire from the modified Ryff and Keyes Scales of Psychological Well-being (higher scores indicating greater purpose in life). Their all-cause and cause-specific mortality were assessed between 2006 and 2010. Results indicated that a stronger life purpose was associated with decreased mortality; besides, the authors mention that among the interventions which may help people to increase life purpose, quality of life and health outcomes there are volunteering, well-being therapy, and meditation (Alimujiang et al., 2019). Practicing yoga, especially in a traditional way, means that your life purpose would be to attain samadhi or other higher states of consciousness; to improve yourself as a human being and to promote virtuous and beneficial habits in your life. Yoga involves yama (abstinences), niyama (observances), asana (postures), pranayama (breath control), pratyahara (withdrawal of the senses), dharana (concentration), dhyana (meditation), and samadhi (absorption) and these practices may lead to self-realization, which is the primary goal of yoga (Taneja & Delhi, 2014). There are many changes in sleep due to aging. Sleep becomes more fragmented and there is an increase in the number of awakenings. There is a reduction in the amount of slow wave sleep (stage 3 and 4), a compensatory increase in the lighter stages of sleep (stage 1 and 2), and a decrease in rapideye-movement (REM) sleep. Sleep efficiency and total sleep time are reduced with age and the number of shifts in sleep stages is increased. Research has suggested that the elderly have a decreased ability to sleep (insomnia) (Cooke & Ancoli-Israel, 2011). Elderly living in old-age homes were recruited to a study aimed at examining the effects of yoga intervention on quality-of-life (QOL) and sleep quality. A block randomization based on the old-age homes was performed. A total of 120 subjects were randomized to whether a yoga group (n = 62) or a waitlist group (n = 58). The yoga intervention was offered daily for 1 month and weekly afterwards, until the completion of 3 months. The elders were encouraged to practice yoga without supervision until for 6 months. Subjects in the yoga group had significantly higher number of years of formal education. Subjects in the yoga group had significant improvement in all the domains of QoL and total sleep quality after the intervention period. Therefore, yoga seems to improve QoL and sleep quality of elderly living in old-age homes (Hariprasad et al., 2013). Among older persons, chronic pain is a common and costly disorder. Fifty percent of communitydwelling older adults reported chronic pain (Helme Rd, 2001). Self-management strategies have an important role in reducing pain and improving function in elders with chronic pain. In a review about such strategies, including yoga, massage therapy, Tai Chi, and music therapy, 96 percent of studies showed

323

 Yoga and Aging

positive outcomes (Jack et al., 2010; Reid et al., 2008). However, there is a lack of systematic reviews especially in the topic of yoga. It is known that the hippocampus is vulnerable for loss of grey matter with aging and that volume reductions in it are one of the earliest indicators of Alzheimer’s dementia (Jack et al., 2010; Raz et al., 2005). In a pilot study, seven healthy elderly subjects received yoga intervention composed by asanas, praṇayama and OM chanting. These practices were taught 5 days a week for 3 months; after those, subjects were instructed to keep practicing in the same manner and frequency, following a manual which described the practices. Magnetic resonance imaging scans were obtained before the beginning of the practices and 6 months later. Voxel-based-morphometric of the brain analyses, comparing hippocampal gray matter volume before and after yoga showed bilateral increases. The authors conclude that yoga has potential to reduce neuro-senescence, however the sample size is very small and there is no control group (Hariprasad et al., 2013). Yoga, Tai Chi and meditation require attentional engagement and have been associated with positive changes in brain structure and function, especially in areas related to awareness, attention, executive functions and memory (Lazar et al., 2005; Wei et al., 2013). In a cross-sectional study the authors compared cortical thickness (CT) in elderly female yoga practitioners with a group of age-matched healthy non-practitioners. Twenty-one participants who had practiced hatha yoga for at least 8 years and 21 women naive to yoga, meditation or any mind-body interventions were matched in age, years of formal education and physical activity level. Inside the magnetic resonance equipment, T1-weighted MPRAGE sequence was acquired for each participant. Yoga practitioners showed significantly greater CT in portions of the lateral middle frontal gyrus, anterior superior frontal gyrus and dorsal superior frontal gyrus. These regions are implicated in attention and working memory two cognitive abilities which may be impaired in elders (Afonso et al., 2017). During the aging process, brain has decreased connectivity of many networks such as the default mode (Mowinckel et al., 2012), working memory (Decarli et al., 2012) and salience (He et al., 2014) as well as decreased functional connectivity between the anterior and posterior structures of the Default Mode Network (Toussaint et al., 2014). In a cross-sectional investigation, long-term female elderly yoga practitioners and healthy paired yoga-naïve controls were compared in resting-state functional connectivity of the medial prefrontal cortex (MPFC) and posterior cingulate cortex-precuneus (PCC-Precuneus). The groups were paired (n = 20, hatha yoga practitioners; practicing a minimum of twice a week with a frequency of at least 8 years) and a control group (n = 20, yoga-naïve, matched by age, years of formal education, and physical activity) and were evaluated for: Mini Mental State Examination, Beck Depression Inventory, Instrumental Activities of Daily Living and open-eyes resting-state functional magnetic resonance imaging fMRI-seed to voxel connectivity analysis (CONN toolbox 17.f). There was no difference between groups in terms of age, years of formal education and questionnaires scores. Elderly women with at least 8 years of yoga practice presented greater intra-network anteroposterior brain functional connectivity of the DMN (between MPFC and the right angular gyrus) which may be related to a healthier cognitive aging process. Given the current scenario in which there are few papers exploring neuro- correlates of yoga practice in the elderly population, the field of research about the neurobiology of yoga may benefit from longitudinal studies with bigger samples. Nevertheless, the increasing evidences of the broad benefits of yoga for the elderly population may allow us recommending its practice as adjunct in the public health context in the near future.

324

 Yoga and Aging

REFERENCES Afonso, R. F., Balardin, J. B., Lazar, S., Sato, J. R., Igarashi, N., Santaella, D. F., Lacerda, S. S., Amaro, E. Jr, & Kozasa, E. H. (2017). Greater cortical thickness in elderly female yoga practitioners-A crosssectional study. Frontiers in Aging Neuroscience, 9, 201. doi:10.3389/fnagi.2017.00201 PMID:28676757 Alimujiang, A., Wiensch, A., Boss, J., Fleischer, N. L., Mondul, A. M., Mclean, K., Mukherjee, B., & Pearce, C. L. (2019). Association between life purpose and mortality among us adults older than 50 years. JAMA Network Open, 2(5), 1–13. doi:10.1001/jamanetworkopen.2019.4270 PMID:31125099 Cesari, M., Vellas, B., Hsu, F. C., Newman, A. B., Doss, H., King, A. C., Manini, T. M., Church, T., Gill, T. M., Miller, M. E., & Pahor, M. (2015). A physical activity intervention to treat the frailty syndrome in older persons - Results from the LIFE-P study. The Journals of Gerontology. Series A, Biological Sciences and Medical Sciences, 70(2), 216–222. doi:10.1093/gerona/glu099 PMID:25387728 Cooke, J. R., & Ancoli-Israel, S. (2011). Normal and abnormal sleep in the elderly. Handbook of Clinical Neurology, 98, 653–665. doi:10.1016/B978-0-444-52006-7.00041-1 PMID:21056216 Cramer, H., Haller, H., Klose, P., Ward, L., Chung, V. C. H., & Lauche, R. (2019). The risks and benefits of yoga for patients with chronic obstructive pulmonary disease : A systematic review and meta-analysis. Clinical Rehabilitation, 33(12), 1847–1862. doi:10.1177/0269215519860551 PMID:31353959 Damoiseaux, J. S. (2017). Effects of aging on functional and structural brain connectivity. NeuroImage, 160, 32–40. doi:10.1016/j.neuroimage.2017.01.077 PMID:28159687 De Oliveira, G., Cunha, C., Fernandes, G., Domingues, J., Oliveira, D. F., Rodrigues, M. R., & Santaella, D. F. (2016). Yoga training has positive effects on postural balance and its influence on activities of daily living in people with multiple sclerosis: A pilot study. Explore (New York, N.Y.), 12(5), 325–332. doi:10.1016/j.explore.2016.06.005 PMID:27426024 Decarli, C., Kawas, C., Morrison, J. H., Reuter-Lorenz, P. A., Sperling, R. A., & Wright, C. B. (2012). Session II: Mechanisms of age-related cognitive change and targets for intervention: Neural circuits, networks, and plasticity. The Journals of Gerontology. Series A, Biological Sciences and Medical Sciences, 67(7), 747–753. doi:10.1093/gerona/gls111 PMID:22570135 Decarli, C., Massaro, J., Harvey, D., Hald, J., Tullberg, M., Au, R., Beiser, A., D’Agostino, R., & Wolf, P. A. (2005). Measures of brain morphology and infarction in the framingham heart study : Establishing what is normal. Neurobiology of Aging, 26(4), 491–510. doi:10.1016/j.neurobiolaging.2004.05.004 PMID:15653178 Everaerd, D., Klumpers, F., Voshaar, R. O., Fernández, G., & Tendolkar, I. (2017). Acute stress enhances emotional face processing in the aging brain. Biological Psychiatry: Cognitive Neuroscience and Neuroimaging, 2(7), 591–598. doi:10.1016/j.bpsc.2017.05.001 PMID:29560910 Ferreira, L. K., Carolina, A., Regina, B., Kovacevic, N., Morais, G., Santos, P. P., … Busatto, G. F. (2016). Aging effects on whole-brain functional connectivity in adults free of cognitive and psychiatric disorders. Cerebral Cortex, 26, 3851–3865.

325

 Yoga and Aging

Fjell, A. M., Sneve, M. H., Grydeland, H., Storsve, A. B., & Walhovd, K. B. (2017). The Disconnected Brain and Executive Function Decline in Aging. Cerebral Cortex (New York, N.Y.), 27, 2303–2317. PMID:27073220 Galantino, M., Tiger, R., Brooks, J., Jang, S., & Wilson, K. (2019). Impact of somatic yoga and meditation on fall risk, function, and quality of life for chemotherapy-induced peripheral neuropathy syndrome in cancer survivors. Integrative Cancer Therapies, 18. doi:10.1177/1534735419850627 PMID:31131640 Gard, T., Taquet, M., Dixit, R., Hölzel, B. K., Dickerson, B. C., & Lazar, S. W. (2015). Greater widespread functional connectivity of the caudate in older adults who practice kripalu yoga and vipassana meditation than in controls. Frontiers in Human Neuroscience, 9, 1–12. doi:10.3389/fnhum.2015.00137 PMID:25852521 Grootjans-van Kampen, I., Engelfriet, P. M., & van Baal, P. H. M. (2014). Disease prevention: Saving lives or reducing health care costs? PLoS One, 9(8), e104469. doi:10.1371/journal.pone.0104469 PMID:25116681 Hariprasad, V. R., Varambally, S., Shivakumar, V., Kalmady, S. V., Venkatasubramanian, G., & Gangadhar, B. N. (2013). Yoga increases the volume of the hippocampus in elderly subjects. Indian Journal of Psychiatry, 55(7), S394–S396. doi:10.4103/0019-5545.116309 PMID:24049206 He, X., Qin, W., Liu, Y., Zhang, X., Duan, Y., Song, J., Li, K., Jiang, T., & Yu, C. (2014). Abnormal salience network in normal aging and in amnestic mild cognitive impairment and Alzheimer’s disease. Human Brain Mapping, 35(7), 3446–3464. doi:10.1002/hbm.22414 PMID:24222384 Helme Rd, G. S. (2001). The epidemiology of pain in elderly people. Clinics in Geriatric Medicine, 17(3), 417–431. doi:10.1016/S0749-0690(05)70078-1 PMID:11459713 Jack, C. R., Wiste, H. J., Vemuri, P., Weigand, S. D., Senjem, M. L., Zeng, G., Bernstein, M. A., Gunter, J. L., Pankratz, V. S., Aisen, P. S., Weiner, M. W., Petersen, R. C., Shaw, L. M., Trojanowski, J. Q., & Knopman, D. S. (2010). Brain beta-amyloid measures and magnetic resonance imaging atrophy both predict time-to-progression from mild cognitive impairment to Alzheimer’s disease. Brain, 133(11), 3336–3348. doi:10.1093/brain/awq277 PMID:20935035 Kanning, M., & Schlicht, W. (2008). A bio-psycho-social model of successful aging as shown through the variable “physical activity”. European Review of Aging and Physical Activity, 5(2), 79–87. doi:10.100711556-008-0035-4 Kirova, A.-M., Bays, R. B., & Lagalwar, S. (2015). Working memory and executive function decline across normal aging, mild cognitive impairment, and alzheimer’s disease. BioMed Research International, 2015, 1–9. doi:10.1155/2015/748212 PMID:26550575 Larson-Meyer, D. E. (2016). A systematic review of the energy cost and metabolic intensity of yoga. Medicine and Science in Sports and Exercise, 48(8), 1. doi:10.1249/MSS.0000000000000922 PMID:27433961 Lazar, S. W., Kerr, C. E., Wasserman, R. H., Gray, J. R., Greve, D. N., Treadway, M. T., McGarvey, M., Quinn, B. T., Dusek, J. A., Benson, H., Rauch, S. L., Moore, C. I., & Fischl, B. (2005). Meditation experience is associated with increased cortical thickness. Neuroreport, 16(17), 1893–1897. doi:10.1097/01. wnr.0000186598.66243.19 PMID:16272874

326

 Yoga and Aging

Li, T., Yao, Y., Cheng, Y., Xu, B., Cao, X., Waxman, D., Feng, W., Shen, Y., Li, Q., Wang, J., Wu, W., Li, C., & Feng, J. (2016). Cognitive training can reduce the rate of cognitive aging : A neuroimaging cohort study. BMC Geriatrics, 16(1), 12. doi:10.118612877-016-0194-5 PMID:26762334 Lockhart, S. N., & DeCarli, C. (2015). Structural imaging measures of brain aging. Neuropsychology Review, 24, 213–223. PMID:25146995 Mason, H., Vandoni, M., Codrons, E., Ugargol, V., & Bernardi, L. (2013). Cardiovascular and respiratory effect of yogic slow breathing in the yoga beginner : What is the best approach? Evidence-Based Complementary and Alternative Medicine, 2013, 743504. doi:10.1155/2013/743504 PMID:23710236 Mody, B. S. (2011). Acute effects of Surya Namaskar on the cardiovascular & metabolic system. Journal of Bodywork and Movement Therapies, 15(3), 343–347. doi:10.1016/j.jbmt.2010.05.001 PMID:21665111 Mowinckel, A. M., Espeseth, T., & Westlye, L. T. (2012). Network-specific effects of age and in-scanner subject motion : A resting-state fMRI study of 238 healthy adults. NeuroImage, 63(3), 1364–1373. doi:10.1016/j.neuroimage.2012.08.004 PMID:22992492 Nguyen-Feng, V. N., Clark, C. J., & Butler, M. E. (2019). Yoga as an intervention for psychological symptoms following trauma: A systematic review and quantitative synthesis. Psychological Services, 16(3), 513–523. doi:10.1037er0000191 PMID:29620390 Ostermann, T., Vogel, H., Boehm, K., & Cramer, H. (2019). Effects of yoga on eating disorders—A systematic review. Complementary Therapies in Medicine, 46, 73–80. doi:10.1016/j.ctim.2019.07.021 PMID:31519291 Persson, J., Nyberg, L., Lind, J., Larsson, A., Nilsson, L. G., Ingvar, M., & Buckner, R. L. (2006). Structure-function correlates of cognitive decline in aging. Cerebral Cortex (New York, N.Y.), 16(7), 907–915. doi:10.1093/cercor/bhj036 PMID:16162855 Raz, N., Lindenberger, U., Rodrigue, K. M., Kennedy, K. M., Head, D., Williamson, A., … Acker, J. D. (2005). Regional brain changes in aging healthy adults: General trends, individual differences and modifiers. Cerebral Cortex, 15, 1676–1689. Reid, M. C., Papaleontiou, M., Ong, A., Breckman, R., Wethington, E., & Pillemer, K. (2008). Selfmanagement strategies to reduce pain and improve function among older adults in community settings: A review of the evidence. Pain Medicine, 9(4), 409–424. doi:10.1111/j.1526-4637.2008.00428.x PMID:18346056 Rouquayrol, M. Z., & Almeida Filho, N. (1999). Epidemiologia e saúde (5th ed.). Medsi. Santaella, D. F., Balardin, J. B., Afonso, R. F., Giorjiani, G. M., Sato, J. R., Lacerda, S. S., Amaro, E. Jr, Lazar, S., & Kozasa, E. H. (2019). Greater anteroposterior default mode network functional connectivity in long-term elderly yoga practitioners. Frontiers in Aging Neuroscience, 11, 1–7. doi:10.3389/ fnagi.2019.00158 PMID:31312135

327

 Yoga and Aging

Santaella, D. F., Devesa, C. R. S., Rojo, M. R., Amato, M. B. P., Drager, L. F., Casali, K. R., Montano, N., & Lorenzi-filho, G. (2011). Yoga respiratory training improves respiratory function and cardiac sympathovagal balance in elderly subjects : A randomised controlled trial. BMJ Open, 1(1), e000085. doi:10.1136/bmjopen-2011-000085 PMID:22021757 Seeman, T. E., Berkman, L. F., Charpentier, P. A., Blazer, D. G., Albert, M. S., & Tinetti, M. E. (1995). Behavioral and psychosocial predictors of physical performance: Macarthur studies of successful aging. The Journals of Gerontology. Series A, Biological Sciences and Medical Sciences, 50(4), M177–M183. doi:10.1093/gerona/50A.4.M177 PMID:7614238 Sevinc, G., Hospital, M. G., Holzel, B., & Hashmi, J. A. (2018). Common and dissociable neural activity following mindfulness-based stress reduction and relaxation response programs. Psychosomatic Medicine, 80(5), 439–451. doi:10.1097/PSY.0000000000000590 PMID:29642115 Sherman, S. A., Rogers, R. J., Davis, K. K., Minster, R. L., Creasy, S. A., Mullarkey, N. C., O’Dell, M., Donahue, P., & Jakicic, J. M. (2017). Energy expenditure in vinyasa yoga versus walking. Journal of Physical Activity & Health, 14(8), 597–605. doi:10.1123/jpah.2016-0548 PMID:28422589 Sieczkowska, S. M., Casagrande, P. O., Coimbra, D. R., Vilarino, G. T., Andreato, L. V., & Andrade, A. (2019). Effect of yoga on the quality of life of patients with rheumatic diseases : Systematic review with meta-analysis. Complementary Therapies in Medicine, 46, 9–18. doi:10.1016/j.ctim.2019.07.006 PMID:31519294 Smith, J. L., & Hollinger-Smith, L. (2015). Savoring, resilience, and psychological well-being in older adults. Aging & Mental Health, 19(3), 192–200. doi:10.1080/13607863.2014.986647 PMID:25471325 Spicuzza, L., Gabutti, A., Porta, C., Montano, N., & Bernardi, L. (2000). Yoga and chemoreflex response to hypoxia and hypercapnia. Lancet, 356(9240), 1495–1496. doi:10.1016/S0140-6736(00)02881-6 PMID:11081541 Streeter, C., Gerbarg, P. L., Nielsen, G. H., Brown, R. P., Jensen, J. E., Silveri, M., & Streeter, C. C. (2018). Effects of yoga on thalamic gamma-aminobutyric acid, mood and depression : Analysis of two randomized controlled trials. Neuropsychiatry, 8, 1923–1939. Streeter, C. C., Whitfield, T. H., Owen, L., Rein, T., Renshaw, P. F., Ciraulo, D. A., & Jensen, J. E. (2010). Effects of yoga versus walking on mood, anxiety, and brain gaba levels. Journal of Alternative and Complementary Medicine (New York, N.Y.), 16(11), 1145–1152. doi:10.1089/acm.2010.0007 PMID:20722471 Taimni, I. K. (2010). The science of yoga (8th ed.). The Theosophical Publishing House. Tak, E., Kuiper, R., Chorus, A., & Hopman-Rock, M. (2013). Prevention of onset and progression of basic ADL disability by physical activity in community dwelling older adults: A meta-analysis. Ageing Research Reviews, 12(1), 329–338. doi:10.1016/j.arr.2012.10.001 PMID:23063488 Taneja, D. K., & Delhi, N. (2014). Yoga and Health. Indian Journal of Community Medicine, 39(2), 68–72. doi:10.4103/0970-0218.132716 PMID:24963220

328

 Yoga and Aging

Telles, S., Gupta, R. K., Bhardwaj, A. K., Singh, N., Mishra, P., Pal, D. K., & Balkrishna, A. (2018). Increased mental well-being and reduced state anxiety in teachers after participation in a residential yoga program. Medical Science Monitor Basic Research, 24, 105–112. doi:10.12659/MSMBR.909200 PMID:30061552 Telles, S., Singh, N., & Puthige, R. (2013). Changes in P300 following alternate nostril yoga breathing and breath awareness. BioPsychoSocial Medicine, 7(1), 11. doi:10.1186/1751-0759-7-11 PMID:23721252 Toussaint, P. J., Maiz, S., Coynel, D., Doyon, J., Messé, A., de Souza, L. C., Sarazin, M., Perlbarg, V., Habert, M.-O., & Benali, H. (2014). Characteristics of the default mode functional connectivity in normal ageing and Alzheimer’s disease using resting state fMRI with a combined approach of entropy-based and graph theoretical measurements. NeuroImage, 101, 778–786. doi:10.1016/j.neuroimage.2014.08.003 PMID:25111470 Wang, M. H., & Yeh, M. L. (2019). Respiratory training interventions improve health status of heart failure patients: A systematic review and network meta-analysis of randomized controlled trails. World Journal of Clinical Cases, 7(18), 2760–2775. doi:10.12998/wjcc.v7.i18.2760 PMID:31616691 Wei, G. X., Xu, T., Fan, F. M., Dong, H. M., Jiang, L. L., Li, H. J., Yang, Z., Luo, J., & Zuo, X. N. (2013). Can Taichi Reshape the Brain? A Brain Morphometry Study. PLoS One, 8(4), e61038. doi:10.1371/ journal.pone.0061038 PMID:23585869 Young, Y., Frick, K. D., & Phelan, E. A. (2009). Can successful aging and chronic illness coexist in the same individual? a multidimensional concept of successful aging. Journal of the American Medical Directors Association, 10(2), 87–92. doi:10.1016/j.jamda.2008.11.003 PMID:19187875 Yuana, P., & Raz, N. (2014). Prefrontal cortex and executive functions in healthy adults: A meta-analysis of structural neuroimaging studies. Neuroscience and Biobehavioral Reviews, 42, 180–192. doi:10.1016/j. neubiorev.2014.02.005 PMID:24568942

329

330

Chapter 19

Special Yoga for Children and Young People With Special Needs Jyoti Jo Manuel Special Yoga, UK

ABSTRACT Special Yoga has been a provider of therapeutic yoga practices and training since its inception. The non-profit organization started with a yoga center in London, UK. The Special Yoga London center offered a therapeutic and nourishing, nurturing space for families of children with special needs who were welcomed with love, compassion, and totally nonjudgmentally with open arms. The work spread globally through the London-based trainings and the therapeutic yoga that was offered to children at the center and within education. This chapter will discuss my experience and understanding of the efficacy of yoga for children and young people, specifically those with cerebral palsy and autism and/or attention deficit hyperactivity disorder (ADHD). The author shares some of the practices, benefits, and case studies of two separate research studies that were undertaken by Special Yoga for each population. The author also shares other case histories of children that they have worked with.

BACKGROUND Special Yoga has been a provider of therapeutic yoga practices and training since its inception. The nonprofit organization started with a yoga centre in London, U.K. The Special Yoga in the London centre offered a therapeutic and nourishing, nurturing space for families of children with special needs who were welcomed with love, compassion and totally non-judgmentally, with open arms. The work spread globally through the London based training and the therapeutic yoga that was offered to children at the centre and within education. Author’s note. This chapter is a personal account and will discuss the author’s personal experience and understanding of the efficacy of yoga for children and young people, specifically those with cerebral palsy and autism and/or attention deficit hyperactivity disorder (ADHD). The author will share DOI: 10.4018/978-1-7998-3254-6.ch019

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Special Yoga for Children and Young People With Special Needs

some of the practices, benefits and case studies of two separate research studies that were undertaken by special yoga for each population. The author will also share other case histories of children that they have worked with.

MEETING THE CHILDREN (Here, the author writes in the first person as this account was particularly meaningful to the author and a first-person description makes that evident). I recently attended a Breath-Body-Mind training with Richard P. Brown MD (Psychiatrist, Columbia University) and Patricia L. Gerbarg MD (Psychiatrist, New York Medical College). One of the first things Dr. Brown shared as he opened the training was “Your presence matters”. This has always been the under- pinning of our work. Working with these beautiful children is work of the heart and they teach us to truly open our hearts. I believe that children with special needs are our teachers (they can teach us so many things). I believe that these children have been brought here to show us the correct way of ‘being’. They embody unconditional love, the primary quality that is deeply needed in the world today. Our mental state does matter significantly. As Mohandas Karamchand Gandhi (1869-1948), says: “Your beliefs become your thoughts, your thoughts become your words, your words become your actions, your actions become your habits, your habits become your values and your values become your destiny.” It is said that 80 percent of communication is non-verbal. Our thoughts are a form of energetic communication. If for example, we are carrying judgments of how we think the children ‘should’ be or ‘should’ behave, or carry expectations of what we want to achieve from the sessions, we are not fully open-hearted to them. When we can meet them with the utmost humility, compassion, empathy (not sympathy), kindness and love, with no expectations, we create a powerful relational field in which the child can receive the benefits of these invaluable practices. Without the practitioner fully embodying these qualities the tools are unquestionably much less useful, and often in my experience may not be effective. The relational field I refer to here is where two individuals meet to form a third field of conscious awareness. It is where the two fields become interconnected. “A relational field that maintains the qualities of warmth, mirroring, empathy, compassion, loving kindness, safety, and trust must be established by the practitioner/teacher with appropriate boundaries (spaciousness) and contact free from the intention of “producing a result” (Shea, 2007; Shaw, 2017). Using principles of neurobiology, we learn how to listen to, and interact with, another person through embodied awareness. The yoga practice of embodied listening allows us to ‘feel’ within ourselves, giving us a pathway of authentic communication. It invites deep listening and to trust our intuitive sense to guide us. Conscious empathetic embodied listening supports the facilitation of the fullest expression of others. Breathing together is a beautiful, simple and profound pathway to create a healthy relational bond between the practitioner and the child. Yoga practice starts with the breath and the impact of the use of correct breathing patterns has well been recorded. Due to its relationship with the nervous system the breath is very powerful and an essential element of an authentic yoga practice. Conscious and regulated breathing is the simplest way to balance the sympathetic and parasympathetic nervous system.

331

 Special Yoga for Children and Young People With Special Needs

Our intention is to help these children to develop breath awareness, and learn pranayama practices that regulate their autonomic nervous system and therefore support their emotional and physiological regulation.

SPECIAL CHILDREN AND YOGA Children and young people with special needs will generally have a compromised nervous system and are often in the stress response (Institute of Medicine, 2001). They are likely to have high amounts of cortisol in the body. Studies show elevated cortisol in children and young people with special needs in situations that would be considered benign by their typically developing peers (Spratt et al., 2012). Furthermore, it is suggested that chronic stress and cortisol levels rise as children get older (Corbett et al., 2010). One of the many things that deep breathing and yoga practices do is to lower cortisol (Innis, 2018). Importantly it also increases levels of GABA, an inhibitory neurotransmitter which promotes calm throughout the system including promoting alpha brainwaves and is the major inhibitor to stress response (Ma et al., 2017; Hagen & Nayar, 2014). Research in 2007 and 2010 by Chris Streeter MD at Boston University found that after just one hour of yoga practice individuals had increased levels of GABA (Streeter et al., 2007; Streeter et al., 2010). Low levels of GABA are found in both anxiety and depression. Deep rhythmic breathing used in yoga practice stimulates the parasympathetic nervous system (through the vagus nerve) brings us into our calm state of rest and digest (Swani, 2015). As we know, participating in regular asana = (yoga posture) practice has also been shown to reduce cortisol levels. Increased cortisol levels have not only been linked to lower immune function, weight gain, and heart disease but they also impact learning and memory. Chronic stress and elevated levels of this stress hormone increase risk for depression and mental illness. Studies are starting to show that children who have participated in regular yoga instruction including pranayama (= voluntarily regulated yoga breathing), asana, meditation techniques, and relaxation are showing both lower salivary cortisol levels and improved performance on behavioral rating scales (Hagen & Nayar, 2014). A review done by Pilkington et al. (2005) has also shown the benefits of yoga on depression. Improved quality of breath led to ability to improved focus, quantity and quality of sleep, and improvement in digestion which helped in reducing depression. Digestive issues are commonly seen in children and young people with special needs (Pilkington et al., 2005). In a study by Woodyard (2011) it was shown that yoga impacts the overall quality of life of people by enhancing muscular strength and body flexibility, promoting and improving respiratory and cardiovascular function, promoting recovery from and treatment of addiction, reducing stress, anxiety, depression, and also alleviating chronic pain (Woodyard, 2011). Although many research studies show the benefits of yoga on able-bodied people there is paucity of studies conducted on persons with special needs. William Oppenhiem (2009) has discussed the use of complementary and alternative (CAM) therapy for people with Cerebral palsy (CP) where he has listed the use of yoga as an alternative therapy. Cerebral Palsy Alliance, Australia one of the forerunners in the area of research for children and people with CP states that the majority of evidence of yoga for CP is available only through individual stories, anecdotes, and case studies that strongly support the benefits of yoga as an exercise but state that availability of large scale studies are not yet available (Oppenheim, 2009; Ahmed et al., 2007). 332

 Special Yoga for Children and Young People With Special Needs

One ongoing study by Mak on the use of Mi-Yoga on children with Cerebral Palsy has been found. This study is being conducted at the Queensland Health Children’s Hospital and it has recruited 36 children in the age group of 6–16 classified as GMFCS I–III and is studying the impact of yoga on their attention at school and the stress on the child as well as the parents/care-givers (Mak et al., 2017; Australian New Zealand Clinical Trials Registry, 2013).

CEREBRAL PALSY AND YOGA Ryan McGraw wrote a thesis on adaptive yoga for disability (Mcgraw, 2006; Riedel, 1987). Cerebral palsy is the leading cause of childhood disability affecting function and development. Cerebral palsy is defined as: “a disorder of the development of movement and posture, causing activity limitations attributed to non-progressive disturbances of the fetal or infant brain that may also affect sensation, perception, cognition, communication, and behavior. Motor control during reaching, grasping, and walking are disturbed by spasticity, dyskinesia, hyperreflexia, excessive co-activation of antagonist muscles, retained developmental reactions, and secondary musculoskeletal malformations, together with paresis and defective programming. Weakness and hypo-extensibility of the muscles are due not only to inadequate recruitment of motor units, but also to changes in mechanical stresses and hormonal factors.” Cerebral palsy related impairment can range from severe to mild, with individual children affected differently, depending upon which parts of the brain are affected. The condition affects muscle tone, movement, and motor skills. Common physical challenges for children with cerebral palsy include muscle spasticity and tightness, low muscle tone, hyperactive reflexes, lack of voluntary motor function, cognitive delay, respiratory issues, seizure disorders, and vision impairment. The young person’s social and emotional development, speech and language development, attention skills, sensory integration, and ability to self- regulate can also be impacted. An increase or decrease in muscle tone impacts postural alignment. Movements can be jerky and involuntary which can in turn compromise breathing capacity. Breath can also be compromised due to the muscle spasticity or low muscle tone in the breathing apparatus. People with cerebral palsy who were affected at birth and then put on ventilators may have a longer term challenge with bronchial hyper reactivity (Kalyani et al., 2011). The emotional and mental health impact for a young person with cerebral palsy often becomes greater in adolescence because there is a greater awareness of what they are unable to achieve both physically and within in their life style (Downs et al., 2018). Additionally, there are simply less facilities for these young people post education, often creating severe isolation and resulting mental health issues. To date there has been very little research into the validity of using yoga therapy for young people with cerebral palsy specifically. ADAPT (formally the Spastic Society of India) under the leadership of Dr. Mithu Alur, Founder Chairperson of ADAPT, is a member of the Central Advisory Board of Education (CABE), New Delhi. Dr. Alur has also been appointed to the Round Table on School Education and Disadvantaged Sections: Women, Persons with Disabilities and SC/ST set up by Shri. Kapil Sibal, former Minister, Human Resource Development, New Delhi. She is a member of the Working Committee of the National Advisory Council (NAC), the Round Table on Elementary Education and the Anil Bordia Committee to synergies the RTE with existing programs. She has also been nominated to the National Monitoring Committee on the Education of Scheduled Castes/Scheduled Tribes/ Persons with Disability. She has been a pioneer in bringing services for people with Cerebral Palsy to India. 333

 Special Yoga for Children and Young People With Special Needs

Dr. Alur’s daughter, MC was introduced to the therapeutic practice of Special Yoga in the U.K. MC is a 50+ year old with quadriplegic CP. Whilst fully cognitive, her body has been affected. She is an author, holds down a job and is an avid scrabble player. She however needs assistance with basic life functions. Whilst MC had been practicing yoga for several years, she did not yet have the tools to regulate her anxiety. We worked intentionally to bring MC into a relaxed state and reduce her stress. The results of this were that MC was more able to access movement and regulate her breath and these practices proved to be a very important rehabilitation element of her recovery from a stroke. Dr. Alur was keen to bring the concept and style of Special Yoga back to India.

RESEARCH PILOT (This is narrated by the author as a first person account) In 2016, I was invited to offer the first Special Yoga training program for parents and therapists at ADAPT in Mumbai so that the yoga therapy program that was benefitting MC could be accessed by children and young adults with special needs in India. The workshops were received with keen enthusiasm by the parents and therapists. The benefits to the children and young adults with cerebral palsy who were part of ADAPT opened up the pathway to create and deliver a research study. The roots of ADAPT have been seeded within the context of ancient Indian philosophies and therefore the practice of yoga has been embedded into the educational practices at ADAPT since the time of its inception. ADAPT is also a centre for research for people with CP. Thus it was decided to implement a yoga therapy intervention with a small group of students with CP at ADAPT, and to track it through a systematic research study. The study was undertaken by a special yoga trained practitioner and aimed to demonstrate the physical and psycho-social impact of yoga on young adults with cerebral palsy. Data was collected pre and post intervention with a sample of 12 students in the age range of 18–35 years with Cerebral Palsy who were enrolled at the ADAPT Centers. The sample received individualized interventions by the special yoga specialist for a period of ten months. Along with the yoga intervention the trainees carried on the with their traditional physiotherapy sessions on a once a week basis. Each of the students served as their own control and changes recorded were compared pre and post on an individual basis. The sample was evaluated on the following parameters; breathing capacity, posture and physical measurements, and psychological impact. The assessments were done using the following tools; breathing capacity was assessed using the calibrated spirometer, chest expansion measurements and length of OM chanting, physical assessments were done through basic body measurements such as height, weight and girth of arm and thighs and postural assessments against a postural grid. The psychosocial impact was measured using the DASS 21 and the Warwick- Edinburgh Mental Well-Being Assessment Scale (WEMWBS). The results showed improvements in postural alignments and body parameters in majority of the sample. Eleven of the 12 trainees showed an increase in their height. The mean weight pre intervention was 50.7 where as post intervention it was 50.9. Nine of the 12 showed reductions in their girth measurements of arm and thighs. Eleven of the 12 showed changes in their vital capacity with the maximum improvement in vital capacity noted to be 1000 ml. All 12 trainees showed improvements in their length of their OM chanting and their chest expansion scores. In the DASS 21 scores showed reduction in stress scores for 9 out of 12 trainees. Average Mental Well Being was seen as the average scores dropped from 53.8 to 50.2.

334

 Special Yoga for Children and Young People With Special Needs

CASE HISTORY I first met one of the young people who participated in the study named X at a training workshop that I was leading for parents and pediatric therapists. X has quadriplegic cerebral palsy with some level of dystonia. X had poor trunk stability, and tension in his shoulders restricting movement in his arms. He, like many with CP clenched his hands while concentrating on making a movement which reduces breath capacity. His gait was unstable, his body showed a pronounced difference between his right and left leg lengths and as such he was unable to walk unaided. He was challenged and restricted in his ability to move hisarms and legs. I spent some time getting to know him, and exploring his breath, moments and also his emotional health. Despite the challenges he had with speech he was able to communicate slowly. He seemingly had no cognitive impairment with the condition although his emotional age did appear a little younger than his 21 years which was probably due to his restricted lifestyle. He expressed he felt that his biggest challenge was around his depression, anger and frustration and his desire for his body to ‘work better’. X was initially unable to breathe into the spirometer, so we were left to measure his breath capacity through the length of his OM. This was the case for most of the students who participated in the study. We asked that he focused on the sensation of the inner vibration as he chanted the sound of OM. Not only did the vibration help with vagal tone, the practice also helped to extend the length of his exhalation. The practice of chanting OM has now been researched and it has been found that the neuro-hemodynamic correlation of ‘OM’ chanting indicates limbic deactivation. Similar observations have been recorded with vagus nerve stimulation treatment used in depression and epilepsy. Many children and young people with CP also suffer from epilepsy (Kumar et al., 2010). The spiritual benefits are mentioned frequently in the Upanishads. Mandukya Upanishad tells us - “the word OM as one indivisible sound is the fourth state of supreme consciousness, a silence, and wholeness beyond birth and death, the symbol of everlasting joy. Those that know OM as the inner Self become the Impersonal Self, the Spirit, so OM is nothing but eternal intermingled Self and Spirit” (Backster, 2003). What was noticeable after the practice for X was a decrease in his depression and a new-found sense of joy. X found that the more he was able to increase and regulate his breath capacity, the more he was able to manage his stress, depression, anger and frustration. Breath awareness and breath control can also positively impact the quantity and intensity of spasmodic movements for those with dystonia. X had experienced significant pain in his hips which is common to those with CP due to dislocation in the hip joint. Often with chronic pain, there is breath holding, and often ‘anticipated’ pain before the pain actually arises. Using deep breathing and awareness, as well as a deep trust with the practitioner gave X the ability to move beyond the ‘anticipated’ pain and develop a wider range of movement. In X as with other people with CP, he had shortness through the fascial front line of the body and tension through the fascial back line exacerbating a thoracic kyphosis. His knee, hip and ankle joints were held in flexion. We used setubhanda and bujangasana to open and lengthen the front fascial plane and the hip flexors, and the iliopsoas which are often chronically shortened and weak from sitting. Within the first couple of sessions, there was a significant difference in his demeanor; he was more engaged, more able to control his dystonic spasms and thus felt more empowered to manage his body. There was a natural joy that arose within him as we found a connection between us to laugh and enjoy the practices together.

335

 Special Yoga for Children and Young People With Special Needs

Another aspect of the practice that we use extensively with this population is through bio- communication with the cells. Limited research has been done here. However, we have consistently seen that when we approach the body with genuine compassion and understanding that it is made up of conscious cells that experience emotions, it opens the possibility for positive change in our relationship with the body. That is not to say that this approach would ‘heal’ cerebral palsy in any way. However, we have seen that building trustthrough engaging the body in positive mental conversation by using different thoughts and words, elicits spontaneous elevated emotions and has also on occasion changed the physical and physiological responses in body control and movement (Backster, 2003).

CASE HISTORIES Many years ago I led a yoga session with Henry a 7 year old boy with hemiplegia. Both he and his mother were referring to the hemiplegic side of his body as his “bad” side. I invited the possibility of changing the dialogue to something kinder. By the end of the session he was able to embrace the possibility of calling this side of his body his ‘special’ side. His mother was initially mystified by this concept but a couple of weeks later she called me to let me know that the whole family and every one had around him embraced the ‘special’ side of his body. The results were that for the first time he was beginning to be able to move that side of his body voluntarily. This was also significant for X as for many who took part in the study. Changing his relationship with his body not only offered him a greater peace and acceptance, but offered a different experience of movement and relaxation. Interestingly but commonly he also had a lot of fear due to his dependency on others to help him to facilitate basic functions. Changing his relational field with his body and incorporating breathing practices daily helped empower him and alleviate some of that fear. The yoga practice has also been very significant for Sophie, a young girl of 11 with quadriplegic CP who has been working with me 1-1 for several months. For the first 6 months she would arrive at the studio crying and distressed, in very high levels of anxiety. She suffered with physical pain and epilepsy. Each session would consist of different practices adapted for her state. Throughout however I used ujjayi breath which she could both hear and feel, and found calming. I also encouraged her to speak more kindly to her body. After approximately 6 months of weekly sessions, she arrived smiling for practice. As she played down on the mat, she started to automatically use coherent breathing. A few months last with consistency in her breathing practice, she is pain and epilepsy free. Her relationship with her condition has improved and she is now empowered through teaching her family and friends coherent breath and gently stretching. Simple but effective asanas were used for this study intended to reduce high muscle tone, increase flexibility and strength, improve motion range in joints and overall mobility, reduce spasticity, to create the possibility of new pathways of movement, and to reform alignment and posture. Yoga poses are always modified to meet the physical needs and ability of the student based not heir level of mobility and motor coordination. Asana practice used in pilot study for cerebral palsy (adapted for each student) were lateral bend (parsva sukhasana), forward bend (pachimotasana), bridge pose (setubhandasana), sitting twists (ardha matsyendrasana), supine twists (adapted for dislocated hips), knee to chest (parva muktasana), cobra pose (bhujangasana) and savasana.

336

 Special Yoga for Children and Young People With Special Needs

Lateral bends and symmetrical stretching of the arms over the head combined with awareness on the breath to open the thoracic cavity, create intercostal space for improvement in breath capacity and the potential of deeper, more rhythmic breathing. Lateral bending is also an important movement to create more balance between the right and left sides of the body by working with each side individually. Gentle and adapted sitting and lying twists helped strengthen the deep multifidus muscle/ fascia, offering more spinal stability. These spiraling motions in particularly are useful for releasing tension held in the fascia. Possibly one of the more challenging aspects of the asana practice for X and again most people with CP is the combination of movement and breath together. For those with limited mobility, facilitation of breath and movement takes a great deal of work, concentration and awareness. It can be a very slow process to learn how to concentrate on breath and movement simultaneously. We facilitate periods of deep relaxation between movements so that the body and brain can assimilate the new pathways that are created through the practice. People with CP spend the majority of their day sitting. Extended periods of sitting exacerbate the forces of gravity. This creates compression in the abdominal cavity confusing the required space for the organs of digestion, respiration, circulation and heart to function well. Extended periods of sitting can also result in joint structure malfunction as there is more bone formation where there is more pressure. The soft tissues become hard and unyielding when held in one position. Movement allows our fascial planes to glide over one another. Releasing chronically held fascial and muscular tension in the body which excessive sitting will create, can also offer emotional and mental relaxation.

Anecdotal Feedback Some of the anecdotal feedback that came as a result of the study: “Doing yoga is the only time I don’t feel any tension in my body” “My back feels relaxed when I’m doing yoga. I feel comfortable in my body” Mother - “She has more upper body and core strength, which makes it easier for her to manage herself and move her body weight. This gives her more independence” “My breathing has improved. I find it easier to talk when my breath is relaxed. When I am able to control and slow my breath down, my breath and speech is more relaxed” “I am more aware of my breathing throughout the day”

YOGA AND AUTISM / ADHD Yoga is also very effective for children with autism and/or ADHD and other developmental and special educational needs (Kenny, 2002). Autism is equally a complex condition and whilst there are similarities in the dysregulation of the nervous system of those with the condition, the symptoms in each have its own variables. Autism is 337

 Special Yoga for Children and Young People With Special Needs

a complex neurobehavioral condition that can include impairments in social interaction and developmental language and communication skills combined with rigid, repetitive behaviors. Because of the range of symptoms, this condition is now called autism spectrum condition (ASC). Autism is a lifelong developmental disability that affects how people perceive the world and interact with others. Autistic people see, hear and feel the world differently to other people. Some autistic people also have learning disabilities, mental health issues or other conditions (Corbett et al., 2010). ADHD is Attention deficit hyperactivity disorder (ADHD) is a behavioral disorder that includes symptoms such as inattentiveness, hyperactivity and impulsiveness. Attention- deficit/hyperactivity disorder (ADHD) is a chronic condition that affects millions of children and often continues into adulthood. ADHD includes a combination of persistent problems, such as difficulty sustaining attention, hyperactivity and impulsive behavior. Children with ADHD may also struggle with low self-esteem, troubled relationships and poor performance in school (Cohen et al., 2018). Educational difficulties are also reported for children with Autism Spectrum Conditions (ASC) and Attention Deficit Hyperactivity Disorder (ADHD). Studies show that children with ASC in mainstream education are underachieving academically compared to typically developing students due to difficulties with attention, emotional and behavioral regulation, anxiety, depression, oppositional and aggressive behaviors (Ashburner et al., 2010). Similarly, children with ADHD are also reported to have difficulties with high rates of anxiety (Delli et al., 2018), poor attention, poor concentration, and low control of impulsivity; having a negative impact on their education (Taanila et al., 2014). Young people with ASC exhibit many behaviors their family, teachers, and other supporters find challenging. At the same time, autistic children often find the world at large a challenge. Many of these children will have language processing difficulties and/or sensory integration dysfunction and will experience great levels of anxiety and sometimes depression. They can be in a constant state of heightened arousal or anxiety and this will inevitably have a negative impact on the child’s health, mood and behavior. Children with ASD present unique challenges to the classroom teacher. The child’s inability to regulate his or her own behavior is often most problematic. At very high or low levels of stress or emotional arousal, complex cognitive processes are compromised. Intervention that addresses this arousal system may then affect functional behavior and educational achievement. Appropriate classroom behaviors include the ability to attend to task; transition from one activity to another; maintain a cal, organized state as demands change; and independently engage in schoolwork. Maintaining a calm, alert state is conducive to optimal learning in the classroom. Few opportunities exist in most educational curricula to train students in the skills required for maintenance of a calm but alert state (Goldberg, 2004). Literature shows that mindfulness based practices such as yoga for children may be useful to improve concentration, preparation, motor planning and control (Naveen et al., 1997); ability to self-regulate across environments and demand quality (Arch & Craske, 2006; Arch et al., 2006); and quantity of sleep (Khalsa, 2004). Within an educational setting yoga has been shown to improve resilience, self regulation, mood, managing emotions and stress (Hagen & Nayar, 2014). Yoga has also been reported to influence cognitive functioning (Ploughman, 2008) and improve executive functions in children (Manjunath & Telles, 2001) due to the reduction in stress and anxiety. There are some small studies that provide evidence that yoga practices specifically help children with ASC and ADHD (Abadi et al., 2008; Van de Weijer-Bergsma et al., 2012; Peck et al., 2005). In the UK, Autistic pupils are three times as likely to be excluded from school as pupils with no special educational needs. Misunderstood behavioral challenges are the most common mental health problem in early childhood, affecting 5-10 percent of young children and can have lasting effects on children’s 338

 Special Yoga for Children and Young People With Special Needs

life chances (Ryan et al., 2017). The NHS publication MHCYP (2017) reported that school exclusion in the UK is much more common for children with a diagnosed condition. A survey by the charity, the National Autistic Society (The National Autistic Society, 2016) showed that pupils with ASC are statistically more likely to be excluded from school this is confirmed by other sources (Barnard, 2000; Brede et al., 2017). Exclusion rates are highest for boys with a hyperactivity or behavioral condition with one child in twenty with these conditions having been excluded from school on three or more occasions; one boy in ten having been permanently excluded from school (National Health Service, 2018). There is a widespread consensus in the research literature that individuals who have been excluded from school experience greater negative outcomes across their lifespan and experience poor mental health. Moreover, school exclusion puts individuals on a pathway to social exclusion which is a traumatic experience shown to contribute to disempowerment of mental health (Barnes, 1999).

RESEARCH PILOT Responding to this substantive increase in exclusions from education, special yoga undertook a research pilot using a therapeutic yoga and mindfulness program for children with Autism/ADHD and other learning challenges in education. Exclusion leads to social isolation and poor outcomes for the future. School exclusion and social exclusion not only creates stress and worry for the individual and their families, it also may lead to mental health problems which puts a huge strain on the health system and economy. This pilot study examined the effects of a weekly therapeutic yoga intervention over 12 weeks for children aged 7-10 years across three main- stream primary schools. A group session of up to 10 pupils and a 1-1 session was delivered in each school. Teachers and parents monitored the children’s behavior, and emotional and cognitive regulation before and after the program using the Behaviour Rating Inventory of Executive Functioning (BRIEF-2) (Gioia et al., 2015). The BRIEF-2 is used to assess children who have either developmental or acquired neurological conditions such as learning disabilities, attention disorders, traumatic brain injury, pervasive developmental disorders, depression, and other developmental, neurological, psychiatric, and medical conditions. In this instance the BRIEF-2 was used to measure symptoms before and after the program from the perspective of the parents and teachers. The results show improved executive functioning in all domains following the yoga intervention. The teachers’ results demonstrated an improvement in all areas with the greatest reduction seen in Shift (a diagnostic marker of Autism Spectrum Conditions); emotional control; self- initiation; planning and organizing; and behavioral, emotional and cognitive regulation. Parents’ results demonstrated a reduction in markers across all domains with the exception of measures of self-monitor and working memory. The conclusion showed that therapeutic yoga within education may be advantageous for stress management, behavior, emotional regulation and cognitive functioning for this population of children. Teachers and teaching staff participated in a three-hour yoga and mindfulness training prior to the commencement of the yoga intervention. The tools shared were intended to support their emotional resilience and mental health. Included were practices to understand the importance of embodied listening to create the relational field they hold with the children, as well as regulatory tools with breath work, simple asanas and sensory integration techniques that they could use during the school day both for themselves and for the children. The teaching staff that attended visibly relaxed during the session and learned quickly the importance of their own personal self care and practice. The program also offered consultation to SEN coordinators, teachers and parents during the intervention. 339

 Special Yoga for Children and Young People With Special Needs

Understanding that behavior is a form of communication offers an alternative approach to managing difficult behaviors that are often due to an underlying anxiety, frustration or sensory sensitivity. When we as practitioners /careers /educators are able to drop our judgements and expectations and truly ‘listen’ with compassion and kindness we will be able to ‘hear’ the young person’s unmet needs. We can then support the potential for change in their emotional state by offering them a safe space within which they have an opportunity to regulate and calm. Ongoing behavioral management whereby the behavior is deemed as something ‘bad’ cannot and will never meet the needs of the child and tackle the underlying issues that the child is experiencing. No child is ‘bad’ irrespective of how ‘bad’ their behaviors may present. Acknowledging that the child is not their behavior honors and respects the child. The child is first and foremost a child. Often autistic people are treated as though they merely do not want to listen or control their behavior, rather than that they are unable to do so. This is particularly relevant in oppositional defiance, tantrums, and high sensitivity. Y is a 9 year old boy who participated in this study. He is stranded in the lengthy pathway for official diagnosis. He is clearly on the autistic spectrum. When we first met him in the school setting, he was sitting separately from everyone, totally disengaged from the world around him. The adults around him were clearly fearful of the potential of his frustration and anger tantrums, where historically he had hurt himself or another, or both. Fear of a child’s behavior does not and cannot offer a ‘safe’ space for the child. We also recognize that his angry outbursts to do not always create a ‘safe’ space for the teaching staff or the other children creating a difficult challenge for the adults around him. Three rural primary schools participated in this pilot study. Each school selected up to 10 children for a group session and one child for a one to one yoga session. One child had a formal diagnosis of ASC, one had a diagnosis of ADHD, at least three children were on the pathway to a diagnosis and all of the other children had social/ emotional or behavioral difficulties. Children were chosen based on the concern that they are having some difficulties at school and may be at risk of exclusion. Twelve weekly sessions of yoga were delivered in the schools by a special yoga practitioner. One group session per school and one 1-1 session per school. The group sessions were delivered on the same day and time every week.

PRACTICE USED FOR RESEARCH PILOT Breathing practices chanting OM and Hari Om, deep pressure/ squeezing arms/legs forward bend (pachimottasana), lateral bends (parsva sukhasana), spinal twist-adapted (ardha matsyendrasana navasana), rolling back and forth, spinal flexion and extension from supine back bend (setubhandasana), knees to chest (apanasana), sun salutation sequence x 2, balancing on one leg (vriksasana, viparita shalabhasana), back bend (dhanurasana, apanasana), relaxation (savasana), meditation and positive affirmation mantra. Prior to the official commencement of the pilot study, we evaluated the children that had been selected through observation within the classroom setting and within an initial yoga session where the children were invited to participate. This observation led us to the practice that was used for the pilot study. An over or an under stimulated sensory system is often the source of many behavioral responses. Children need to move regularly, to have safe ‘mindful moments’ to feel their bodies, be aware of their senses and learn ways to self-regulate their sensory systems. Yoga and mindfulness help children to recognize how their bodies are feeling and how to link these sensations to their emotional learning and communication skills. We offered the children a grounded and embodied practice through asanas, pranayama, sensory integration practices, self massage, meditation, positive mantras, chanting and deep relaxation. Feeling 340

 Special Yoga for Children and Young People With Special Needs

safe, connected and calm automatically addresses a number of the ‘behaviors’. Children learn in multisensory ways and the ability to decode and process sensory feedback from all the senses, including the proprioceptive, vestibular and tactile systems are vital for the holistic health of the child. The young people participating in the research pilot were taught coherent breathing, bhramari breath (= bumble bee breath), and nadishodana (= subtle energy channel cleansing breath). Regulating the length of the in breath and out breath with counting was also used to slow and regulate the breath patterns. The asanas included a balance of simple but effective practices to help the children to be more embodied and grounded, build core strength, increase flexibility and open the energetic pathways. We oftenused the chanting of AUM within the asana practice to elongate the exhalation to stimulate the parasympathetic nervous system. Therapeutic yoga practices support our emotional, physical and physiological sense of safety. Safety in the body is key to authentic embodied listening. According to the neuroscientist Stephen Porges, the neurobiology of our body needs to feel safe in order for us to socially engage (Porges, 2017). The work of Porges’ polyvagal theory informs much of our practice to enhance the regulation and wellbeing of children with autism and/or ADHD. Porges states that the neural regulation of our bodily organs influences our emotional responses and behavior toward others and our environment. Many people with autism have difficulty regulating their behavior and emotions. Their initial reaction to threat is often anger, irritability or aggression that may be expressed as an uncontrollable tantrum. These responses constitute a fight or flight reaction that can be difficult to manage. Porges’ research has documented parallels between the autonomic system, the system that controls the fight or flight system and the functioning of our organs, and difficulties with learning and socializing. Essentially, if we cannot regulate our physiological state, we cannot socialize and connect with others. The autonomic nervous system controls unconscious bodily functions such as heart rate, breathing, and the fight or flight response. Based on his research, he hypothesizes that the system functions like a neural platform that can facilitate or suppress various types of behavior, from spontaneous social interactions to aggressive and oppositional behavior and complete shutdown. These of course are many of the challenges that autistic children can present with. The polyvagal theory offers an explanation for some of autism’s core features, from social difficulties to gut dysfunction and sensory sensitivity. A therapeutic yoga program and can ease the severity of these issues (Porges, 2005). We have consistently found that children with ADHD respond almost immediately to a practice of Ha breath and tapping, taught by Dr. Brown as part of the BBM program (Bryant-Davis, 2011; Telles et al., 2015). Simple repetitive mantras help bring focus to scattered minds which is often seen in ADHD and ASC. Rhythmic chanting used in a ‘call and response’ manner, keeps the child’s attention focused. In the pilot study we used the chant ‘Hari Om’ with a simple repetitive rhythm. We added hand movements and mudras (= hand gesture) to help with motor planning and coordination. We added cross body movements which are often challenging for those out of regulation. These movements improve communication and integration of the left and right hemispheres of the brain, supporting brain-body connection, brain function, and fine and gross motor development. Crossing the midline is a necessary action for development and learning, bilateral coordination and thus many physical activities. It also sup- ports thinking and concentration, because such actions necessitate communication between the left and right hemispheres of the brain, which can directly impact writing, reading, and attention skills. The “mid-line” being an imaginary line down the middle of the body (from the brain to the feet), separating the left and right halves of the body.

341

 Special Yoga for Children and Young People With Special Needs

CASE HISTORY Although Y didn’t actively participate in the initial session, he was observing from a distance with some level of interest. When we spent time with him, it was clear that he felt deeply disturbed by the lack of control he experienced with his angry tantrums. We worked 1-1 with him as he was unable to access the group session. The individual time with this highly intelligent young man gave us the opportunity to share very basic neuroscience of the prefrontal cortex and the amygdala which helped enormously in his ability to regulate. M, an autistic child who also participated in the study was disconnected from himself and others, unable to make eye contact, and without a sense of where his physical body was in space. Using selfsqueezing and self massage with rhythmic chanting as well as tapping on the body helped to establish his proprioceptive sense as well as creating a regulatory internal rhythm. Yoga Nidra and a more basic deep relaxation practice were imperative for both cohorts. There was a noticeable improvement for the children with ADHD for whom settling into relaxation was impossible in the very early stages of the pilot. By the end of the intervention, all the children were able to find longer moments of rest and relaxation, often lying prone as opposed to the classical savasana.

Anecdotal Feedback The anecdotal feedback from the children, parents and teachers who were involved with the pilot study included: “Before yoga I felt very upset because my daddy is sick. After yoga i feel calm and peaceful” “Before yoga I feel hyper, after yoga I feel nice and calm. I feel sleepy”. Parent - “my child has finally slept through the night for the first time in 9 years. Yoga has changed our lives.” The teachers’ feedback for Y on the Brief evaluation form showed significant improvement in selfregulation, and improved ability to stay on task and initiate.

RECOMMENDATIONS Yoga has multitudinous benefits for children and young people with special needs, and all can find a pathway for practice. I believe that the key to working with these children lies in our ability to meet the child rather than their disability, without judgment or expectations, but with compassion and kindness and with an open heart. Keep the practice simple and embodied. Keep in mind always that we are not here to ‘fix’ these children. They are perfect as they are. We are just here to open a pathway with practices to help them to thrive and reach their potential. THE AUTHOR: I stepped into my first yoga class in 1974 and knew immediately that I had found the path to my own healing of early life trauma. I am passionate about its potential to enhance the lives

342

 Special Yoga for Children and Young People With Special Needs

of children with special needs. In 2003, I founded Special Yoga with a mission to reach as many children and young people with special or additional needs with yoga (Special Yoga, n.d.).

REFERENCES Abadi, M. S., Madgaonkar, J., & Venkatesan, S. (2008). Effect of yoga on children with attention deficit/ hyperactivity disorder. Psychological Studies, 53, 154. Ahmed, S., Ross, M., Trout, K., & Zhang, J. (2007). Simulation of surface water for un- gauged areas with storage-attenuation wetlands. Journal of the American Water Resources Association, 43(2), 546–556. doi:10.1111/j.1752-1688.2007.00043.x Arch, J. J., & Craske, M. G. (2006). Mechanisms of mindfulness: Emotion regulation following a focused breathing induction. Behaviour Research and Therapy, 44(12), 1849–1858. doi:10.1016/j. brat.2005.12.007 PMID:16460668 Arch, J. J., Craske, M. G., Stein, M. B., Sherbourne, C. D., & Roy-Byrne, P. P. (2006). Correlates of alcohol use among anxious and depressed primary care patients. General Hospital Psychiatry, 28(1), 37–42. doi:10.1016/j.genhosppsych.2005.08.007 PMID:16377363 Ashburner, J., Ziviani, J., & Rodger, S. (2010). Surviving in the mainstream: Capacity of children with autism spectrum disorders to perform academically and regulate their emotions and behavior at school. Research in Autism Spectrum Disorders, 4(1), 18–27. doi:10.1016/j.rasd.2009.07.002 Australian New Zealand Clinical Trials Registry. (2013). MiYoga: Mindfulness yoga for children with cerebral palsy. https://anzctr.org.au/Trial/Registration/TrialReview.aspx?id=362966&isReview=true Backster, C. (2003). Primary perception: Biocommunication with plants, living foods, and human cells. White Rose Millennium Press. Barnard, J. (2000). Inclusion and autism: Is it working? National Autistic Society. Barnes, H. (1999). Exclusion and mental health—The relationship context of mental health practice. Social Work Education, 18(4), 401–415. doi:10.1080/02615479911220401 Brede, J., Remington, A., Kenny, L., Warren, K., & Pellicano, E. (2017). Excluded from school: Autistic students’ experiences of school exclusion and subsequent re- integration into school. Autism & Developmental Language Impairments, 2, 1–20. doi:10.1177/2396941517737511 Bryant-Davis, T. (2011). Surviving sexual violence: A guide to recovery and empowerment. Rowman & Littlefield Publishers. Cohen, S. C., Harvey, D. J., Shields, R. H., Shields, G. S., Rashedi, R. N., Tancredi, D. J., ... Schweitzer, J. B. (2018). The Effects of Yoga on Attention, Impulsivity and Hyperactivity in Pre-school Age Children with ADHD Symptoms. Journal of Developmental and Behavioral Pediatrics, 39, 200–209. doi:10.1097/ DBP.0000000000000552 PMID:29538185

343

 Special Yoga for Children and Young People With Special Needs

Corbett, B. A., Schupp, C. W., Simon, D., Ryan, N., & Mendoza, S. (2010). Elevated cortisol during play is associated with age and social engagement in children with autism. Molecular Autism, 1(1), 13. doi:10.1186/2040-2392-1-13 PMID:20875126 Delli, C. K. S., Polychronopoulou, S. A., Kolaitis, G. A., & Antoniou, A. S. G. (2018). Review of interventions for the management of anxiety symptoms in children with ASD. Neuroscience and Biobehavioral Reviews, 95, 449–463. doi:10.1016/j.neubiorev.2018.10.023 PMID:30395924 Downs, J., Blackmore, A. M., Epstein, A., Skoss, R., Langdon, K., Jacoby, P., Whitehouse, A. J. O., Leonard, H., Rowe, P. W., & Glasson, E. J. (2018). The prevalence of mental health disorders and symptoms in children and adolescents with cerebral palsy: A systematic review and meta‐analysis. Developmental Medicine and Child Neurology, 60(1), 30–38. doi:10.1111/dmcn.13555 PMID:28914445 Gioia, G. A., Isquith, P. K., Guy, S. C., & Kenworthy, L. (2015). BRIEF-2: Behavior Rating Inventory of Executive Function. Psychological Assessment Resources. Goldberg, L. (2004). Creative Relaxation SM: A Yoga-based program for regular and exceptional student education. International Journal of Yoga Therapy, 14(1), 68–78. doi:10.17761/ijyt.14.1.50524j511674v292 Hagen, I., & Nayar, U. S. (2014). Yoga for children and young people’s mental health and well-being: Research review and reflections on the mental health potentials of yoga. Frontiers in Psychiatry, 5, 35. doi:10.3389/fpsyt.2014.00035 PMID:24765080 Innis, G. (2018). Understanding cortisol, the stress hormone. https://www.canr.msu.edu/news/ understanding_cortisol_the_stress_hormone#:~:text=Practice%20mindfulness%20and%2For%20 meditation,provide%20a%20sense%20of%20calm Institute of Medicine. (2001). Neurological, psychiatric, and developmental disorders. National Academy Press. Kalyani, B. G., Venkatasubramanian, G., Arasappa, R., Rao, N. P., Kalmady, S. V., Behere, R. V., ... Gangadhar, B. N. (2011). Neurohemodynamic correlates of ‘OM’ chanting: A pilot functional magnetic resonance imaging study. International Journal of Yoga, 4(1), 3–6. doi:10.4103/0973-6131.78171 PMID:21654968 Kenny, M. (2002). Integrated movement therapy™: Yoga-based therapy as a viable and effective intervention for autism spectrum and related disorders. International Journal of Yoga Therapy, 12(1), 71–79. doi:10.17761/ijyt.12.1.r978vxt214683904 Khalsa, S. B. S. (2004). Treatment of chronic insomnia with yoga: A preliminary study with sleep–wake diaries. Applied Psychophysiology and Biofeedback, 29(4), 269–278. doi:10.100710484-004-0387-0 PMID:15707256 Kumar, S., Nagendra, H. R., Manjunath, N. K., Naveen, K. V., & Telles, S. (2010). Meditation on OM: Relevance from ancient texts and contemporary science. International Journal of Yoga, 3(1), 2. doi:10.4103/0973-6131.66771 PMID:20948894

344

 Special Yoga for Children and Young People With Special Needs

Ma, X., Yue, Z. Q., Gong, Z. Q., Zhang, H., Duan, N. Y., Shi, Y. T., Wei, G.-X., & Li, Y. F. (2017). The effect of diaphragmatic breathing on attention, negative affect and stress in healthy adults. Frontiers in Psychology, 8, 874. doi:10.3389/fpsyg.2017.00874 PMID:28626434 Mak, C., Whittingham, K., Cunnington, R., & Boyd, R. N. (2017). Mi Yoga: a randomised controlled trial of a mindfulness movement programme based on hatha yoga principles for children with cerebral palsy: a study protocol. BMJ Open, 7(7), e015191. doi:10.1136/bmjopen-2016-015191 PMID:28698326 Manjunath, N. K., & Telles, S. (2001). Improved performance in the Tower of London test following yoga. Indian Journal of Physiology and Pharmacology, 45, 351–354. PMID:11881575 Mcgraw, R. (2006). Adaptive Yoga for Individuals with Cerebral Palsy [Master of Science dissertation]. University of Illinois, Chicago, IL, USA. National Health Service. (2018). Mental Health of Children and Young People in England. 2017. https:// dera.ioe.ac.uk/32622/1/MHCYP%202017%20Summary.pdf Naveen, K. V., Nagendra, R. N. H., & Telles, S. (1997). Yoga breathing through a particular nostril increases spatial memory scores without lateralized effects. Psychological Reports, 81(2), 555–561. doi:10.2466/pr0.1997.81.2.555 PMID:9354111 Oppenheim, W. L. (2009). Complementary and alternative methods in cerebral palsy. Developmental Medicine and Child Neurology, 51, 122–129. doi:10.1111/j.1469-8749.2009.03424.x PMID:19740219 Peck, H. L., Kehle, T. J., Bray, M. A., & Theodore, L. A. (2005). Yoga as an intervention for children with attention problems. School Psychology Review, 34, 415–424. Pilkington, K., Kirkwood, G., Rampes, H., & Richardson, J. (2005). Yoga for depression: The research evidence. Journal of Affective Disorders, 89(1-3), 13–24. doi:10.1016/j.jad.2005.08.013 PMID:16185770 Porges, S. W. (2005). The vagus: A mediator of behavioral and physiologic features associated with autism. The Neurobiology of Autism, 2, 65–77. Porges, S. W. (2017). The pocket guide to the polyvagal theory: The transformative power of feeling safe. WW Norton & Co. Riedel, F. (1987). Long term effects of artificial ventilation in neonates. Acta Paediatrica (Oslo, Norway), 76(1), 24–29. doi:10.1111/j.1651-2227.1987.tb10409.x PMID:3551491 Ryan, R., O’Farrelly, C., & Ramchandani, P. (2017). Parenting and child mental health. London Journal of Primary Care, 9(6), 86–94. doi:10.1080/17571472.2017.1361630 PMID:29181091 Shaw, J. E. (2017). Introducing Biodynamic Craniosacral Therapy (BCST) into the landscape of jungian thought. Journal of Jungian Scholarly Studies, 12, 45–55. doi:10.29173/jjs27s Shea, M. J. (2007). Biodynamic craniosacral therapy. North Atlantic Books. Special Yoga. (n.d.). About us. https://specialyoga.org.uk/about-us/

345

 Special Yoga for Children and Young People With Special Needs

Spratt, E. G., Nicholas, J. S., Brady, K. T., Carpenter, L. A., Hatcher, C. R., Meekins, K. A., Furlanetto, R. W., & Charles, J. M. (2012). Enhanced cortisol response to stress in children in autism. Journal of Autism and Developmental Disorders, 42(1), 75–81. doi:10.100710803-011-1214-0 PMID:21424864 Streeter, C. C., Jensen, J. E., Perlmutter, R. M., Cabral, H. J., Tian, H., Terhune, D. B., Ciraulo, D. A., & Renshaw, P. F. (2007). Yoga Asana sessions increase brain GABA levels: A pilot study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 13(4), 419–426. doi:10.1089/acm.2007.6338 PMID:17532734 Streeter, C. C., Whitfield, T. H., Owen, L., Rein, T., Karri, S. K., Yakhkind, A., Perlmutter, R., Prescot, A., Renshaw, P. F., Ciraulo, D. A., & Jensen, J. E. (2010). Effects of yoga versus walking on mood, anxiety, and brain GABA levels: A randomized controlled MRS study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 16(11), 1145–1152. doi:10.1089/acm.2010.0007 PMID:20722471 Swani, P. (2015). Neurophysiological effects of yogic breathing techniques. International Research Journal of Management Sociology & Humanity, 6, 490–495. Taanila, A., Ebeling, H., Tiihala, M., Kaakinen, M., Moilanen, I., Hurtig, T., & Yliherva, A. (2014). Association between childhood specific learning difficulties and school performance in adolescents with and without ADHD symptoms: A 16-year follow-up. Journal of Attention Disorders, 18(1), 61–72. doi:10.1177/1087054712446813 PMID:22751677 Telles, S., Gerbarg, P., & Kozasa, E. H. (2015). Physiological effects of mind and body practices. BioMed Research International, 2015, 2015. doi:10.1155/2015/983086 PMID:26171397 The National Autistic Society. (2016). School report 2016. https://network.autism.org.uk/sites/default/ files/ckfinder/files/NAS%20SEN%20School%20report%202016%20FINAL.PDF Van de Weijer-Bergsma, E., Formsma, A. R., de Bruin, E. I., & Bögels, S. M. (2012). The effectiveness of mindfulness training on behavioral problems and attentional functioning in adolescents with ADHD. Journal of Child and Family Studies, 21(5), 775–787. doi:10.100710826-011-9531-7 PMID:22993482 Woodyard, C. (2011). Exploring the therapeutic effects of yoga and its ability to increase quality of life. International Journal of Yoga, 4(2), 49. doi:10.4103/0973-6131.85485 PMID:22022122

346

347

Chapter 20

Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder Pauline Jensen University of Sydney, Australia

ABSTRACT Research conducted in both the field of yoga and the field of behavioural disorders in children and adolescents leads to the speculation that the benefits of yoga practice demonstrated with respect to physiological, psychological, emotional, and psychosocial functioning may be applicable to the impairments evidenced in these areas in behavioural disorders. The intervention—20 weekly one-hour sessions of yoga—required a large commitment for the participants, their families, and the yoga instructor. The results suggested that yoga appears to exert its impact on stabilising the emotions (a secondary symptom) and reducing oppositional behaviour, frequently co-morbid (40%) with attention deficit hyperactivity disorder (ADHD). Given the limitations of this study, the results do indicate some significant changes in the behaviour of some of the boys with ADHD. In conclusion, yoga shows promise as a non-invasive, inexpensive, adjuvant treatment for boys with ADHD.

BACKGROUND Following working with children diagnosed or presenting symptoms of Attention Deficit Hyper Activity Disorder (ADHD) in the NSW Public School system for a couple of years; the present results are derived from what became a 12 year research project, investigating the effects of yoga on children and adolescents with attention deficit hyperactivity disorder (called ADHD) and other behaviour disorders. Children with ADHD have either predominant hyperactivity (externalising behaviour) or predominant inattentive behaviour (internalising behaviour) or the combined type. These behaviours are particularly disturbing or dysfunctional in the average classroom environment where expectations are to listen to the teacher and peers; sit still, quiet and attentive for extended periods of time; answer questions in the DOI: 10.4018/978-1-7998-3254-6.ch020

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

appropriate way; be organised with pencils, etc. and correct books; and apply oneself after a lesson has been taught. For most children these behaviours are readily acquired. So what makes it so hard for a child with ADHD? When working memory (necessary for accessing and manipulating information to solve problems), (Pliszka et al., 1996; Swanson et al., 2007) and when remembering instructions is difficult; when finding correct books, pencils etc. takes longer than expected; when remembering and understanding what the lesson was about; restlessness can manifest resulting in delays or refusal to get started on a set tasks, let alone completing any set tasks. Hyperactivity and impulsivity is correlated with high baseline levels of noradrenalin (Pliszka et al., 1996), resulting in difficulty with application to a singular task which requires focus and concentration. A lack of self-regulation (Greene & Ablon, 2001; Jarman, 1996; Porter, 1996; Barkley, 1996) is displayed as an inability to internalise speech which manifests as needing to verbalise all thoughts which has many implications in terms of classroom expectations and social mores. Where anxiousness and diagnosed anxiety co-exists, perhaps as a consequence of knowing that they are failing at school, continually ‘getting into trouble’ and negative interactions with peers, self-esteem can plummet. These challenging behaviours are not only present at school but at home, creating many difficulties for parents, siblings and the extended family where one parent may have had ADHD as a child or still be experiencing it (Cantwell, 1996). ADHD is often co-morbid with other disruptive behaviours including Oppositional Defiant Disorder (called ODD here) and the more severe Conduct Disorder (called CD here) (Biederman, et al., 1996; Bird, et al., 1988; Anderson et al., 1987; Jensen et al., 1997). These additional diagnoses can manifest in extreme argumentativeness and risk-taking behaviours such as fighting, stealing, lying, fire-lighting and animal cruelty. In NSW, students with these co-morbid disorders are removed from mainstream schools (with parental permission) and placed in behaviour schools or similar facilities where student/teacher ratios are around 7:1, with a teacher’s aide and a full time counsellor in the school but these facilities are in high demand and student’s tenure there is limited (Planning and Innnovation DET, 2006). Many of these students may eventually commit crimes and become part of the Juvenile Justice System (Kenny et al., 2008). The lifetime trajectory for these young people is obviously not good. While making many informal and formal observations of students with ADHD, ODD and CD and endeavouring to assist both students and teachers with managing these behaviours in the mainstream school setting, I began to wonder how practicing yoga may assist these students in self-management of behaviour; reduce impulsivity, hyperactivity, anxiety and anger; and build confidence and self- acceptance. Thus my research journey began. I was not an academic in the university system. I was a school teacher with a recently gained graduate diploma in Aboriginal studies and a Yoga Teacher Training certificate. During a Yoga Symposium organised by Satyananda Yoga in Sydney, NSW, I gleamed from the many presenting professionals that yoga was being applied in the various fields of medicine, psychology, education and philosophy. I made a commitment to pursue researching the effects of yoga on young people with ADHD and other behavioural disorders. I began by composing a research proposal which I sent to the Education Department and the Department of Behavioural Sciences at my closest University- University of Sydney. I was contacted by Dianna Kenny Ph.D. who showed interest in my proposal. She became my supervisor over the next 12 years in completing a Masters of Applied Science and a Ph.D.

348

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

FIRST STUDY –A PILOT STUDY Aims Ascertain whether ADHD children could be taught the yogic techniques to be used in a main study; assess the acceptance of the yoga program by parents and their children; assess compliance and dropout for a 9 week and a 20 week program of weekly yoga practice; and determine whether discernible reductions in inattentive, hyperactive and impulsive behaviour and the other behaviours included in the Conners’ Parent Rating Scale-Revised - Long (CPRS-R:L-R- L) (Conners, 1997) in children with ADHD could be obtained using yoga techniques.

Methods The yoga treatment protocol involved breathing exercises, asanas-body postures, concentration and relaxation exercises. Seven single case studies using an AB design were conducted during 1998 and 1999. Participants were recruited from an ADHD Support Group. Children aged 8 to 14 years, diagnosed with ADHD, together with a parent, were taught yoga weekly for nine weeks (n = 4) followed by a 20 week (n = 3) program taught by a qualified yoga teacher and engaged in daily home practice, which was recorded in a diary. Parental perceptions of their child’s behaviour were assessed using Conners’ Parent Rating Scales (CPRS-R:L-R- L) (Conners, 1997) at pre-(•--•) and post(▴—▴) test. Discussion on the data for the seven children, who were all taking stimulant medication are presented below.

Results Common improvements were seen in Anxious/Shy subscale, Oppositional subscale, Social problems subscale, Conners’ global index for restless/impulsive, Conners’ global index for emotional lability and Conners’ global index -total. These Results are presented in Figure 1. Figure 1 is an example of how results were presented. This example represents parental perceptions using the Connors parent rating scales of a 13 years old boy after attending 9 sessions of yoga and practising regularly at home at a maximum level of practice, demonstrating improvement in all but one subscale.

349

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

Figure 1. Changes in parental perceptions using the Connors Parent Rating Scales before and after 9 sessions of yoga

Connors Parent Rating Scales Opp = Oppositional; CoPr = Cognitive Problems; Hyp = Hyperactivity; An/Sh = Anxious/Shy; Pef = Perfectionism; Soc/Pr = Social Problems; Ps/So = Psychsomatic; A In = ADHD Index; Gl InR/I = Conners’Global Index Restless/Impulsive; Gl In Em. L = Conners’Global Index Emotional Lability; Gl In T = Conners’Global Index Total; D1V I = DSM1V Inattention; D1V H-I = DSM1V Hyperactive/ Impulsive; D1VT = DSM1V Total

Discussion and Conclusion The outcomes of this pilot study indicated that the aims were achievable. The participants were willing and able to practice the yoga program offered weekly and all made an effort to do home practice. Attendance for this group was satisfactory. The improvements seen on some of the Conners’ parent rating scales were encouraging. Parental perception using the CPRS-R:L-R’s for three of these subjects was favourable on most of the scales; for two there was improvement in some scales; and for one perceptions remained unchanged. Overall, the positive outcomes and results indicated that this study was worthy of replication on a larger scale. The children in the pilot study were a heterogeneous group of children differing in sex, age and number of sessions attended. Overall the main conclusion that can be drawn from these results are: 1. 2. 3. 4.

350

Children with ADHD can be taught yogic techniques; A yoga program is accepted by parents and children; Compliance of up to a 20 week yoga program is feasible; and From a parental perspective yoga is effective for some children and not for others.

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

SECOND STUDY: THE EFFECTS OF YOGA ON BOYS WITH ADHD This study examined the impact of yoga on the behaviour of boys aged 8-12 years, diagnosed with attention deficit hyperactivity disorder by paediatricians specialising in ADHD.

Aims To test a more rigorous study design; expand number of measures used; to compare a social skills program to a yoga program.

Methods The design of this study was a small randomised control trial involving partial cross-over in which boys (n = 16), stabilized on medication, were randomly assigned to either a yoga group (n = 6) (2 withdrew) for 20 weeks with weekly sessions offered or a cooperative games group (n = 7) for 20 weeks with 4 monthly sessions offered. After the first yoga trial was completed five boys from the control group entered the second yoga program of 20 weeks with weekly sessions, with two additional participants. A total of 11 boys completed up to 20 weeks of yoga therapy. Participants were boys diagnosed and treated by paediatricians at a Sydney Children’s hospital. Inclusionary criteria was based on choosing an homogenous group in relation to age range 8-12 years, gender- male, diagnosis ADHD (anxiety and learning problems accepted), English speaking, no overt family pathology, not undergoing other treatment (except ADHD medication), IQ within average range and parental cooperation

The Yoga Program The yoga program that was developed and successfully trialed in the pilot study was employed. It comprised of breathing practices, asanas (postures), concentration practices and yogic relaxation. The breathing practices involved respiratory training with selective use of oral and nasal passages for respiratory flow intended to increase the child’s breath awareness as well as train him to breathe naturally through both nards. Asanas involved stretching, load bearing, backward, forward, lateral flexion and extensions and inversions performed in sitting, standing, supine and prone positions. Relaxation training: involved becoming progressively aware of and relaxing body parts and tensing and relaxing muscles. Because this study involved young boys with attention, hyperactivity and impulsivity problems, the treatment program needed to cater for these problems. The program integrating many yoga techniques offer variety and activity. This requirement did not compromise the content of the program. Drawing upon many components of yoga was appropriate to ascertain overall outcomes and was also in partial accordance with the eight limbs of yoga advocated by Patanjali (Eliade, 1973). The program was taught by a qualified, experienced teacher of yoga. Participants were asked to practice the same techniques on a daily basis at home for at least 15 minutes per day which was recorded.

351

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

The Control Group Program The control group was taught and practiced cooperative games (as opposed to competitive games). Participants were encouraged to listen to each other and the teacher, follow instructions, take turns, establish and follow the rules and learn that working together resulted in having fun and a sense of belonging.

Measures All participants were pre- and post-tested on behavior rating scales using the Conners’ parent rating scale - revised: long version (CPRS-R:L) and the Conners’ teacher rating scale - revised:long version (CTRS-R:L); the Test of variables of attention (TOVA)- a continuous performance test and a motion logger the Actigraph- an objective measure of hyperactivity. Children underwent testing at time 1 (following stabilization on the medication or medication review), again at time 2 (at the end of the yoga training or control period) and again at time 3 for the cross over group.

Results Boys’ attendance at the yoga classes varied from 5 (this boy continued practice at home) to 19 sessions with a mean of 13.9 classes of the 20 scheduled yoga sessions. Mean days practiced (including classes) was 54.70 (SD = 40.95) ranging from 14 days to 116 days but quality and duration of practice at home could not be determined. Of the five sessions offered to the control groups, boys attended an average of three sessions. Preliminary analyses of the data indicated that despite randomization, there were significant differences between the control and yoga groups on pre-test subscale scores on almost all Conners’ rating scales for both parents and teachers. Further, because the standard deviations for some subscales were relatively large and because the data set was small, potential differences could be obscured by factorial (2x2) analysis. To reduce the degrees of freedom, a series of one–way repeated measures analyses of variance (ANOVA) were conducted on the outcome measures. Since the risk of an inflated type- I error was considered to be minimal, p was retained at 0.05 for this exploratory study. Results on the Conners’ parent rating scale (CPRS-R:L) – Conners’ global emotional lability subscale, indicated a significant main effect for time (F1,16 = 6.386, p = 0.022, Eta2 = .285); and a significant group by time interaction (F1,16 = 8.683, p = 0.009, Eta2 = .352). On the Conners’ global index total subscale, there was a significant main effect for time, (F1,16 = 12.263, p = 0.003, Eta2 = .434) and a significant group by time interaction (F1,16 = 7.294, p = 0.016, Eta2 = .313). On the Conners’ global restless/impulsive subscale, there was a significant main effect for time, (F1,16 = 6.058, p = 0.026, Eta2 = .275) and a trend for a group by time interaction (F1,16 = 2.165, p = 0.160, Eta2 = .119). On the oppositional scale there was a trend toward a main effect for time (F1,16 = 3.393, p = 0.084, Eta2 = .175; and there was a significant (borderline) group by time interaction (F1,16 = 4.649, p = 0.047, Eta2 = .225). All interaction effects favoured the yoga group except for Perfectionism subscale in which the control group interaction effects were significant. This result is not relevant to the study as perfectionism is not regarded as a primary or secondary symptom of ADHD. There were no significant improvements on any of the Conners’ teacher rating scale (CTRS-R:L) nor any significant effects on the Test of Variables of Attention (TOVA). The graphed data obtained from motion logger Actigraphs did not reveal any significant improvement in hyperactivity. The lack of improvement on the Conners’ teacher rating scale could be attributed to the 352

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

fact that all participants except one in the yoga group were on medication when the teachers rated their behaviour. By contrast, parents rated their sons’ behaviour when in a non-medicated state. The TOVA produced a number of invalid results and was not considered a reliable measure of attention and impulsivity in this study. There were many technical problems with the Actigraphs which could be attributed to being administered in a naturalistic setting and may have been better utilised in a clinical setting. Figure 2. Effect for Group By Time Interaction - CPRS-R:L - Global Index – Rest/Imp.

CONCLUSION The research conducted in both the field of yoga and the field of ADHD gives rise to some speculation that the benefits of yoga practice on neurophysiological, physiological, psychological, emotional and psychosocial functioning may be applicable to the impairments evidenced in these areas in ADHD. Although the measures used in this current study did not demonstrate overall significant changes to the core symptoms of attention deficit hyperactivity disorder (mentioned as ADHD below), the results suggested that, according to parental assessments, yoga appears to exert its impact on stabilising the emotions (secondary symptom) and reducing oppositional behaviour, frequently co-morbid (40 percent) with ADHD. The employment of more sensitive measures may indicate a more accurate assessment of outcome.

353

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

Figure 3. Effect for Group By Time Interaction - CPRS-R:L - Global Emotional Lability

The intervention – 20 weekly one-hour sessions of yoga - required a large commitment for the participants, their families and the yoga instructor. Greater compliance and closer monitoring of yoga practice could be achieved by conducting the sessions and practices within the school setting. Non-compliance with practice requirements may have, to some extent, compromised the results of the study. The current study’s integrated yoga program offered variety and activity, essential elements in catering for the needs of children who suffer from hyperactivity, impulsivity and inattention and who need to be highly motivated to stay on task. Although variables can be measured from an integrated yoga program and can be replicated, outcomes of specific practices can only be ascertained in studies designed to measure specific outcomes for example, the practice of specific yogic breathing techniques demonstrated changes in cerebral functioning in lateralisation, oxygen consumption and brain wave frequencies. As was demonstrated in the literature these parameters are affected in ADHD. This level of application may be appropriate in future studies. Several variables require measurement in order to make an accurate diagnosis of ADHD. To date there is no one diagnostic tool that reliably diagnoses the disorder. Therefore, several measures are currently being employed and others are being investigated. Those used in this study measured behavioural changes in contrast to the parameters, measured in many of the studies presented in the literature review. Although the results did not provide definitive evidence for the efficacy of yoga as an adjuvant treatment for ADHD, they demonstrate that for some children it was of benefit. The group results indicate, that compared to controls emotional lability significantly improved, oppositional behaviour indicated

354

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

borderline significant improvement and in restless/impulsive behaviour there was a trend toward significant improvement as observed by their mothers who were rating their boys when the effects of medication had dissipated. In addition the study demonstrated that ADHD children are capable of and willing to practice yoga on a regular basis. Figure 4. Effect for Group By Time Interaction - CPRS-R:L - Global Index Total

With larger numbers, more regular supervised practice and more reliable and specific measures, further studies may indicate that the changes observed in this underpowered study, may in fact be robust. Given the limitations of this study, the results do indicate some significant changes in the behaviour of some of the boys with ADHD. In conclusion yoga shows promise as a non-invasive, inexpensive, adjuvant treatment for boys with ADHD and further studies are warranted from the findings reported here.

355

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

THIRD STUDY: YOGA AS AN ADJUVANT THERAPY FOR STUDENTS ENROLLED IN SPECIAL SCHOOLS FOR DISRUPTIVE BEHAVIOUR Aims and Design This controlled field study, using a within and between repeated measures design examined the impact of yoga on the behaviour of students aged 8-16 years, enrolled at special schools for disruptive behaviour with the New South Wales Department of Education, Australia. Seventy-eight participants (five female) enrolled in the study, sixteen students acted as their own controls, fifty–five participated in yoga intervention only and seven were in the control condition only. Altogether seventy-one (71) students participated in the yoga intervention and twenty-three (23) in the control condition.

Methods The yoga intervention, a 13-week comprehensive program consisting of two to three 30-40 minute sessions per week, was taught by a qualified, experienced yoga teacher who was also a specialist teacher for behaviourally disordered students (PSJ). The control group experienced the standard school program provided by the special school. Control and yoga participants were pre- and post-tested on the Conners’ teacher and parent rating scales–revised long version (CTRS-R: L, CRRS-R: L), the Test of everyday attention for children (TEA-Ch), the Trait component of the state-trait anxiety inventory (STAI) or the state-trait anxiety inventory for children (STAIC) and the self description questionnaire I or II (SDQI & II). Behaviour observations were conducted using the behaviour assessment system for children-portable observation program (BASC-POP) in both the classroom (for control and yoga groups) and the yoga classes by blind independent observers and by the main researcher. School staff wrote comprehensive daily notes, in yoga classes and recorded students’ on- and off-task behaviours. Measures designed by the researchers consisted of the feelings faces scale (FFS) that was completed after the last yoga class for the week by all students; a yoga survey (YS) requesting student perceptions of the benefits of yoga and individual assessments of yoga competence (IAYC) that were completed at the end of the yoga intervention by two subgroups. A physical, emotional and mental states (PEMS) measure was administered pre- and post-yoga sessions for a subgroup. Breathing patterns before, during and after the yoga relaxation session were recorded using Respiratory Inductive Plethysmography (RIP) bands in a subgroup and compared with three young people without disruptive behaviour. Results were analysed using the general liner Model for all pre- and post-test measures. Mean scores were calculated for the FFS, the yoga survey and the IAYC. Visual analysis of the RIP results was conducted by researchers.

Results Of 71 participants in the yoga group 12 (16.9%) attended from 7 to 10 classes; 36 (50.5%) attended from 11 to 20 classes and 23 (32.5%) attended from 21 to 35 classes. Total absences from the yoga classes (39.76%) were due to sickness and truancy (32.35%); lack of interest, (45.71%); work experience, home school visits or other school programs, (15.89%); and suspension from school (6.04%). Of the 33 students in the control group, 10 (32.35%) discontinued due to leaving the school (n = 6) or truancy (n = 4); 23 (67.65%) remained in the control group.

356

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

Major findings were as follows: On the Conners’ teacher rating scales revised-: long version (CTRS-R: L), significant improvement over time was found for the yoga group (n = 64) in the Oppositional subscale. No other significant changes were seen over time or in group by time interactions for the yoga (n = 64) or the control groups (n = 20). On the Conners’ parent rating scales revised long version (CRRS-R: L), significant improvements over time were seen in ten out of fifteen subscales for the control group (n = 10) and deterioration in six subscales for the yoga intervention (n = 16). Group by time interaction, favouring the control group was seen in thirteen subscales. Significant improvements on the Test of Everyday Attention for Children, (TEA-Ch) were seen on two subtests of focused attention and two subtests of sustained attention (one borderline) for the yoga group and two subtests of focused attention, two subtests of sustained attention and one of switching attention for the control group. No significant changes were observed on the State-Trait Anxiety Inventory (STAI) or the State-Trait Anxiety Inventory for Children (STAIC) nor on the Self Description Questionnaire I or II SDQ (I & II) but pre-test levels were within normal limits. Subgroup analysis of the CTRS-R:L of students who participated in over 20 classes (n = 14) indicated more pronounced significance on the oppositional subscale. On the Behaviour Assessment System for Children-Portable Observation Program, (BASC-POP) significant group by time interaction reductions were observed in oppositional behaviour with a trend in hyperactive behaviours favouring the yoga group in the classroom. Over time, the yoga group’s (n = 19) classroom behaviours indicated significant reductions in impulsive behaviour and borderline reductions in hyperactivity and total ADHD behaviours. Numbers assessed on this measure were reduced due to one rater proving unreliable (and whose ratings were discarded) and due to technical problems. The control group (n = 16) showed no significant changes in classroom behaviours. Subgroup analysis of the BASC-POP for students who acted as their own controls (n = 8) indicated significantly less ADHD behaviours in yoga classes at the end of the program compared with all other assessment times. In the yoga classes (n = 21) at the beginning of the intervention ADHD behaviours were 33% of classroom behaviours compared with 25 percent at the end (n = 20) of the intervention. Staff observations of yoga classes indicated ontask behavioural descriptors outnumbered off-task descriptors by approximately 4:1. Weekly self-reports on the Feelings Faces Scale (n ≤ 35), immediately after yoga each week, showed an overall positive response in mood, enjoyment of the program, and confidence in yoga practice. Self-report on the physical, emotional, mental states measure, showed significant positive changes in physical, emotional and mental states from the beginning of yoga sessions to the end of sessions in a subgroup of students (n ≤ 13). The Yoga survey indicated benefits for 63 percent to 80 percent of the respondents (n = 27) in six out of the seven items. On the individual assessment of yoga competence students (n = 11) scored a mean of 79.64 percent (SD = 9.44). Breathing patterns, for students with disruptive behaviour (n = 7), indicated greater stability during the relaxation compared with before and to a lesser degree after the relaxation but were not as stable as the breathing patterns of students without disruptive behaviour (n = 3) throughout the testing period.

357

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

Figure 5. Changes in self report emotional state pre and post yoga classes

Figure 6. One minute screen shots of breathing effort pre, during and post yognidra for individual student with conduct disorder, learning disorder and emotional disorder

358

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

Table 1. Responses of the yoga survey Question

N

Yes

%

No

%

1. Do you have more energy?

27

20

74.1

7

25.9

2. Are you a calmer person?

27

18

66.7

9

33.3

3. Are you more flexible?

27

21

77.8

6

22.2

4. Do you feel stronger?

25

20

80

5

20

5. Are you able to relax easier?

27

17

63

10

37

6. Do you have less mood swings?

26

12

46.2

14

53.8

7. Do you experience as much anger?

27

10

37

17

63

DISCUSSION Yoga as an intervention for students enrolled at behaviour school appears to have immediate positive effects as perceived by students immediately after sessions, in observations of behaviour during the yoga class, in assessed ability during a yoga class and in the stabilizing effects on breathing effort during relaxation. Collecting data on a regular basis appears to be a method of overcoming spasmodic attendance and early withdrawal. Few significant results were found on standardized measures. Results on these tests were affected by a number of methodological issues such as (i) fluctuations in attendance, (ii) withdrawals from the program weeks before post-program assessments, and (iii) the intervention not being long and intense enough for parents and teachers to perceive significant changes in the environments in which the students had been ‘acting out’ for most of their childhood.

REFERENCES Anderson, J. C., Williams, S., McGee, R., & Silva, P. A. (1987). DSM-III disorders in preadolescent children: Prevalence in a large sample from the general population. Archives of General Psychiatry, 44(1), 69–76. doi:10.1001/archpsyc.1987.01800130081010 PMID:2432848 Barkley, R. A. (1996). The North American perspective on attention deficit hyperactivity disorder. The Educational and Developmental Psychologist, 13(1), 2–23. doi:10.1017/S0816512200027358 Biederman, J., Faraone, S. V., Milberger, S., Jetton, J. G., Chen, L., Mick, E., Greene, R. W., & Russell, R. L. (1996). Is childhood oppositional defiant disorder a precursor to adolescent conduct disorder? Findings from a four-year follow-up study of children with ADHD. Journal of the American Academy of Child and Adolescent Psychiatry, 35(9), 1193–1204. doi:10.1097/00004583-199609000-00017 PMID:8824063 Bird, H. R., Canino, G., Rubio-Stipec, M., Gould, M. S., Ribera, J., Sesman, M., ... Moscoso, M. (1988). Estimates of the prevalence of childhood maladjustment in a community survey in Puerto Rico: The use of combined measures. Archives of General Psychiatry, 45, 1120–1126. doi:10.1001/archpsyc.1988.01800360068010 PMID:3264147

359

 Yoga as an Intervention for Students With Attention Deficit Hyperactivity Disorder

Cantwell, D. P. (1996). Attention deficit disorder: A review of the past 10 years. Journal of the American Academy of Child and Adolescent Psychiatry, 35(8), 978–987. doi:10.1097/00004583-199608000-00008 PMID:8755794 Conners, C. K. (1997a). Conners rating scale -revised. Multi-Health Systems Inc. Conners, C. K. (1997b). Conners rating scales- revised manual. Multi-Health Systems Inc. Eliade, M. (1973). Yoga immortality and freedom. Bollingen Series LVI. Greene, R. W., & Ablon, J. S. (2001). What does the MTA study tell us about effective psychosocial treatment for ADHD? Journal of Clinical Child Psychology, 30(1), 114–121. doi:10.1207/S15374424JCCP3001_13 PMID:11294069 Jarman, F. C. (1996). Current approaches to management of attention deficit hyperactivity disorder. The Educational and Developmental Psychologist, 13(1), 46–55. doi:10.1017/S0816512200027395 Jensen, P. S., Martin, D., & Cantwell, D. P. (1997). Comorbidity in ADHD: Implications for research, practice, and DSM-V. Journal of the American Academy of Child and Adolescent Psychiatry, 36(8), 1065–1079. doi:10.1097/00004583-199708000-00014 PMID:9256586 Kenny, D. T., Nelson, P. K., Schreiner, I., Lennings, C. J., & Butler, T. (2008). Young offenders on community orders: Health, welfare and criminogenic needs. Sydney University Press. Planning and Innnovation. (2006). Statistical Bulletin. NSW Department of Education and Training. Pliszka, S. R., McCRACKEN, J. T., & Maas, J. W. (1996). Catecholamines in attention-deficit hyperactivity disorder: Current perspectives. Journal of the American Academy of Child and Adolescent Psychiatry, 35(3), 264–272. doi:10.1097/00004583-199603000-00006 PMID:8714313 Porter, L. (1996). Student behaviour. Theory and practice for teachers. Allen and Unwin. Swanson, J. M., Kinsbourne, M., Nigg, J., Lanphear, B., Stefanatos, G. A., Volkow, N., Taylor, E., Casey, B. J., Castellanos, F. X., & Wadhwa, P. D. (2007). Etiologic subtypes of attention-deficit/hyperactivity disorder: Brain imaging, molecular genetic and environmental factors and the dopamine hypothesis. Neuropsychology Review, 17(1), 39–59. doi:10.100711065-007-9019-9 PMID:17318414

360

361

Chapter 21

Yoga:

A Multi-Dimensional Therapeutic Approach to Autism Spectrum Disorder Kankan Gulati National Institute of Mental Health and Neurosciences, Bengaluru, India Praerna H. Bhargav National Institute of Mental Health and Neurosciences, Bengaluru, India Shalu Elizabeth Abraham National Institute of Mental Health and Neurosciences, Bengaluru, India Hemant Bhargav National Institute of Mental Health and Neurosciences, Bengaluru, India

ABSTRACT Autism spectrum disorder (ASD), a neurodevelopmental disorder, manifests as impairment in social communication an interaction with restrictive and repetitive patterns of behaviour. Yoga therapy, a mind-body intervention, employs a multi-dimensional approach to reduce psychological distress and bring balance and harmony at the levels of body, breath and mind through physical postures, breathing practices, chanting, and relaxation techniques, respectively, thus enhancing overall well-being. Various yoga studies have shown promise in improving symptoms of ASD by improvement in sensory processing, gross motor skills, balance and coordination, cognition, imitation skills, and the ability to connect in relationships. This chapter aims to provide an overview of the potential role of Yoga therapy in the management of ASDs with emphasis on future standardized yoga trials with robust methodology and long-term follow-ups to establish the clinical utility of Yoga therapy for the same. Also, a tentative yoga lifestyle module for ASD with necessary contra-indications and practical tips has been provided.

DOI: 10.4018/978-1-7998-3254-6.ch021

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga

INTRODUCTION Autism Spectrum Disorder (ASD), previously termed as pervasive developmental disorder (Towbin, 2005), is a phenotypically heterogeneous group (Jeste & Geschwind, 2014) of neurodevelopmental disorders (Livingston & Happé, 2017). Prior to DSM V (American Psychiatric Association, 2013), ASD was conceptualised as five discrete disorders including Autistic Disorder, Asperger’s Disorder, Pervasive Developmental Disorder Not Otherwise Specified (PDD-NOS), Childhood Disintegrative Disorder and Rett’s Disorder (Phetrasuwan et al., 2009; American Psychiatric Association, 2013). The symptoms are identified in early childhood typically during the second year of life (Wetherby et al., 2004) and persists throughout adult life (Thapar et al., 2017). Recent clinical consensus has shifted the conceptualisation of the disorder into a continuum model in which heterogeneity of the symptoms is recognised as inherent within the disorder. The core diagnostic impairments are collapsed into two main domains: (1) Impairment in social communication and interaction (Criteria A); (2) Restrictive repetitive patterns of behaviour (Criteria B). Aberrant language development is no longer considered as a defining criterion (Grzadzinski et al., 2013).

PREVALENCE Autism Spectrum Disorder with a prevalence rate of over 1percent worldwide (Essa & Qoronfleh, 2020) diagnosed four times more often in boys than in girls (Fombonne, 2006). The latest ASD prevalence estimate, as measured by the Autism and Developmental Disabilities Monitoring (ADDM) Network, is 18.5 per 1,000 children aged 8 years in 2016 which is 175 percent higher than (2.8 times) the first estimates reported by the ADDM Network in 2000 and 2002 (Centres for Disease Control and Prevention, 2016). Another study done on children aged between 0-17 years, diagnosed with ASD, from South Asian countries including Bangladesh, India, Sri Lanka, informed that the percentage prevalence rate ranged from 0.09 percent to 1.07 percent in these countries (Hossain et al., 2017).

CAUSES ASD is considered to be a multifactorial disease (Akshoomoff et al., 2002) and includes various genetic, biological and environmental risk factors (Hallmayer et al., 2011).

Genetic Factors Recent neuroimaging studies on ASD has helped to evolve a better understanding of the genetic basis of brain development and also the structural and functional abnormalities associated with it (Berg & Dobyns, 2015). Previous studies on twins and families suggest that genetic factors play a dominant role, like a child having a sibling with ASD is reported to be at higher risk of developing ASD (Korvatska et al., 2002; Spence, 2004). Genetic defects give rise to abnormal molecular brain growth factors causing developmental dysregulation and also result in anatomical abnormalities (Akshoomoff et al., 2002). It has been found that individuals with genetic or chromosomal conditions like fragile X syndrome or tuberous sclerosis are more susceptible to develop ASD (Cohen et al., 2005; Hall et al., 2008). 362

 Yoga

Biological Factors Brain Development Previous neuroimaging studies provide evidence that compartmental volume of the various sub-regions of the brain follows an unconventional development pattern in ASD (Ecker et al., 2012; Raznahan et al., 2010; Wallace et al., 2010). As a child with ASD matures into an adult, the regional and whole-brain volumes change dynamically (Lange et al., 2015). Many MRI studies have reported that children between the age of 2 - 4 years with ASD and without any psychiatric diagnosis have larger brain volume as compared to healthy controls (Courchesne, 2002; Courchesne et al., 2001; Carper & Courchesne, 2005) which seems to fade away by the age of 6-8 years (Courchesne et al., 2001), after that there is no considerable increase and possibly a decrease in the brain volume at an older age (Courchesne et al., 2011). The changes can be more evidently seen in the frontal and temporal lobes as compared to the parietal and occipital lobes (Carper & Courchesne, 2005). However, one latest study findings challenge the “normalization” prediction of the brain overgrowth hypothesis by demonstrating that brain enlargement persists across childhood into early adulthood (Yankowitz et al., 2020).

Neurotransmitters Neurotransmitters are neuronal signalling molecules which play a key role in adequate brain development. The important functions of neurotransmitters include maintaining memory, learning, behaviour and motor activity (Choudhury et al., 2012). In early developmental phases, neurotransmitters help in modulating various processes like the neuronal migration, differentiation, growth cone motility and synaptogenesis (Haydon et al., 1987; Kwong et al., 2000). Imbalances in the level and regulation of various neurotransmitters such as Glutamate (Nicoletti et al., 2011), GABA (gamma-aminobutyric acid) (Gajcy et al., 2010; SidAhmed‐Mezi et al., 2010), Glycine (Smith et al, 2019), Norepinephrine (Ressler & Nemeroff, 1999), Dopamine (Lud Cadet et al., 2010), Serotonin (Arango et al., 2003; Oquendo & Mann, 2000) and Acetylcholine (Greenberg et al., 2010) can lead to various developmental disorders including ASD.

Age at Conception Some studies support the evidence that children born to older parents are at higher risk for developing such disorders (Durkin et al., 2008).

Environmental Factors Some preliminary evidence suggest the link between consumption of thalidomide and anticonvulsants by the mother during pregnancy and future risks of developing ASD in children (Christensen et al., 2013; Strömland et al., 1994). Also, some findings suggest that certain time period such as those before, during and immediately after birth is critical in the development of ASD (Gardener et al., 2011).

363

 Yoga

CLINICAL SYMPTOMS Impaired Social Interaction Deficits in social interaction are the core underlying feature of ASD (White et al., 2007) that manifests from early years of life (Kanner, 1943) and have been recognized as a major cause of impairment (Carter et al., 2005). Individuals with ASD have a lack of “theory of mind” (difficulty in making attribution about intentions and motivation) which produce difficulties interpreting the social behaviours of others (Sadock et al., 2009). Symptoms include lack of social interest or smiling, deficiency in social responsiveness, preferring to be alone (Zwaigenbaum et al., 2007), reduced eye contact, the paucity of response to their own name and limited social engagement (Baranek, 1999). The symptoms show no remission even with ageing. In fact, the propensity of social deficits adds to an increased level of distress as these children approach adolescence, facing increased & expanding social demands (White et al., 2007).

Underdeveloped Communication Skills Previous researches suggest that in young children with ASD, verbal development is predicted by both verbal and nonverbal skills such as joint attention, play development, imitation and motor skills (which remain under-developed) (Paul et al., 2008; Thurm et al., 2007; Charman et al. 2003; Charman et al. 2005; Weismer et al., 2010; Manwaring et al., 2017; Stone & Yoder 2001). Usually delayed language milestones concern the parents to seek clinical consultation (Bolton et al., 2012; De Giacomo & Fombonne, 1998; Richards et al., 2016). Verbal communication impairment involves difficultly in initiating and sustaining a conversation such that even normal conversations like the exchange of ideas become challenging for them. Non-verbal communications such as maintaining eye contact, body language, gestures and facial expressions during social interaction are also not up to the mark (Eigsti et al., 2011).

Behavioural Abnormalities Another core feature of ASD is restricted, repetitive and stereotyped behaviour (Asperger, 1944; Kanner, 1943). Restrictedness implies to the sameness of interests and activities, narrow focus and the demand to maintain a consistent environment. Repetitive motor actions, words or phrases, routines and rituals are also a common feature (Leekam et al., 2011).

Associated Physiological Dysfunction – Gastrointestinal Dysfunction Amongst physiological dysfunction, gastrointestinal dysfunction is commonly found in children with ASD. Approximately 9 to 70 percent of the children with ASD suffer from gastrointestinal dysfunctions like gastrointestinal reflux, diarrhoea, constipation and abdominal pain (Adams et al., 2006; Black et al., 2002; Horvath & Perman, 2002; Levy et al., 2007; Molloy & Manning Courtney, 2003; Nikolov et al., 2009; Valicenti-McDermott et al., 2006). Some develop even more severe symptoms such as lymphoid nodular hyperplasia, ulcerative colitis or hiatal hernia (Ming et al., 2008). A strong positive correlation has been found between autism severity and severity of gastrointestinal dysfunction i.e., children with more severe symptoms of ASD are likely to have more severe gastrointestinal symptoms and vice-versa. It is very much possible that the gastrointestinal problems, to some extent, aggravate the symptoms of 364

 Yoga

ASD (Adams et al., 2011; Tomova et al., 2015). The causes of increased gastrointestinal problems in ASD are beyond the scope of this chapter.

Sleep Difficulties Another common problem faced by autistic children is disturbance of sleep (Sotoodeh et al., 2017) with prevalence rate ranging from 44 percent to 83 percent of the total population. Complaints related to sleep include difficulty initiating sleep, restlessness during sleep, not falling asleep in one’s own bed, frequent waking-ups and decreased sleep during night-time (Gail et al., 2004). These sleep problems may act as predictors of the intensity of autistic behavioural issues during the day time (Schreck et al., 2004).

AVAILABLE TREATMENT Available conventional treatment aims at managing the core symptoms affecting the quality of life in such individuals (Broadstock et al., 2007). The treatment also depends on physical, behavioural and associated co-morbid conditions (Kerbeshian et al., 2001; Larkin, 1997) that may include mental retardation, neurological conditions such as seizure disorder (Volkmar & Nelson,1990), Tourette syndrome (Kerbeshian & Burd, 1996) and other psychiatric conditions like bipolar disorder, depression (Korkmaz, 2000), obsessive–compulsive disorder (Russell et al., 2013) and schizophrenia (Stahlberg et al., 2004). Pharmacological treatment predominantly involves the use of antipsychotics, antidepressants, stimulants, lithium and anxiolytics (Kerbeshian et al., 2001). But these medications lack the ability to bring a complete cure, and may cause some biological and psychological side effects such as weight gain (Hellings et al., 2001; McDougle et al., 1998), fatigue, lethargy (Remington et al., 2001), vomiting, excessive sedation (Campbell et al., 1993), skin rashes (Hellings et al., 2005) and even severe tantrums, screaming, hyperactivity and aggression (Kern et al., 2002). Thus, often the families of children seek other non-pharmacological treatments. Commonly explored options are: nutritional or diet therapy, probiotic supplementation, psychosocial interventions and yoga therapy (Levy & Hyman, 2005; Harrington et al., 2006; Hanson et al., 2007; Mari-Bauset et al., 2014).

Diet Therapy The nutritional or dietary intervention involves the adoption of elimination diets and supplementation interventions (Alpert, 2007; Elder et al., 2015). An elimination diet consists of a gluten-free diet and/or casein-free diet for ASD. It basically involves eliminating all the sources of gluten (a mixture of proteins found in wheat, oats, barley, or rye) and casein (a peptide commonly found in milk) from the daily diet of a person with ASD (Ly et al., 2017). Researches on the elimination diet reveal mixed results in terms of its efficacy (Hyman et al., 2016; Johnson et al., 2011; Elder et al., 2006; Knivsberg et al., 2002; Whiteley et al., 2010). Also, the cost-effectiveness and feasibility further need to be assessed as previous literature mentioned an attrition rate of up to 25 percent in short-term trials (Ly et al., 2017). Other limitations include stigmatization, diversion of treatment resources and nutritional deficiencies in children following an elimination diet (Hediger et al., 2008). In a study, it was observed that seventy-eight per cent of the 12 percent autistic children (who were on elimination diet) were made to ingest supplements in excess (Stewart et al., 2015). This can lead to its own repercussions. For e.g., excess of Vitamin A can 365

 Yoga

lead to problems like nausea, headache, fatigue, loss of appetite, psychiatric changes, osteoporosis, and irritability (National Institutes of Health, & National Institutes of Health, 2011).

Probiotic Supplementation Another popular therapeutic intervention is manipulating the gut microbiome using probiotic supplements. In children with ASD, gut microbiome profiles were strongly correlating with the severity of neurological and behavioural symptoms (Adams et al., 2011; Chaidez et al., 2014; Liu et al., 2016). Therefore, microbiome manipulation seems to be an emerging therapeutic modality for this disorder (Liu et al., 2019). Though studies have reported optimized cognitive development and improvement in behaviour using probiotic supplementation but conclusive evidences are lacking (Rianda et al., 2019; Liu et al., 2019). Also, some minor side-effects such as diarrhoea, bloating, abdominal cramps and skin rashes have been recorded in children with ASD using probiotics (Shaaban et al., 2017).

Psychosocial Interventions Psychological and psychosocial interventions for ASD aim to treat the core and associated symptoms (DeFilippis & Wagner, 2016). These predominantly include therapeutic techniques such as Applied Behaviour Analysis (ABA), Cognitive behaviour therapy (CBT) and social cognition training. All these have been found to be useful to some extent but these therapies also have their own limitations in terms of difficulty with adherence and sustained irreversible improvement.

YOGA THERAPY Yoga, an ancient Indian science, (Telles & Naveen, 1997) has evolved in last few decades as a popular mind-body intervention with a significant role in the promotion of physical and mental health (Jain & Talukdar, 1993). Yoga therapy is being defined as “the use of yoga practices for the prevention and treatment of medical conditions” (Stephens, 2017). Various physiological and psychological benefits have been recorded in the past such as regulation of blood glucose levels, blood pressure, enhanced alertness, better mood, improvement in subjective well-being, self-acceptance and promotion of positive feelings with the decline in the levels of aggression, depression and anxiety (Ross & Thomas, 2010; Field, 2016). Also it is helpful in achieving & maintaining optimum health of cardiovascular, digestive, endocrine and excretory systems (Bhavanani, 2014). Yoga therapy is a holistic approach which targets healing at the five levels of existence (Nagendra, 2013) that are explained in a Vedic text called the Taittiriya Upanishad. The five levels of existence comprise of the physical layer or Annamaya kosha, the layer of vitality or Pranamaya kosha, the layer of emotions or Manomaya kosha, the layer of inner wisdom or Vijnanamaya kosha and the layer of bliss or Anandamaya kosha. All these layers affect each other and the overall health of an individual (Or-Bach, 2018). Yoga therapy involves the use of a combination of practices to harmonize and balance each layer of existence. For e.g., asanas (bodily postures) for Annamaya kosha, pranayama (breath control and techniques) for Pranamaya kosha, dhyana (meditation/chants) for Manomaya kosha, jnana (knowledge/ counselling) for Vijnanamaya kosha and karma yoga (working in blissful awareness/forgiveness, compassion and letting go) for the Anandamaya kosha (Nagarathna & Nagendra, 2008). 366

 Yoga

Yoga, which means “to connect”, can be a promising psycho-social therapy for such children. It employs the use of various poses, breathing practices and visualizations which can be accomplished by the children themselves (or with the help of the parents) and helps in improving sensory processing, gross motor skills and the ability to communicate and connect in relationships. In contrast to other therapies, it helps in letting such children calm by themselves and reduces dependence on someone else to provide this comfort (Radhakrishna et al., 2010; Gulati et al., 2018). Figure 1 describes the holistic approach of yoga therapy towards the five levels of existence for children with ASD.

Therapeutic Role of Yoga in ASD Yoga has shown to improve symptoms of ASD (Hekmat Pou et al., 2016) and reduce the psychological distress of individuals with autism (Deorari & Bhardwaj, 2014). In the past, preliminary researches have been conducted to probe the effect of yoga on scales of sensory dysfunction, behavioural problems and impaired communication with beneficial results (Artchoudane et al., 2019). Yoga brings improvement in the overall functioning of the brain which in turn helps in improving the overall behaviour (Radhakrishna et al., 2010).

Behavioural Issues Yoga being a positive therapeutic intervention, may help in reducing all the negatives such as stress, feelings of worry, irritable behaviour (Rosenblatt et al., 2011), lethargy, social withdrawal, hyperactivity and non-compliance and may bring positivity by upgrading mood (Litchke et al., 2018) and inducing a feeling of calmness in autistic children and their parents (Radhakrishna et al., 2010; Koenig et al., 2012; Hourston & Atchley, 2017; de Bruin et al., 2015; Ridderinkhof et al., 2018). Thus, it may help in promoting a reduction in maladaptive and disruptive behaviour and increasing positive behavioural responses (Milton et al., 2019) in children with autism (Koenig et al., 2012; Lee, 2017). Autonomic dysfunction is associated with ASD (Schumann et al., 2009) with an increased sympathetic tone and reduced parasympathetic activity (Ming et al., 2011). Three months of yoga therapy has shown to increase parasympathetic activity and decrease the sympathetic activity of the nervous system in children with autism (Vidyashree et al., 2019). Also, yoga practice has been found to reduce self-stimulatory behaviour (Nevola, 2017) and elevating functional performance in this population (Koenig et al., 2012). The core symptoms of autism which include peculiar, odd, asocial and inappropriate behaviour showed improvement following a combined intervention comprising of yoga, dance and music therapy in latency-aged group autistic children (Rosenblatt et al., 2011). Also, yoga is reported to be beneficial in not only reducing anxiety but also stress, tension and headache (Kenny, 2002) in these children. A recent study assessed the effect of yoga therapy on sensory processing disorder in 21 children with ASD between 6 to 20 years of age. It was found that 2 weeks of yoga intervention given for 90 min a day, for 5 days/week, resulted in a reduction in aggression, increase in calmness, improved behavior & increased response to verbal instructions as reported by the parents & teachers of the participants. Although the study did not show any significant changes in the primary variables, probably due to the short duration of intervention but was able to bring out noticeable improvements in behavioral aspects of such children with an intervention of duration as short as two weeks (Gulati et al., 2018).

367

 Yoga

Figure 1. Yoga Therapy practices for all the five vital levels of existence in children with ASD

Social Interaction The practice of yoga may boost the ability to maintain eye contact in verbal (Radhakrishna, 2010; Kaur & Bhat, 2019) as well as non-verbal children (Kenny, 2002). Relationship and communication of such children with their family members improved considerably by active involvement and participation of other family members in practising yoga together (Garcia et al., 2019). This remained not just confined to family, as the practice of yoga stimulated the overall social and emotional growth by the promotion of expression of emotions and enhancement of the feeling of empathy towards others (Litchke et al., 2018). Multimodal Mandala yoga practice comprising of group poses, colour and tracing sheets, rhythmic chanting and yoga games (delivered to children with autism on an hourly basis, twice a week for 4 weeks), showed enhanced mood and emotional expression increased empathy toward others and

368

 Yoga

improved teamwork skills. The study concluded that multimodal Mandala yoga training has implication for developing positive social and emotional skills. Thus, improvement in all these abilities put together may improve the overall socializing ability and quality of life of an autistic child (Sotoodeh et al., 2017).

Communication Skills The relationship between movement and language development is well established in child development (Jensen, 2000) as it improves general cognition of a child. Spontaneous and complex changes in speech have been reported by speech pathologists when a child practices physical movements. When physical movement was combined with language learning it strengthens recalling ability and supplements the use of new and correct language (Kenny, 2002). Also, the positive association between movement before an academic task and on-task behaviour implicates that yoga postures practised before an academic task may also improve on-task behaviours, potentially (Miramontez & Schwartz, 2017). Therefore, yoga therapy comprising of physical movements in the form of asanas, may serve as a tool to enhance communication skills and language in autistic children (Radhakrishna et al., 2010; Lee, 2017). Reports also suggest that comprehension and verbal reactivity to spatial commands also increased post yoga intervention (Connor, 2005) along with verbal reception and expression (Radhakrishna et al., 2010). Yoga therapy also improved various features of autism including the ability of a child to carry out routine activities, to pay heed to visual schedule, verbal identification and motor imitation of poses from picture cards, initiation and engagement in play routines which required social interaction (Kenny, 2002; Connor, 2005).

Cognitive Issues Deficits in sensory integration lead to low levels of concentration and attention in people detected with ASD (Simonoff et al., 2008). Sensory integration is a neurological process that helps in organizing & processing information received from the environment (Kranowitz & Archer, 1998) in order to produce suitable responses (Ayres, 1995). One of the important cognitive features found to be impaired in children with ASD is sustained attention, which is essentially required to engage in a particular task (Christakou et al., 2013). Evidence show that yoga therapy may address & correct these core symptoms i.e. attention (Behar, 2006; Koenig et al., 2012; Koterba,2007; Peck et al., 2005; Porter, 2013; Scroggins et al., 2012), concentration, focus, ability to organize (Oldenberg, 2007), alertness (Radhakrishna et al., 2010), sensitivity (Ramanathan et al., 2019), sensory integration, sensory processing (Peck et al., 2005) and sensory defensiveness (Goldeberg, 2004) ability to engage with others (Koenig et al., 2012; Koterba, 2007; Scroggins et al., 2016; Radhakrishna, 2010; Radhakrishna et al., 2010) at school, (Koenig et al., 2012; Peck et al., 2005; Porter, 2013; Ehleringer, 2010; Goldberg, 2004) at home, (Scroggins et al., 2016; Radhakrishna et al., 2010; Ehleringer, 2010) and in the community as well. With better engagement, the team-work skills (Litchke et al., 2018) and on-task performance may also improve with yoga (Miramontez & Schwartz, 2017). Significant reduction in auditory and visual reaction time has been observed in children post practise of yoga, denoting faster reactivity, though there is no evidence of slowing of reaction time in ASD (Ferraro, 2016). Reaction time is the time taken to respond to a stimulus and helps in assessing the quality and speed of information processing (Moradi & Esmaeilzadeh, 2017). Thus faster reactivity means faster

369

 Yoga

processing of information which can be accredited to increased alertness and concentration as a result of yoga practice (Ramanathan et al., 2019).

Sensory-Motor Skills Yoga therapy may help in developing body awareness and improving sensory-motor skills (Behar, 2006). The sense of proprioception helps in processing information about body position received through the muscles, ligaments and joints. The vestibular system that helps in maintaining balance, (Day & Fitzpatrick, 2005) is also responsible for regulating attention/concentration and behavioural stability (Brewster et al., 1999). The practice of yogasanas increases vestibular and proprioceptive stimulation and makes one more aware of the body movements (Kaur & Bhat, 2019), thus, developing body awareness (Radhakrishna et al., 2010). Also increased sitting develops tolerance and various body postures help to build balance (Radhakrishna et al., 2010). The kinaesthetic information obtained from the joints and muscles may promote awareness of how the body is located in space (Proske et al., 1988). By stimulation of the kinaesthetic or the tactile senses during the practise of asanas, the self-efficacy may improve as the yoga postures involve a combination of body positioning, breathing patterns and balance (Milton et al., 2019). Children who are initially not even aware of their breath, gradually learn about the process of breathing along with the technique of deep breathing (Radhakrishna et al., 2010). Fine motor skills are usually underdeveloped in autistic individuals. Yoga practices are helpful in improving both gross (Radhakrishna, 2010) and fine motor skills (Kaur & Bhat, 2019). Motor coordination skills of a child are positively correlated to the successful execution of imitation skills (Bhat et al., 2011; Mostofsky et al., 2006; Dziuk et al., 2007; McDuffie et al., 2007). In order to imitate, social monitoring skills are required, that is to visually monitor the actions of social companion (Dziuk et al., 2007; Dadgar et al., 2017; McDuffie et al., 2007). On practising yoga regularly, the ability to imitate their social companions was found to increase considerably in such children (Radhakrishna, 2010; Kaur & Bhat, 2019). Specifically, mantra chanting has been found to be helpful in developing oro-facial movement and enhance vocal imitation skills along with the strengthening of the relationship between student and the therapist (Radhakrishna et al., 2010). Also, the functional use of objects by children was found to be better post yoga practise of six months (Ramanathan et al., 2019).

Self-Image The practise of yoga may also help in developing self-identity, with the help of which the children may be able to relate better to themselves and interact more comfortably with others. Study on a small sample of ASD children followed up for 9 months has found that yoga helped boost their self-confidence, socialcommunicative behaviour and strengthened their ability to make decisions (Radhakrishna et al., 2010). The practice of yoga has also been demonstrated to improve self-monitoring behaviour (Connor, 2005) and reduce self-stimulatory and self-injurious behaviours (Radhakrishna et al., 2010).

Sleep Issues With yoga practise an autistic child is able to sleep longer without frequent waking-ups, easily falls asleep at night and gets out of bed timely in the morning with no day time sleeping and reduces the habit of

370

 Yoga

snoring and breathing from the mouth during sleep. This also helps the parents and other family members sleep better as the child does not disturb anyone at night (Narasingharao et al., 2017).

Gastrointestinal Symptoms Yoga therapy has proved to be useful in promoting better dietary patterns. Children who performed yoga avoided over-eating and consumed a more balanced diet (Narasingharao et al., 2017). Number of studies have shown beneficial effects of yoga in improving digestion (Bhandari et al., 2012) and constipation (Rao et al., 2016). To summarize, regular yoga therapy intervention for 6 to 9 months may help in developing sustained attention, concentration, focus, along with improvement in self-esteem, social communication skills, social competence and behaviour (Galantino et al., 2008; Harrison & Hare, 2004; Jensen & Kenny, 2004). Yoga may also enhance motor sequences, body awareness, balance, coordination and imitation skills (Kenny, 2000) with improvement in overall quality of life in individuals with autism. All the major domains where yoga therapy has shown to bring improvement thus contributing in holistic development of such children has been provided in Figure 2. Along with this an evidence based table summarizing the problem domains in children with ASD and various yoga practices addressing the same along with its mechanism is provided in Table 1.

Yoga Practice Module for Autism Based on the literature review and our personal clinical experience with such children, a forty-five minute yoga module for autism has been provided below. The module starts with loosening practices which emphasize body movement and breath-synchronisation, this would improve attention, balance and coordination. This is followed by breathing practices in different animal postures. Such breathing practices would help in improving imitation skills, body awareness and mind-body connection. After these preparatory practices, yogasanas (yoga postures) are advised in different positions – standing, sitting, supine and prone. Most of these practices resemble one or the other creature – animal, bird or a particular pose such as warrior pose, bridge pose etc. After yogasanas, which are maintained for 10-15 seconds, in the beginning, a brief relaxation for 2 minutes is advised in Shavasana pose. During relaxation, children are asked to chant the sound “OM” gently for 9-10 rounds for deeper relaxation. After relaxation, Pranayama practices are advised which follow the sequence of faster breathing techniques followed by the slower ones i.e. Bhastrika (bellows breath 20 counts), followed by Nadi shuddhi Pranayama (alternate nostril breathing 9 rounds) and then Bhramari in Shanmukhi Mudra (Humming breath). These practices help improve cognitive functions, achieve better autonomic balance, reduce restlessness and bring calmness of the mind. In the end, mantra chanting is emphasized for 10 minutes to induce further relaxation and enhance cognitive functions: chanting which involves sounds A, U, M (with focus on the vibrations of the sounds) and some other mantras (such as Gayatri Mantra, Om Mani Padme Hum etc.) based on the socio-cultural background of the child can be used. The starting and ending of the session should involve a starting & closing prayer as per the cultural background of the subjects involved. Apart from this regular module, yoga postures should also be performed in the form of “Mandala” i.e. group of children forming a single geometrical shape together e.g., assuming flower geometries, simple pyramids or poses (such as Bhujangasana alternating with Parvatasana) in circular geometry. Kriya of “Jyoti Trataka” which involves gazing at a candle flame in a dark room 371

 Yoga

without blinking of the eyes for 10-15 minutes should also be practiced at least once a week to enhance cognitive and attentional faculties. This 45 minutes module should be practiced, preferably with parents and/or teachers, 5 day per week for a period of 6 to 9 months to obtain observable changes in imitation skills and social connect. The proposed yoga module for Autism is provided in Table 2. Figure 2. Multi-dimensional therapeutic approach of yoga therapy towards ASD

CONTRAINDICATIONS FOR YOGA PRACTICES IN AUTISM •

372

All complex yoga postures and pranayama requiring multiple set of instructions and that are difficult to learn should be avoided, e.g., practices such as Ujjayi and Kapalbhati. Practices with mudras, bandhas and meditative practices that involve withdrawal of the senses and intense focus at a particular point.

 Yoga

• •

• •

Risky postures which may lead to fall/injury such as Purna-chakrasana, Purna-ustrasana, Shirsasana and Sarvangasana. None of the postures should be maintained for more than 30 seconds (except relaxing asanas) and care should be taken to avoid very slow practices. To keep the interest of the children, practices should shift dynamically from one to another but instructions should be kept simple with clear demonstrations, starting with practices which are very easy to learn, gradually adding one step at a time slowly over weeks to months. The sequence of the practices should not be changed too frequently. The same sequence should be followed with a very slow gradual step by step up-gradation. All the Kriyas except Trataka should be avoided as they may prove to be risky due to difficulty following the procedure.

PRACTICAL TIPS OF DEALING CHILDREN WITH AUTISM WHILE TEACHING YOGA • •



• •

• •



Slow & gradual approach after understanding an individual’s needs - Each child is different with different needs, so the approach should be such that it caters to individual’s needs by being patient and understanding. Rapport building and connection development – It is important to form bond and trust with the child before starting the actual yoga session and this may take one, two or initial few sessions before they get ready to participate. One can take the help of chanting, music, story-telling, or massaging to accomplish this. Regularity in time, place and room/environmental settings – As children with ASD show restricted repetitive behaviour. So keeping regularity in time and place helps by not disturbing them and keeping the spatial set-up conditions same such as low lights, plain walls with nothing around that and placing yoga mats in the same place keeps them from being distracted. Providing assistance in group sessions - it is important to provide assistance in the beginning either by the teacher, parent or care-taker. Ideally, there should be one teacher or assistant per child in case of a group session. Making use of visual, auditory and tactile stimulations to enhance better learning – Generally, children with ASD tend to learn visually, so demonstration of the postures by the yoga teacher keeping their body at the level of their eyes and moving with students helps in better learning. Also, other visual tools like yoga cards and use of background music and gentle touch by the teacher can also help in enhanced learning. Increasing sitting tolerance by practicing seated & grounding postures - Children with high energy may need to only practice seated and grounding postures to begin with to soothe the nervous system. Keeping the same order in the sequence of practices – As these children respond to routines, yoga should be offered as an orderly and consistent activity with repeating postures in the same order every week. This will boost their confidence as they gradually learn the sequence and know what is coming next in subsequent session. If parents also participate with children, children learn faster. Parents can also be trained to conduct sessions for children at home. 373

 Yoga

Table 1. Evidence based summary of the problem domains in children with ASD and various yoga practices addressing the same along with its mechanism S.No.

1.

2.

3.

4.

5.

Domains

Social interaction

Verbal communication skills

Non-verbal communication skills

Sensory integration

Fine & gross motor skills

Challenges/Problems

Mechanism

Yoga practices found to be useful

References

Stimulates social growth by promotion of emotional expression and feeling of empathy towards others

Mandala yoga comprising of group poses, rhythmic chanting, and yoga games

Litchke et al., 2018

Enhances communication skills and language

Physical movement stimulates vestibular and proprioceptive system & strengthens recalling ability & language learning

Yogasanas including inverted postures, handstands, Ardhasherasanana (half inverted position), Ardhachakrasana (half wheel position), Trikonasana (triangular position), Uttanasana

Jensen, 2000; Kenny, 2002; Miramontez & Schwartz, 2017; Lee, 2017; Radhakrishna et al., 2010, Brownstone, 2001

Improves eye contact, visual cues, fine tunes sensitivity, recognition of facial expression, body language, appropriate facial expressions and expression of emotions

Non-competitive approach of yoga, calming effect on nervous system bringing parasympathetic dominance, practicing within their comfortable limits, increased awareness about body and self, thoughts and feelings, and reinforcement of ego

Integrated Movement Therapy, Classic relaxation pose, cleansing, calming breath, gradual stretching, Pranayama, Radhakrishna, 2010; kneeling poses, inverted Kaur, 2019; Kenny, 2002 poses such as Ardha Sarvangasana, standing poses like Tadasana and Vrksasana, and yoga games that incorporate the poses

Solutions/Addresses

Improves overall socializing ability, enhances teamwork skills, and improves social skills related to Social withdrawal, odd, spatial relationships, asocial and inappropriate improves emotional and poor interaction, expression, response limited social to initiation, initiating engagement, and poor interaction, and affective understanding and bonding in relationships perspective taking, improves ability to trust, share, initiate, and reciprocate, improves social cohesiveness

Delayed language milestones

Low social monitoring skills like poor eye contact, lack of facial expression, and impaired sensitivity

Deficits in sensory integration leads to poor concentration and reduced attention

Imitation skills remain impaired

Stimulation of vestibular and proprioceptive Improves overall system and increased sensory integration with body awareness, increased concentration, coordination & balance attention, focusing along with calming ability, alertness, effect brings overall sensitivity, organization, improvement in brain and on-task performance function helps in improved sensory integration

Physical postures – Ardhasirsasana, Trikonasana, Ardhachandrasana Pranayama – Alternate nostril yoga breathing, and Ujjayi

Oldenberg, 2007; Goldeberg, 2004; Kenny, 2002; Peck et al., 2005; Ramanathan et al., 2019

Significantly improved ability to follow motor patterns and to create and follow motor sequences, and improved fine and gross motor skills

Jogging, bending exercises, twisting, Trikonasana, Parshavakonasana, Veerabhadrasana, neck exercises, back bending exercises, relaxation exercises, Sukhasana, Shavasana, blowing exercises, in and out breathing, sectional, breathing

Radhakrishna, 2010; Kaur, 2019; Kenny, 2002

Stimulation of mirror neuron activity by guided imitation of therapist body positions

continued on following page

374

 Yoga Table 1. Continued S.No.

6.

7.

8.

9.

10.

11.

12.

Domains

Behaviour

Mood symptoms

Challenges/Problems

Restricted (maladaptive, disruptive, irritable, aggressive), repetitive, stereotyped behaviour, hyperactivity and conduct behaviours

Anxiety, sadness, depression, and lethargy

Mechanism

Yoga practices found to be useful

References

Increases positive behavioural responses, and decrease in disruptive behaviour

Decrease hyperactivity by calming down the nervous system by increased activity of the parasympathetic nervous system through integration of the right and left sides of the brain. Breath control helps to manage unnecessary and unwanted movements of the body and the mind

Inverted postures, alternate-nostril yogabreathing, alternate movement patterns, or patterns that cross the midline, such as Parivrittatrikonâsana and variations, Pranayama, Trikonasana, Parshavakonasana, Veerabhadrasana, neck exercises, back bending exercises, relaxation exercises, Sukhasana, Shavasana, blowing exercises, and sectional breathing

Koenig et al., 2012; Lee 2017; Milton, 2019; Kenny, 2002; Radhakrishna et al., 2010; Lee, 2017

Enhances and upgrades positive mood, induces calmness, improvement in happiness, and depressive symptoms

Inverted poses, Cat pose, Cow pose, Child’s pose, Stork pose, Dancer Stimulation of vestibular pose, Cobra pose, and proprioceptive Downward-dog pose, system and toning of the Frog pose, Airplane chest & spinal muscles pose, Back stretch pose, curbs hyperactivity Mountain pose Tree along with calming pose, Crane pose, Front effect & emotional stretch, Sleeping baby regulation pose, Pranayama, deep relaxation in Yoga nidra and chanting - kirtan Inverted positions, alternate-nostril yogabreathing, alternate movement patterns, or patterns that cross the midline, such as Parivrittatrikonâsana and variations, Pranayama, and IAYT procedures

Vidyashree et al., 2019

Solutions/Addresses

Rosenblatt et al., 2011; Kenny, 2002; Koenig et al., 2012; Brownstone, 2001; Litchke et al., 2018

Autonomic function

Autonomic dysfunction due to increased sympathetic and decreased parasympathetic tone

Restores autonomic balance by increasing parasympathetic and decreasing sympathetic activity

Calms down the nervous system by increased activity of the parasympathetic nervous system through integration of the right and left sides of the brain

Self-image

Poor self-confidence, self-esteem, and impaired ego

Improves selfconfidence, strengthens ego, improves self-esteem, self – acceptance, and selfmonitoring

Increases strength, coordination, courage, while encouraging appropriate risk-taking behaviours, and reinforcement of ego

Improved selfmonitoring skills, better communicative abilities, sense of achievement. All these boosts identity and improves self-image

Radhakrishna et al., 2010; Connor, 2005, Galantino et al., 2008; Harrison & Hare, 2004; Jensen & Kenny, 2004

Developing body awareness and improved balance & coordination between different body parts

Jogging, bending exercises, twisting, Trikonasana, Parshavakonasana, Veerabhadrasana, neck exercises, back bending exercises, relaxation exercises, Sukhasana, Shavasana, blowing exercises, in and out breathing, and sectional breathing

Kaur, 2019; Radhakrishna et al., 2010

Enhances restfulness, quality of sleep and makes falling asleep easier thus reducing day time sleeping

Starting prayer, breathing exercises, preparatory/dynamic practice, wind releasing practices, Sun-salutation (10 step and 12 step), relaxation, asana in standing, sitting, prone & supine posture, Pranayama, relaxation, chanting shloka and ending prayer.

Narasingharao et al., 2017

Starting prayer, breathing exercises, preparatory/dynamic practice, wind releasing Corrects wrong eating practices, Sun-salutation habits, overeating, (10 step and 12 step), absent mindedness relaxation, asana in during eating, promoting standing, sitting, prone balanced diet & supine posture, Pranayama, relaxation, chanting shloka and ending prayer.

Narasingharao et al., 2017

Sense of proprioception

Sleep

Gastrointestinal functions

Underdeveloped

Difficult initiating sleep, disturbed sleep, frequent waking-ups, impaired sleep quality, reduced duration, day time sleeping

Impaired

Improves sense of proprioception

Improves overall quality of sleep

Improves

375

 Yoga

Table 2. Proposed 45 minute yoga module for children with ASD S. no.

List of Practices

Practice English names

Practice Hindi names

Number of rounds

Duration

References

1.

Starting prayer

Opening prayer

Aarambha prarthana

1

1 minute

Gulati et al., 2018; Vidyashree et al., 2019; Narasingharao et al., 2017

2.

Joint loosening practices

Neck movements Shoulder rotation Wrist joint movements Waist rotation Knee stretching Ankle joint movements

Griva Shakti vikasaka Skandha Shakti vikasaka Manibandha Shakti vikasaka Kati Shakti vikasaka Janu Shakti vikasaka Gulfa Shakti vikasaka

5 rounds each

5 minutes

Gulati et al., 2018; Radhakrishna et al., 2010

3.

Balancing practices

Standing deltoid stretch Walking with deltoid stretch

Uttana Tadasana Uttana Tadasana Mein gati

5 rounds each

2 minutes

Gulati et al., 2018; Ramanathan et al., 2019

4.

Breathing practices

Tiger breathing Rabbit breathing Dog breathing Lion roar breathing

Vyaghrasana svasana kriya Sasankasana svasana kriya Svanasana svasana kriya Simha mudra

5 rounds each

8 minutes

Narasingharao et al., 2017; Rosenblatt et al., 2011; Radhakrishna et al., 2010

5.

Yoga postures Standing Sitting Supine Prone

Triangle pose Tree pose Warrior pose Rabbit pose Thunderbolt pose Half head stand pose Wind releasing pose Bridge pose Cobra pose Cat pose

Trikonasana Vrikshasana Virabhadrasana Sasankasana Vajrasana Ardha sirsasana Pawanmuktasana Setubandhasana Bhujangasana Marjarasana

3 rounds each 3 rounds each 3 rounds each 3 rounds each

3 minutes 3 minutes 3 minutes 3 minutes

Radhakrishna et al., 2010; Kaur, 2019; Gulati et al., 2018; Rosenblatt et al., 2011; Kenny, 2002; Narasingharao et al., 2017

6.

Relaxation with Om chanting

Corpse pose

Savasana

Om chanting 6 times

2 minutes

Radhakrishna et al., 2010; Vidyashree et al., 2019

7.

Pranayama

Bellow breathing Alternate nostril yoga-breathing Humming bee breathing

Bhastrika Nadi shuddhi Bhramari

20/cycle – 3 rounds 9 rounds 5 rounds

2 minutes 3 minutes 1 minute

Ramanathan et al., 2019; Vidyashree et al., 2019; Gulati et al., 2018; Narasingharao et al., 2017

8.

Relaxation by chanting

Chanting of syllables A,U,M, A-U-M chanting / Mantra chanting

‘Nadanusandhana / Gayathri mantra

9 rounds

8 minutes

Radhakrishna et al., 2010; Narasingharao et al., 2017; Ramanathan et al., 2019

9.

Ending prayer

Closing prayer

“Sarve bhavantu sukhinah..”

1

1 minute

Gulati et al., 2018; Vidyashree et al., 2019; Narasingharao et al., 2017

10.

Yoga postures in Groups

Flower geometries, simple pyramids, postures in circles

Mandala yoga

-

-

Litchke et al., 2018

LIMITATIONS IN EVIDENCE-BASED YOGA THERAPY FOR ASD Most of the yoga studies provide only short-term effects of the intervention. Those studies where longterm follow-up was done have methodological issues. Long-term follow-up studies with randomized controlled design and robust methodology should be performed in future to establish definite clinical value of yoga in ASD. Also, very limited evidence is available towards the neurobiological effects of yoga in ASD. This too needs further exploration in future studies. Different studies followed different yoga program causing heterogeneity in interventions across studies, hindering the clinical application of yoga therapy. So the development of a standardized evidence based yoga program for ASD should be the first step towards future exploration in this area.

376

 Yoga

CONCLUSION Yoga therapy, being holistic and patient empowering in its approach has the potential to improve cognitive and social deficits and improve the overall quality of life of individuals with ASD. Future researches should explore this further through good quality trials with long-term follow-up and robust research designs.

REFERENCES Adams, J. B., Holloway, C. E., George, F., & Quig, D. (2006). Analyses of toxic metals and essential minerals in the hair of Arizona children with autism and associated conditions, and their mothers. Biological Trace Element Research, 110(3), 193–209. doi:10.1385/BTER:110:3:193 PMID:16845157 Adams, J. B., Johansen, L. J., Powell, L. D., Quig, D., & Rubin, R. A. (2011). Gastrointestinal flora and gastrointestinal status in children with autism–comparisons to typical children and correlation with autism severity. BMC Gastroenterology, 11(1), 22. doi:10.1186/1471-230X-11-22 PMID:21410934 Akshoomoff, N., Pierce, K., & Courchesne, E. (2002). The neurobiological basis of autism from a developmental perspective. Development and Psychopathology, 14(3), 613–634. doi:10.1017/S0954579402003115 PMID:12349876 Alpert, M. (2007). The autism diet. Scientific American, 296(4), 19–20. doi:10.1038cientificameric an0407-19 PMID:17479620 American Psychiatric Association. (2013). Diagnostic and statistical manual of mental disorders (DSM5®). American Psychiatric Publishing. Arango, V., Huang, Y. Y., Underwood, M. D., & Mann, J. J. (2003). Genetics of the serotonergic system in suicidal behavior. Journal of Psychiatric Research, 37(5), 375–386. doi:10.1016/S0022-3956(03)000487 PMID:12849930 Artchoudane, S., Bhavanani, A. B., Ramanathan, M., & Mariangela, A. (2019). Yoga as a therapeutic tool in autism: A detailed review. Yoga Mimamsa, 51(1), 3. doi:10.4103/ym.ym_3_19 Asperger, H. (1944). The “autistic psychopaths” in childhood. Archives of Psychiatry and Nervous. Diseases (Basel, Switzerland), 117(1), 76–136. Ayres, A. J. (1995). Sensory Integration and the Child. Western Psychological Services. Baranek, G. T. (1999). Autism during infancy: A retrospective video analysis of sensory-motor and social behaviors at 9–12 months of age. Journal of Autism and Developmental Disorders, 29(3), 213–224. doi:10.1023/A:1023080005650 PMID:10425584 Behar, M. (2006). Yoga therapy for autistic children. Yoga Therapy in Practice, 2(2), 18–19. Berg, A. T., & Dobyns, W. B. (2015). Progress in autism and related disorders of brain development. Lancet Neurology, 14(11), 1069–1070. doi:10.1016/S1474-4422(15)00048-4 PMID:25891008

377

 Yoga

Bhandari, R. B., Bālakr̥shṇa (Ācārya), Bhandari, C. B., Sharma, G. D., Singh, K., Pandya, P., & Katiyar, V. K. (2012). Implications of Corporate Yoga: A review. INTECH Open Access Publisher. Bhat, A. N., Landa, R. J., & Galloway, J. C. (2011). Current perspectives on motor functioning in infants, children, and adults with autism spectrum disorders. Physical Therapy, 91(7), 1116–1129. doi:10.2522/ ptj.20100294 PMID:21546566 Bhavanani, A. B. (2014). Yoga: A novel integrative therapy. Nisargopachar Varta, 6(5), 13–15. Black, C., Kaye, J. A., & Jick, H. (2002). Relation of childhood gastrointestinal disorders to autism: Nested case-control study using data from the UK General Practice Research Database. BMJ (Clinical Research Ed.), 325(7361), 419–421. doi:10.1136/bmj.325.7361.419 PMID:12193358 Bolton, P. F., Golding, J., Emond, A., & Steer, C. D. (2012). Autism spectrum disorder and autistic traits in the Avon Longitudinal Study of Parents and Children: Precursors and early signs. Journal of the American Academy of Child and Adolescent Psychiatry, 51(3), 249–260. doi:10.1016/j.jaac.2011.12.009 PMID:22365461 Brewster, K., Russell, J. L., & Webb, W. G. (1999). Neurology for the Speech-Language Pathologist (3rd ed.). Butterworth-Heinemann Medical. Broadstock, M., Doughty, C., & Eggleston, M. (2007). Systematic review of the effectiveness of pharmacological treatments for adolescents and adults with autism spectrum disorder. Autism, 11(4), 335–348. doi:10.1177/1362361307078132 PMID:17656398 Brownstone, A. (2001). Therapeutic mechanisms of yoga âsana. International Journal of Yoga Therapy, 11(1), 11–14. doi:10.17761/ijyt.11.1.y1711u773575m48t Campbell, M., Anderson, L. T., Small, A. M., Adams, P., Gonzalez, N. M., & Ernst, M. (1993). Naltrexone in autistic children: Behavioral symptoms and attentional learning. Journal of the American Academy of Child and Adolescent Psychiatry, 32(6), 1283–1291. doi:10.1097/00004583-19931100000024 PMID:8282676 Carper, R. A., & Courchesne, E. (2005). Localized enlargement of the frontal cortex in early autism. Biological Psychiatry, 57(2), 126–133. doi:10.1016/j.biopsych.2004.11.005 PMID:15652870 Carter, A. S., Davis, N. O., Klin, A., & Volkmar, F. R. (2005). Social Development in Autism. In F. R. Volkmar, R. Paul, A. Klin, & D. Cohen (Eds.), Diagnosis, Development, Neurobiology, and Behavior (pp. 312–334). John Wiley & Sons Inc. Centres for Disease Control and Prevention. (2016). Autism data visualization tool. Retrieved from https://www.cdc.gov/ncbddd/autism/data/index.html#data Chaidez, V., Hansen, R. L., & Hertz-Picciotto, I. (2014). Gastrointestinal problems in children with autism, developmental delays or typical development. Journal of Autism and Developmental Disorders, 44(5), 1117–1127. doi:10.100710803-013-1973-x PMID:24193577 Charman, T., Baron-Cohen, S., Swettenham, J., Baird, G., Drew, A., & Cox, A. (2003). Predicting language outcome in infants with autism and pervasive developmental disorder. International Journal of Language & Communication Disorders, 38(3), 265–285. doi:10.1080/136820310000104830 PMID:12851079

378

 Yoga

Charman, T., Taylor, E., Drew, A., Cockerill, H., Brown, J. A., & Baird, G. (2005). Outcome at 7 years of children diagnosed with autism at age 2: Predictive validity of assessments conducted at 2 and 3 years of age and pattern of symptom change over time. Journal of Child Psychology and Psychiatry, and Allied Disciplines, 46(5), 500–513. doi:10.1111/j.1469-7610.2004.00377.x PMID:15845130 Choudhury, P. R., Lahiri, S., & Rajamma, U. (2012). Glutamate mediated signaling in the pathophysiology of autism spectrum disorders. Pharmacology, Biochemistry, and Behavior, 100(4), 841–849. doi:10.1016/j.pbb.2011.06.023 PMID:21756930 Christakou, A., Murphy, C. M., Chantiluke, K., Cubillo, A. I., Smith, A. B., Giampietro, V., Daly, E., Ecker, C., Robertson, D., Murphy, D. G., & Rubia, K. (2013). Disorder-specific functional abnormalities during sustained attention in youth with attention deficit hyperactivity disorder (ADHD) and with autism. Molecular Psychiatry, 18(2), 236–244. doi:10.1038/mp.2011.185 PMID:22290121 Christensen, J., Grønborg, T. K., Sørensen, M. J., Schendel, D., Parner, E. T., Pedersen, L. H., & Vestergaard, M. (2013). Prenatal valproate exposure and risk of autism spectrum disorders and childhood autism. Journal of the American Medical Association, 309(16), 1696–1703. doi:10.1001/jama.2013.2270 PMID:23613074 Cohen, D., Pichard, N., Tordjman, S., Baumann, C., Burglen, L., Excoffier, E., ... Héron, D. (2005). Specific genetic disorders and autism: Clinical contribution towards their identification. Journal of Autism and Developmental Disorders, 35(1), 103–116. doi:10.100710803-004-1038-2 PMID:15796126 Connor, C. (2005). So much stress; so little time. Autism News Orange County, 2, 18–20. Courchesne, E. (2002). Abnormal early brain development in autism. Molecular Psychiatry, 7(2), S21– S23. doi:10.1038j.mp.4001169 PMID:12142938 Courchesne, E., Campbell, K., & Solso, S. (2011). Brain growth across the life span in autism: Age-specific changes in anatomical pathology. Brain Research, 1380, 138–145. doi:10.1016/j.brainres.2010.09.101 PMID:20920490 Courchesne, E., Karns, C. M., Davis, H. R., Ziccardi, R., Carper, R. A., Tigue, Z. D., ... Lincoln, A. J. (2001). Unusual brain growth patterns in early life in patients with autistic disorder: An MRI study. Neurology, 57(2), 245–254. doi:10.1212/WNL.57.2.245 PMID:11468308 Dadgar, H., Rad, J. A., Soleymani, Z., Khorammi, A., McCleery, J., & Maroufizadeh, S. (2017). The relationship between motor, imitation, and early social communication skills in children with autism. Iranian Journal of Psychiatry, 12(4), 236. PMID:29472949 Day, B. L., & Fitzpatrick, R. C. (2005). The vestibular system. Current Biology, 15(15), R583–R586. doi:10.1016/j.cub.2005.07.053 PMID:16085475 de Bruin, E. I., Blom, R., Smit, F. M., van Steensel, F. J., & Bögels, S. M. (2015). MYmind: Mindfulness training for youngsters with autism spectrum disorders and their parents. Autism, 19(8), 906–914. doi:10.1177/1362361314553279 PMID:25348866 De Giacomo, A., & Fombonne, E. (1998). Parental recognition of developmental abnormalities in autism. European Child & Adolescent Psychiatry, 7(3), 131–136. doi:10.1007007870050058 PMID:9826299

379

 Yoga

DeFilippis, M., & Wagner, K. D. (2016). Treatment of autism spectrum disorder in children and adolescents. Psychopharmacology Bulletin, 46(2), 18. PMID:27738378 Deorari, M., & Bhardwaj, I. (2014). Effect of yogic intervention on autism spectrum disorder. Yoga Mimamsa, 46(3), 81. doi:10.4103/0044-0507.159744 Durkin, M. S., Maenner, M. J., Newschaffer, C. J., Lee, L. C., Cunniff, C. M., Daniels, J. L., Kirby, R. S., Leavitt, L., Miller, L., Zahorodny, W., & Schieve, L. A. (2008). Advanced parental age and the risk of autism spectrum disorder. American Journal of Epidemiology, 168(11), 1268–1276. doi:10.1093/aje/ kwn250 PMID:18945690 Dziuk, M. A., Larson, J. G., Apostu, A., Mahone, E. M., Denckla, M. B., & Mostofsky, S. H. (2007). Dyspraxia in autism: Association with motor, social, and communicative deficits. Developmental Medicine and Child Neurology, 49(10), 734–739. doi:10.1111/j.1469-8749.2007.00734.x PMID:17880641 Ecker, C., Suckling, J., Deoni, S. C., Lombardo, M. V., Bullmore, E. T., Baron-Cohen, S., ... Williams, S. C. (2012). Brain anatomy and its relationship to behaviour in adults with autism spectrum disorder: A multicentre magnetic resonance imaging study. Archives of General Psychiatry, 69(2), 195–209. doi:10.1001/archgenpsychiatry.2011.1251 PMID:22310506 Ehleringer, J. (2010). Yoga for children on the autism spectrum. International Journal of Yoga Therapy, 20(1), 131–139. doi:10.17761/ijyt.20.1.eu176u2721423510 Eigsti, I. M., de Marchena, A. B., Schuh, J. M., & Kelley, E. (2011). Language acquisition in autism spectrum disorders: A developmental review. Research in Autism Spectrum Disorders, 5(2), 681–691. doi:10.1016/j.rasd.2010.09.001 Elder, J. H., Kreider, C. M., Schaefer, N. M., & de Laosa, M. B. (2015). A review of gluten-and caseinfree diets for treatment of autism: 2005–2015. Nutrition and Dietary Supplements, 7, 87. doi:10.2147/ NDS.S74718 PMID:28111520 Elder, J. H., Shankar, M., Shuster, J., Theriaque, D., Burns, S., & Sherrill, L. (2006). The gluten-free, casein-free diet in autism: Results of a preliminary double blind clinical trial. Journal of Autism and Developmental Disorders, 36(3), 413–420. doi:10.100710803-006-0079-0 PMID:16555138 Essa, M. M., & Qoronfleh, M. W. (Eds.). (2020). Personalized food intervention and therapy for autism spectrum disorder management (Vol. 24). Springer Nature. doi:10.1007/978-3-030-30402-7 Ferraro, F. R. (2016). No evidence of reaction time slowing in autism spectrum disorder. Autism, 20(1), 116–122. doi:10.1177/1362361314559986 PMID:25512973 Field, T. (2016). Yoga research review. Complementary Therapies in Clinical Practice, 24, 145–161. doi:10.1016/j.ctcp.2016.06.005 PMID:27502816 Fombonne, E. (2006). Past and Future Perspectives on Autism Epidemiology. In S. O. Moldin & J. L. R. Rubenstein (Eds.), Understanding autism: From basic neuroscience to treatment (pp. 25–48). Taylor & Francis Group. doi:10.1201/9781420004205.ch2 Gail Williams, P., Sears, L. L., & Allard, A. (2004). Sleep problems in children with autism. Journal of Sleep Research, 13(3), 265–268. doi:10.1111/j.1365-2869.2004.00405.x PMID:15339262

380

 Yoga

Gajcy, K., Lochynski, S., & Librowski, T. (2010). A role of GABA analogues in the treatment of neurological diseases. Current Medicinal Chemistry, 17(22), 2338–2347. doi:10.2174/092986710791698549 PMID:20491640 Galantino, M. L., Galbavy, R., & Quinn, L. (2008). Therapeutic effects of yoga for children: A systematic review of the literature. Pediatric Physical Therapy, 20(1), 66–80. doi:10.1097/PEP.0b013e31815f1208 PMID:18300936 Garcia, J. M., Baker, K., Diaz, M. R., Tucker, J. E., Kelchner, V. P., & Rice, D. J. (2019). Implementation fidelity of a mindfulness-based yoga program for children with autism spectrum disorder and their families: A pilot study. Advances in Neurodevelopmental Disorders, 3(1), 54–62. doi:10.100741252-018-0091-3 Gardener, H., Spiegelman, D., & Buka, S. L. (2011). Perinatal and neonatal risk factors for autism: A comprehensive meta-analysis. Pediatrics, 128(2), 344–355. doi:10.1542/peds.2010-1036 PMID:21746727 Goldberg, L. (2004). Creative RelaxationSM: A Yoga-based program for regular and exceptional student education. International Journal of Yoga Therapy, 14(1), 68–78. doi:10.17761/ijyt.14.1.50524j511674v292 Greenberg, D. S., Toiber, D., Berson, A., & Soreq, H. (2010). Acetylcholinesterase variants in Alzheimer’s disease: From neuroprotection to programmed cell death. Neurodegenerative Diseases, 7(1-3), 60–63. doi:10.1159/000285507 PMID:20173328 Grzadzinski, R., Huerta, M., & Lord, C. (2013). DSM-5 and autism spectrum disorders (ASDs): An opportunity for identifying ASD subtypes. Molecular Autism, 4(1), 1–6. doi:10.1186/2040-2392-4-12 PMID:23675638 Gulati, K., Loganathan, N., Mooventhan, A., Lahiri, A., & Telles, S. (2018). Effect of yoga therapy on the symptoms of sensory processing disorder in autistic individuals. Yoga Mimamsa, 50(2), 60. doi:10.4103/ym.ym_9_18 Hall, S. S., Lightbody, A. A., & Reiss, A. L. (2008). Compulsive, self-injurious, and autistic behavior in children and adolescents with fragile X syndrome. American Journal of Mental Retardation, 113(1), 44–53. doi:10.1352/0895-8017(2008)113[44:CSAABI]2.0.CO;2 PMID:18173299 Hallmayer, J., Cleveland, S., Torres, A., Phillips, J., Cohen, B., Torigoe, T., ... Lotspeich, L. (2011). Genetic heritability and shared environmental factors among twin pairs with autism. Archives of General Psychiatry, 68(11), 1095–1102. doi:10.1001/archgenpsychiatry.2011.76 PMID:21727249 Hanson, E., Kalish, L. A., Bunce, E., Curtis, C., McDaniel, S., Ware, J., & Petry, J. (2007). Use of complementary and alternative medicine among children diagnosed with autism spectrum disorder. Journal of Autism and Developmental Disorders, 37(4), 628–636. doi:10.100710803-006-0192-0 PMID:16977497 Harrington, J. W., Rosen, L., Garnecho, A. N. A., & Patrick, P. A. (2006). Parental perceptions and use of complementary and alternative medicine practices for children with autistic spectrum disorders in private practice. Journal of Developmental and Behavioral Pediatrics, 27(2, Supplement 2), S156–S161. doi:10.1097/00004703-200604002-00014 PMID:16685182

381

 Yoga

Harrison, J., & Hare, D. J. (2004). Brief report: Assessment of sensory abnormalities in people with autistic spectrum disorders. Journal of Autism and Developmental Disorders, 34(6), 727–730. doi:10.100710803004-5293-z PMID:15679192 Haydon, P. G., McCobb, D. P., & Kater, S. B. (1987). The regulation of neurite outgrowth, growth cone motility, and electrical synaptogenesis by serotonin. Journal of Neurobiology, 18(2), 197–215. doi:10.1002/neu.480180206 PMID:2883253 Hediger, M. L., England, L. J., Molloy, C. A., Kai, F. Y., Manning-Courtney, P., & Mills, J. L. (2008). Reduced bone cortical thickness in boys with autism or autism spectrum disorder. Journal of Autism and Developmental Disorders, 38(5), 848–856. doi:10.100710803-007-0453-6 PMID:17879151 Hekmat Pou, D., Khonsari, A. H., Shamsikhani, S., Matouri Pour, P., & Gholami, M. (2016). The Effect of super brain yoga on children with autism disorder. Complementary Medicine Journal, 6(3), 1549–1559. Hellings, J. A., Weckbaugh, M., Nickel, E. J., Cain, S. E., Zarcone, J. R., Reese, R. M., Hall, S., Ermer, D. J., Tsai, L. Y., Schroeder, S. R., & Cook, E. H. (2005). A Double-Blind, Placebo-Controlled Study of Valproate for Aggression in Youth with Pervasive Developmental Disorders. Journal of Child and Adolescent Psychopharmacology, 15(4), 682–692. doi:10.1089/cap.2005.15.682 PMID:16190799 Hellings, J. A., Zarcone, J. R., Crandall, K., Wallace, D., & Schroeder, S. R. (2001). Weight gain in a controlled study of risperidone in children, adolescents and adults with mental retardation and autism. Journal of Child and Adolescent Psychopharmacology, 11(3), 229–238. doi:10.1089/10445460152595559 PMID:11642473 Horvath, K., & Perman, J. A. (2002). Autism and gastrointestinal symptoms. Current Gastroenterology Reports, 4(3), 251–258. doi:10.100711894-002-0071-6 PMID:12010627 Hossain, M. D., Ahmed, H. U., Uddin, M. J., Chowdhury, W. A., Iqbal, M. S., Kabir, R. I., ... Hossain, S. W. (2017). Autism Spectrum disorders (ASD) in South Asia: A systematic review. BMC Psychiatry, 17(1), 1–7. doi:10.118612888-017-1440-x PMID:28826398 Hourston, S., & Atchley, R. (2017). Autism and mind–body therapies: A systematic review. Journal of Alternative and Complementary Medicine (New York, N.Y.), 23(5), 331–339. doi:10.1089/acm.2016.0336 PMID:28437148 Hyman, S. L., Stewart, P. A., Foley, J., Peck, R., Morris, D. D., Wang, H., & Smith, T. (2016). The gluten-free/casein-free diet: A double-blind challenge trial in children with autism. Journal of Autism and Developmental Disorders, 46(1), 205–220. doi:10.100710803-015-2564-9 PMID:26343026 Jain, S. C., & Talukdar, B. (1993). Evaluation of yoga therapy programme for patients of bronchial asthma. Singapore Medical Journal, 34, 306–306. PMID:8266199 Jensen, E. (2000). Brain-Based Learning. Brain Store Inc. Jensen, P. S., & Kenny, D. T. (2004). The effects of yoga on the attention and behavior of boys with attention-deficit/hyperactivity disorder (ADHD). Journal of Attention Disorders, 7(4), 205–216. doi:10.1177/108705470400700403 PMID:15487477

382

 Yoga

Jeste, S. S., & Geschwind, D. H. (2014). Disentangling the heterogeneity of autism spectrum disorder through genetic findings. Nature Reviews. Neurology, 10(2), 74–81. doi:10.1038/nrneurol.2013.278 PMID:24468882 Johnson, C. R., Handen, B. L., Zimmer, M., Sacco, K., & Turner, K. (2011). Effects of gluten free/ casein free diet in young children with autism: A pilot study. Journal of Developmental and Physical Disabilities, 23(3), 213–225. doi:10.100710882-010-9217-x Kanner, L. (1943). Autistic disturbances of affective contact. Nervous Child, 2(3), 217–250. Kaur, M., & Bhat, A. (2019). Creative yoga intervention improves motor and imitation skills of children with autism spectrum disorder. Physical Therapy, 99(11), 1520–1534. doi:10.1093/ptj/pzz115 PMID:31765484 Kenny, M. (2002). Integrated movement therapy™: Yoga-based therapy as a viable and effective intervention for autism spectrum and related disorders. International Journal of Yoga Therapy, 12(1), 71–79. doi:10.17761/ijyt.12.1.r978vxt214683904 Kerbeshian, J., & Burd, L. (1996). Case study: Comorbidity among Tourette’s syndrome, autistic disorder, and bipolar disorder. Journal of the American Academy of Child and Adolescent Psychiatry, 35(5), 681–685. doi:10.1097/00004583-199605000-00024 PMID:8935216 Kerbeshian, J., Burd, L., & Avery, K. (2001). Pharmacotherapy of autism: A review and clinical approach. Journal of Developmental and Physical Disabilities, 13(3), 199–228. doi:10.1023/A:1016686802786 Kern, J. K., Miller, V. S., Evans, P. A., & Trivedi, M. H. (2002). Efficacy of porcine secretin in children with autism and pervasive developmental disorder. Journal of Autism and Developmental Disorders, 32(3), 153–160. doi:10.1023/A:1015441428154 PMID:12108616 Knivsberg, A. M., Reichelt, K. L., Høien, T., & Nødland, M. (2002). A randomised, controlled study of dietary intervention in autistic syndromes. Nutritional Neuroscience, 5(4), 251–261. doi:10.1080/10284150290028945 PMID:12168688 Koenig, K. P., Buckley-Reen, A., & Garg, S. (2012). Efficacy of the Get Ready to Learn yoga program among children with autism spectrum disorders: A pretest–posttest control group design. The American Journal of Occupational Therapy, 66(5), 538–546. doi:10.5014/ajot.2012.004390 PMID:22917120 Korkmaz, B. (2000). Infantile autism: Adult outcome. Seminars in Clinical Neuropsychiatry, 5(3), 164–170. doi:10.1053cnp.2000.6727 PMID:11291011 Korvatska, E., Van de Water, J., Anders, T. F., & Gershwin, M. E. (2002). Genetic and immunologic considerations in autism. Neurobiology of Disease, 9(2), 107–125. doi:10.1006/nbdi.2002.0479 PMID:11895365 Koterba, R. (2007). Superbrain yoga in children with autism and ADHD. Winter: Prana World, 14-17. Kranowitz, C. S., & Archer, E. (1998). The out-of-sync child. Berkley Publishing Group. Kwong, W. H., Chan, W. Y., Lee, K. K. H., Fan, M., & Yew, D. T. (2000). Neurotransmitters, neuropeptides and calcium binding proteins in developing human cerebellum: A review. The Histochemical Journal, 32(9), 521–534. doi:10.1023/A:1004197210189 PMID:11127973

383

 Yoga

Lange, N., Travers, B. G., Bigler, E. D., Prigge, M. B., Froehlich, A. L., Nielsen, J. A., ... Alexander, A. A. (2015). Longitudinal volumetric brain changes in autism spectrum disorder ages 6–35 years. Autism Research, 8(1), 82–93. doi:10.1002/aur.1427 PMID:25381736 Larkin, M. (1997). Approaches to amelioration of autism in adulthood. Lancet, 349(9046), 186. doi:10.1016/S0140-6736(05)60993-2 Lee, L. N. (2017). The Effects of Yoga on a Male Child with Autism Spectrum Disorder (Masters dissertation). Retrieved from https://thescholarship.ecu.edu/bitstream/handle/10342/6475/LEE-MASTERSTHESIS-2017.pdf?sequence=1&isAllowed=y Leekam, S. R., Prior, M. R., & Uljarevic, M. (2011). Restricted and repetitive behaviors in autism spectrum disorders: A review of research in the last decade. Psychological Bulletin, 137(4), 562–593. doi:10.1037/a0023341 PMID:21574682 Levy, S. E., & Hyman, S. L. (2005). Novel treatments for autistic spectrum disorders. Mental Retardation and Developmental Disabilities Research Reviews, 11(2), 131–142. doi:10.1002/mrdd.20062 PMID:15977319 Levy, S. E., Souders, M. C., Ittenbach, R. F., Giarelli, E., Mulberg, A. E., & Pinto-Martin, J. A. (2007). Relationship of dietary intake to gastrointestinal symptoms in children with autistic spectrum disorders. Biological Psychiatry, 61(4), 492–497. doi:10.1016/j.biopsych.2006.07.013 PMID:17207470 Litchke, L. G., Liu, T., & Castro, S. (2018). Effects of multimodal mandala yoga on social and emotional skills for youth with autism spectrum disorder: An exploratory study. International Journal of Yoga, 11(1), 59. PMID:29343932 Liu, J., Wan, G. B., Huang, M. S., Agyapong, G., Zhang, X. Y., Liu, Y. W., ... Kong, X. J. (2019). Probiotic therapy for treating behavioral and gastrointestinal symptoms in autism spectrum disorder: A systematic review of clinical trials. Current Medical Science, 39(2), 173–184. doi:10.100711596-0192016-4 PMID:31016507 Liu, J., Zhang, M., & Kong, X. (2016). Gut Microbiome and Autism: Recent Advances and Future Perspectives. North American Journal of Medical Sciences, 9(3), 104–115. Livingston, L. A., & Happé, F. (2017). Conceptualising compensation in neurodevelopmental disorders: Reflections from autism spectrum disorder. Neuroscience and Biobehavioral Reviews, 80, 729–742. doi:10.1016/j.neubiorev.2017.06.005 PMID:28642070 Lud Cadet, J., Jayanthi, S., McCoy, M., Beauvais, G., & Sheng Cai, N. (2010). Dopamine D1 receptors, regulation of gene expression in the brain, and neurodegeneration. CNS & Neurological Disorders-Drug Targets, 9(5), 526-538. Ly, V., Bottelier, M., Hoekstra, P. J., Vasquez, A. A., Buitelaar, J. K., & Rommelse, N. N. (2017). Elimination diets’ efficacy and mechanisms in attention deficit hyperactivity disorder and autism spectrum disorder. European Child & Adolescent Psychiatry, 26(9), 1067–1079. doi:10.100700787-017-0959-1 PMID:28190137

384

 Yoga

Manwaring, S. S., Mead, D. L., Swineford, L., & Thurm, A. (2017). Modelling gesture use and early language development in autism spectrum disorder. International Journal of Language & Communication Disorders, 52(5), 637–651. doi:10.1111/1460-6984.12308 PMID:28120370 Mari-Bauset, S., Zazpe, I., Mari-Sanchis, A., Llopis-González, A., & Morales-Suarez-Varela, M. (2014). Evidence of the gluten-free and casein-free diet in autism spectrum disorders: A systematic review. Journal of Child Neurology, 29(12), 1718–1727. doi:10.1177/0883073814531330 PMID:24789114 McDougle, C. J., Holmes, J. P., Carlson, D. C., Pelton, G. H., Cohen, D. J., & Price, L. H. (1998). A double-blind, placebo-controlled study of risperidone in adults with autistic disorder and other pervasive developmental disorders. Archives of General Psychiatry, 55(7), 633–641. doi:10.1001/archpsyc.55.7.633 PMID:9672054 McDuffie, A., Turner, L., Stone, W., Yoder, P., Wolery, M., & Ulman, T. (2007). Developmental correlates of different types of motor imitation in young children with autism spectrum disorders. Journal of Autism and Developmental Disorders, 37(3), 401–412. doi:10.100710803-006-0175-1 PMID:16900404 Milton, L. E., Bantel, S., Calmer, K., Friedman, M., Haley, E., & Rubarts, L. (2019). Yoga and autism: Students’ perspectives on the get ready to learn yoga program. The Open Journal of Occupational Therapy, 7(4), 1–10. doi:10.15453/2168-6408.1560 Ming, X., Bain, J. M., Smith, D., Brimacombe, M., Gold von-Simson, G., & Axelrod, F. B. (2011). Assessing autonomic dysfunction symptoms in children: A pilot study. Journal of Child Neurology, 26(4), 420–427. doi:10.1177/0883073810381921 PMID:21196528 Ming, X., Brimacombe, M., Chaaban, J., Zimmerman-Bier, B., & Wagner, G. C. (2008). Autism spectrum disorders: Concurrent clinical disorders. Journal of Child Neurology, 23(1), 6–13. doi:10.1177/0883073807307102 PMID:18056691 Miramontez, S. K., & Schwartz, I. S. (2017). The effects of physical activity on the on-task behavior of young children with autism spectrum disorders. International Electronic Journal of Elementary Education, 9(2), 405–418. Molloy, C. A., & Manning-Courtney, P. (2003). Prevalence of chronic gastrointestinal symptoms in children with autism and autistic spectrum disorders. Autism, 7(2), 165–171. doi:10.1177/1362361303007002004 PMID:12846385 Moradi, A., & Esmaeilzadeh, S. (2017). Simple reaction time and obesity in children: Whether there is a relationship? Environmental Health and Preventive Medicine, 22(1), 1–6. doi:10.118612199-0170612-0 PMID:29165104 Mostofsky, S. H., Dubey, P., Jerath, V. K., Jansiewicz, E. M., Goldberg, M. C., & Denckla, M. B. (2006). Developmental dyspraxia is not limited to imitation in children with autism spectrum disorders. Journal of the International Neuropsychological Society, 12(3), 314. doi:10.1017/S1355617706060437 PMID:16903124 Nagendra, H. R. (2013). Integrated yoga therapy for mental illness. Indian Journal of Psychiatry, 55(3), S337. PMID:24049195

385

 Yoga

Nagendra, H. R., & Nagarathna, R. (2008). Integrated approach of yoga therapy for positive health. Swami Vivekananda Yoga Prakashana. Narasingharao, K., & Pradhan, B. (2017). Efficacy of structured yoga intervention for sleep, gastrointestinal and behaviour problems of ASD children: An exploratory study. Journal of Clinical and Diagnostic Research: JCDR, 11(3), VC01. doi:10.7860/JCDR/2017/25894.9502 PMID:28511484 National Institutes of Health. (2011). Office of dietary supplements: dietary supplement fact sheet: folate. Retrieved from http://ods.od.nih.gov/factsheets/folate.asp#en10 Nevola, E. (2017). A clinical case study: Using yoga to improve functional communication in an adolescent with autism spectrum disorder. Academic Festival, 71. Nicoletti, F., Bockaert, J., Collingridge, G. L., Conn, P. J., Ferraguti, F., Schoepp, D. D., Wroblewski, J. T., & Pin, J. P. (2011). Metabotropic glutamate receptors: From the workbench to the bedside. Neuropharmacology, 60(7-8), 1017–1041. doi:10.1016/j.neuropharm.2010.10.022 PMID:21036182 Nikolov, R. N., Bearss, K. E., Lettinga, J., Erickson, C., Rodowski, M., Aman, M. G., ... Arnold, L. E. (2009). Gastrointestinal symptoms in a sample of children with pervasive developmental disorders. Journal of Autism and Developmental Disorders, 39(3), 405–413. doi:10.100710803-008-0637-8 PMID:18791817 Oldenberg, L. (2004). Use of yoga with occupational therapy. International Journal of Yoga Therapy, 12, 71–77. Oquendo, M. A., & Mann, J. J. (2000). The biology of impulsivity and suicidality. The Psychiatric Clinics of North America, 23(1), 11–25. doi:10.1016/S0193-953X(05)70140-4 PMID:10729928 Or-Bach, C. (2018). A yoga therapy model for supporting cancer patients: Weaving pancha-kosha and vijnana yoga principles. Retrieved from https://vijnanayoga.com/wp-content/uploads/2018/12/A-YogaTherapy-Model-for-Supporting-Cancer-Patients.pdf Paul, R., Chawarska, K., Cicchetti, D., & Volkmar, F. (2008). Language outcomes of toddlers with autism spectrum disorders: A two year follow‐up. Autism Research, 1(2), 97–107. doi:10.1002/aur.12 PMID:19360656 Peck, H. L., Kehle, T. J., Bray, M. A., & Theodore, L. A. (2005). Yoga as an intervention for children with attention problems. School Psychology Review, 34(3), 415–424. Phetrasuwan, S., Miles, M. S., Mesibov, G. B., & Robinson, C. (2009). Defining autism spectrum disorders. Journal for Specialists in Pediatric Nursing, 14(3), 206–209. doi:10.1111/j.1744-6155.2009.00200.x PMID:19614831 Porter, J. (2013). Yoga as an effective behavioral intervention for children diagnosed with an autism spectrum disorder. Graduate Annual, 1(1), 9. Proske, U., Schaible, H. G., & Schmidt, R. F. (1988). Joint receptors and kinaesthesia. Experimental Brain Research, 72(2), 219–224. doi:10.1007/BF00250245 PMID:3066633

386

 Yoga

Radhakrishna, S. (2010). Application of integrated yoga therapy to increase imitation skills in children with autism spectrum disorder. International Journal of Yoga, 3(1), 26. doi:10.4103/0973-6131.66775 PMID:20948898 Radhakrishna, S., Nagarathna, R., & Nagendra, H. R. (2010). Integrated approach to yoga therapy and autism spectrum disorders. Journal of Ayurveda and Integrative Medicine, 1(2), 120. doi:10.4103/09759476.65089 PMID:21836799 Ramanathan, M., Eswari, R., Bhavanani, A. B., Prathima, G. S., & Sanguida, A. (2019). Yoga training enhances auditory and visual reaction time in children with autism spectrum disorder: A case-control study. SBV Journal of Basic Clinical and Applied Health Science, 2(1), 8–13. doi:10.5005/jp-journals-10082-02103 Rao, J., Metri, K. G., Singh, A., & Nagaratna, R. (2016). Effect of integrated approach of Yoga therapy on chronic constipation. Voice of Research, 5, 23–26. Raznahan, A., Toro, R., Daly, E., Robertson, D., Murphy, C., Deeley, Q., Bolton, P. F., Paus, T., & Murphy, D. G. (2010). Cortical anatomy in autism spectrum disorder: An in vivo MRI study on the effect of age. Cerebral Cortex (New York, N.Y.), 20(6), 1332–1340. doi:10.1093/cercor/bhp198 PMID:19819933 Remington, G., Sloman, L., Konstantareas, M., Parker, K., & Gow, R. (2001). Clomipramine versus haloperidol in the treatment of autistic disorder: A double-blind, placebo-controlled, crossover study. Journal of Clinical Psychopharmacology, 21(4), 440–444. doi:10.1097/00004714-200108000-00012 PMID:11476129 Ressler, K. J., & Nemeroff, C. B. (1999). Role of norepinephrine in the pathophysiology and treatment of mood disorders. Biological Psychiatry, 46(9), 1219–1233. doi:10.1016/S0006-3223(99)00127-4 PMID:10560027 Rianda, D., Agustina, R., Setiawan, E. A., & Manikam, N. R. M. (2019). Effect of probiotic supplementation on cognitive function in children and adolescents: A systematic review of randomised trials. Beneficial Microbes, 10(8), 873–882. doi:10.3920/BM2019.0068 PMID:31965841 Richards, C., Moss, J., Nelson, L., & Oliver, C. (2016). Persistence of self-injurious behaviour in autism spectrum disorder over 3 years: A prospective cohort study of risk markers. Journal of Neurodevelopmental Disorders, 8(1), 21. doi:10.118611689-016-9153-x PMID:27158272 Ridderinkhof, A., de Bruin, E. I., Blom, R., & Bögels, S. M. (2018). Mindfulness-based program for children with autism spectrum disorder and their parents: Direct and long-term improvements. Mindfulness, 9(3), 773–791. doi:10.100712671-017-0815-x PMID:29875881 Rosenblatt, L. E., Gorantla, S., Torres, J. A., Yarmush, R. S., Rao, S., Park, E. R., Denninger, J. W., Benson, H., Fricchione, G. L., Bernstein, B., & Levine, J. B. (2011). Relaxation response–based yoga improves functioning in young children with autism: A pilot study. Journal of Alternative and Complementary Medicine (New York, N.Y.), 17(11), 1029–1035. doi:10.1089/acm.2010.0834 PMID:21992466 Ross, A., & Thomas, S. (2010). The health benefits of yoga and exercise: A review of comparison studies. Journal of Alternative and Complementary Medicine (New York, N.Y.), 16(1), 3–12. doi:10.1089/ acm.2009.0044 PMID:20105062

387

 Yoga

Russell, A. J., Jassi, A., Fullana, M. A., Mack, H., Johnston, K., Heyman, I., Murphy, D. G., & Mataix‐Cols, D. (2013). Cognitive behavior therapy for comorbid obsessive‐compulsive disorder in high‐functioning autism spectrum disorders: A randomized controlled trial. Depression and Anxiety, 30(8), 697–708. doi:10.1002/da.22053 PMID:23389964 Sadock, B. J., Sadock, V. A., & Kaplan, H. I. (2009). Kaplan and sadock’s concise textbook of child and adolescent psychiatry. Lippincott Williams & Wilkins. Schreck, K. A., Mulick, J. A., & Smith, A. F. (2004). Sleep problems as possible predictors of intensified symptoms of autism. Research in Developmental Disabilities, 25(1), 57–66. doi:10.1016/j. ridd.2003.04.007 PMID:14733976 Schumann, C. M., Barnes, C. C., Lord, C., & Courchesne, E. (2009). Amygdala enlargement in toddlers with autism related to severity of social and communication impairments. Biological Psychiatry, 66(10), 942–949. doi:10.1016/j.biopsych.2009.07.007 PMID:19726029 Scroggins, M. L., Litchke, L. G., & Liu, T. (2016). Effects of multisensory yoga on behavior in a male child with Apert and Asperger syndrome. International Journal of Yoga, 9(1), 81. doi:10.4103/09736131.171716 PMID:26865777 Shaaban, S. Y., El Gendy, Y. G., Mehanna, N. S., El-Senousy, W. M., El-Feki, H. S., Saad, K., & ElAsheer, O. M. (2018). The role of probiotics in children with autism spectrum disorder: A prospective, open-label study. Nutritional Neuroscience, 21(9), 676–681. doi:10.1080/1028415X.2017.1347746 PMID:28686541 SidAhmed-Mezi, M., Pumain, R., Louvel, J., Sokoloff, P., & Laschet, J. (2010). New therapeutic targets to develop molecules active in drug‐resistant epilepsies. Epilepsia, 51, 43–47. doi:10.1111/j.15281167.2010.02608.x PMID:20618399 Simonoff, E., Pickles, A., Charman, T., Chandler, S., Loucas, T., & Baird, G. (2008). Psychiatric disorders in children with autism spectrum disorders: Prevalence, comorbidity, and associated factors in a population-derived sample. Journal of the American Academy of Child and Adolescent Psychiatry, 47(8), 921–929. doi:10.1097/CHI.0b013e318179964f PMID:18645422 Smith, A. M., King, J. J., West, P. R., Ludwig, M. A., Donley, E. L., Burrier, R. E., & Amaral, D. G. (2019). Amino acid dysregulation metabotypes: Potential biomarkers for diagnosis and individualized treatment for subtypes of autism spectrum disorder. Biological Psychiatry, 85(4), 345–354. doi:10.1016/j. biopsych.2018.08.016 PMID:30446206 Sotoodeh, M. S., Arabameri, E., Panahibakhsh, M., Kheiroddin, F., Mirdoozandeh, H., & Ghanizadeh, A. (2017). Effectiveness of yoga training program on the severity of autism. Complementary Therapies in Clinical Practice, 28, 47–53. doi:10.1016/j.ctcp.2017.05.001 PMID:28779937 Spence, S. J. (2004). The genetics of autism. Seminars in Pediatric Neurology, 11(3), 196–204. doi:10.1016/j.spen.2004.07.003 PMID:15575414

388

 Yoga

Stahlberg, O., Soderstrom, H., Rastam, M., & Gillberg, C. (2004). Bipolar disorder, schizophrenia, and other psychotic disorders in adults with childhood onset AD/HD and/or autism spectrum disorders. Journal of Neural Transmission (Vienna, Austria), 111(7), 891–902. doi:10.100700702-004-0115-1 PMID:15206005 Stephens, I. (2017). Medical yoga therapy. Children (Basel, Switzerland), 4(2), 12. doi:10.3390/children4020012 PMID:28208599 Stewart, P. A., Hyman, S. L., Schmidt, B. L., Macklin, E. A., Reynolds, A., Johnson, C. R., James, S. J., & Manning-Courtney, P. (2015). Dietary supplementation in children with autism spectrum disorders: Common, insufficient, and excessive. Journal of the Academy of Nutrition and Dietetics, 115(8), 1237–1248. doi:10.1016/j.jand.2015.03.026 PMID:26052041 Stone, W. L., & Yoder, P. J. (2001). Predicting spoken language level in children with autism spectrum disorders. Autism, 5(4), 341–361. doi:10.1177/1362361301005004002 PMID:11777253 Strömland, K., Nordin, V., Miller, M., Akerström, B., & Gillberg, C. (1994). Autism in thalidomide embryopathy: A population study. Developmental Medicine and Child Neurology, 36(4), 351–356. doi:10.1111/j.1469-8749.1994.tb11856.x PMID:8157157 Telles, S., & Naveen, K. V. (1997). Yoga for rehabilitation: An overview. Indian Journal of Medical Sciences, 51(4), 123–127. PMID:9355699 Thapar, A., Cooper, M., & Rutter, M. (2017). Neurodevelopmental disorders. The Lancet. Psychiatry, 4(4), 339–346. doi:10.1016/S2215-0366(16)30376-5 PMID:27979720 Thurm, A., Lord, C., Lee, L. C., & Newschaffer, C. (2007). Predictors of language acquisition in preschool children with autism spectrum disorders. Journal of Autism and Developmental Disorders, 37(9), 1721–1734. doi:10.100710803-006-0300-1 PMID:17180717 Tomova, A., Husarova, V., Lakatosova, S., Bakos, J., Vlkova, B., Babinska, K., & Ostatnikova, D. (2015). Gastrointestinal microbiota in children with autism in Slovakia. Physiology & Behavior, 138, 179–187. doi:10.1016/j.physbeh.2014.10.033 PMID:25446201 Towbin, K. E. (2005). Pervasive developmental disorder not otherwise specified. In F. R. Volkmar, R. Paul, A. Klin, & D. Cohen (Eds.), Diagnosis, development, neurobiology, and behavior (pp. 165–200). John Wiley & Sons Inc. Valicenti-McDermott, M., McVicar, K., Rapin, I., Wershil, B. K., Cohen, H., & Shinnar, S. (2006). Frequency of gastrointestinal symptoms in children with autistic spectrum disorders and association with family history of autoimmune disease. Journal of Developmental and Behavioral Pediatrics, 27(2), S128–S136. doi:10.1097/00004703-200604002-00011 PMID:16685179 Vidyashree, H. M., Maheshkumar, K., Sundareswaran, L., Sakthivel, G., Partheeban, P. K., & Rajan, R. (2019). Effect of yoga intervention on short-term heart rate variability in children with autism spectrum disorder. International Journal of Yoga, 12(1), 73. doi:10.4103/ijoy.IJOY_66_17 PMID:30692787

389

 Yoga

Volkmar, F. R., & Nelson, D. S. (1990). Seizure disorders in autism. Journal of the American Academy of Child and Adolescent Psychiatry, 29(1), 127–129. doi:10.1097/00004583-199001000-00020 PMID:2295565 Wallace, G. L., Dankner, N., Kenworthy, L., Giedd, J. N., & Martin, A. (2010). Age-related temporal and parietal cortical thinning in autism spectrum disorders. Brain, 133(12), 3745–3754. doi:10.1093/ brain/awq279 PMID:20926367 Weismer, S. E., Lord, C., & Esler, A. (2010). Early language patterns of toddlers on the autism spectrum compared to toddlers with developmental delay. Journal of Autism and Developmental Disorders, 40(10), 1259–1273. doi:10.100710803-010-0983-1 PMID:20195735 Wetherby, A. M., Woods, J., Allen, L., Cleary, J., Dickinson, H., & Lord, C. (2004). Early indicators of autism spectrum disorders in the second year of life. Journal of Autism and Developmental Disorders, 34(5), 473–493. doi:10.100710803-004-2544-y PMID:15628603 White, S. W., Keonig, K., & Scahill, L. (2007). Social skills development in children with autism spectrum disorders: A review of the intervention research. Journal of Autism and Developmental Disorders, 37(10), 1858–1868. doi:10.100710803-006-0320-x PMID:17195104 Whiteley, P., Haracopos, D., Knivsberg, A. M., Reichelt, K. L., Parlar, S., Jacobsen, J., Seim, A., Pedersen, L., Schondel, M., & Shattock, P. (2010). The ScanBrit randomised, controlled, single-blind study of a gluten-and casein-free dietary intervention for children with autism spectrum disorders. Nutritional Neuroscience, 13(2), 87–100. doi:10.1179/147683010X12611460763922 PMID:20406576 Yankowitz, L. D., Herrington, J. D., Yerys, B. E., Pereira, J. A., Pandey, J., & Schultz, R. T. (2020). Evidence against the “normalization” prediction of the early brain overgrowth hypothesis of autism. Molecular Autism, 11(1), 1–17. doi:10.118613229-020-00353-2 PMID:32552879 Zwaigenbaum, L., Thurm, A., Stone, W., Baranek, G., Bryson, S., Iverson, J., ... Rogers, S. (2007). Studying the emergence of autism spectrum disorders in high-risk infants: Methodological and practical issues. Journal of Autism and Developmental Disorders, 37(3), 466–480. doi:10.100710803-006-0179-x PMID:16897376

390

391

Chapter 22

Yoga and Polyvagal Theory:

Effects and Applications for Well-Being Marlysa Sullivan Maryland University of Integrative Health, USA

ABSTRACT This chapter explores yoga as a salutogenic intervention supportive of eudaimonic well-being with its wide-spread health effects for various patient populations. Autonomic nervous system regulation and resilience are considered as important meditators for the promotion of biopsychosocial health. Polyvagal theory offers a novel perspective on how underlying neural platforms support combined physiological, psychological, and behavioral states—inclusive of eudaimonic well-being. This chapter describes the convergence of neurophysiological ideas of neuroception, interoception, and neural platforms with yoga foundational concepts such as discriminative wisdom and the gunas. This translatory language of eudaimonic well-being and polyvagal theory offers a framework for yoga to be understood and integrated into current healthcare and research contexts while maintaining its unique perspective and foundational wisdom.

INTRODUCTION: CONCEPTUALIZING WELL-BEING Health and well-being can be considered from two complementary perspectives. The pathogenic viewpoint looks to identify specific causes of a condition or disease. Interventions are then identified to target the pathology. Alternatively, the salutogenic perspective seeks to explore and then direct interventions to support the contributors of optimal health. These two perspectives, salutogenesis and pathogenesis, can be utilized together to facilitate both the alleviation of pathology and the optimization of health (Sullivan & Robertson, 2020). Yoga Therapy aligns with this salutogenic perspective as it addresses the many contributors to complex or chronic conditions and seeks to promote self-management and lifestyle change. Ananda Balayogi Bhavanani (DSc (Yoga), Director CYTER of Sri Balaji Vidyapeeth) teaches this importance of yoga therapy as a salutogenic intervention as opposed to applying the practices for symptom management from a pathogenic perspective- calling the latter yogopathy (Bhavanani, 2011; Bhavanani et al., 2019). When DOI: 10.4018/978-1-7998-3254-6.ch022

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Yoga and Polyvagal Theory

yoga therapy is practiced consistent with this salutogenic framework the tools (e.g. asana, pranayama, meditation, ethical principles) are applied synergistically to uncover the sources of suffering for the promotion of greater well-being. Eudemonic well-being is a term that is useful for the practical application of the salutogenic perspective. Additionally, this term offers a translatory language for the wisdom teachings of yoga to be understood in current healthcare/scientific contexts. This term comes from Aristotle’s teachings of eudemonia describing the sentiment of a person looking back on their life and feeling as if they reached their optimal potential and lived an excellent, purpose-filled life (Aristotle et al., 2004). Two types of happiness are described by Aristotle- eudemonic and hedonic. Eudemonic happiness stems from this realizing of one’s potential and is described as a non-transitory, steadfast contentment or joy. Hedonic happiness is considered a more transitory emotional experience, such as cheerfulness (Aristotle et al., 2004). The significance of eudemonic well-being is demonstrated by the growing body of research that supports its widespread and positive health effects for various patient populations. These benefits include: 1. Lower levels of fatigue, disability, pain intensity, pain medication use, depressive symptoms, and improved patient functioning and life satisfaction in people with chronic pain conditions (Dezutter et al., 2013; Dezutter et al., 2015; Schleicher et al., 2005). 2. Decreased all-cause mortality independent of factors such as age, physical inactivity, conditions such as cardiovascular disease or cancers (Keyes & Simoes, 2012; Pressman et al., 2019). 3. Eudemonic well-being is differentiated from hedonic wellbeing with its effect on inflammatory and immune processes. While eudemonic well-being is associated with a down-regulation of gene profiles related to healthier inflammatory and immune processes affecting neurodegenerative, cardiovascular conditions and cancers- hedonic well-being is associated with its up-regulation (Cole et al., 2015; Fredrickson et al., 2013; Fredrickson et al., 2015). 4. Eudemonic well-being is correlated with decreased levels of salivary cortisol, pro-inflammatory cytokines, and lessened cardiovascular disease risk, hedonic well-being has minimal effect on these biomarkers (Ryff et al., 2004). These myriad and distinct health benefits demonstrate eudemonic well-being as an important contributor to physical, psychological, and social health.

YOGA AND EUDEMONIC WELL-BEING Yoga has demonstrated an effect for improving eudemonic wellbeing (Ivtzan & Papantoniou, 2014). Additionally the component measures of eudemonic well-being (Huta & Waterman, 2014) share a similar focus with foundational philosophical teachings of yoga (Sullivan, 2019; Sullivan & Robertson, 2020). Eudemonic constructs of authenticity, personal expressiveness, self-actualization, and self-realization are both foundational to and can be facilitated through yoga’s teachings. Yoga provides a methodology to foster insight into how the relationship to bodily state or condition, thoughts, emotions, and behaviors contribute to suffering. This process helps facilitate greater authenticity, self-expression, and self-actualization as the person finds positive ways to relate to their body and mind as well as to explore actions that improve well-being. 392

 Yoga and Polyvagal Theory

The emphasis on ethical inquiry in both yoga and eudemonia helps to bring the person in harmony with their values. This value-driven process empowers the individual for right action for a meaningful, purpose-filled, flourishing life and to realize their potential. The teaching of dharma in yoga encompasses this harmonious right action to support the individual, as well as people and the life circumstance in which they are situated. The yamas and niyamas (social and personal disciplines relating to interactions with others and ourselves, respectively) encompass ethical qualities which are foundational to yoga practice and help align the person with dharma (Mallinson & Singleton, 2017; Smith, 2009; Stoler-Miller, 1998). Yoga also teaches the recognition of inherent connection between all beings which, with the yamas and niyamas, support meaningful and healthy relationships—a key component of eudemonic well-being. A recent paper explored this context of eudemonic well-being in relationship to yoga therapy (Sullivan et al., 2018). Yoga’s methodology offers a path for the individual to inquire and develop insight into the relationships with the body, mind, and environment that may either promote suffering or its alleviation. The person is supported in practices to align with values, meaning, and purpose such that steadfast contentment and equanimity is realized even amidst adverse or difficult life circumstances. Resultantly the person experiences eudemonic well-being and a healthy, adaptive relationship to the body, mind, and life circumstance (Sullivan, 2019; Sullivan et al., 2018; Sullivan & Robertson, 2020).

WELL-BEING AND THE AUTONOMIC NERVOUS SYSTEM (ANS) The ANS serves as a mediator between the external environment, bodily state, thoughts and emotions (Cole, 2013; Cole, 2014; Porges, 2011). As such, the ANS is key to creating synchronistic change in the body-mind and behavior. Positive states such as eudemonic well-being may be catalyzed through autonomic regulation and its support of improved biopsychosocial health. This chapter explores the function of the ANS and polyvagal theory to illuminate these connections between physiological, psychological, and behavioral states for the cultivation of eudemonic well-being and improved whole-person health. The convergence with yoga’s philosophy, teachings, and practices will also be described to explore its unique framework to support these intentions and goals.

The ANS as a Link Between Body, Mind and Behavior The ANS is responsible for evaluating and generating appropriate responses to what is perceived in the body, mind, and environment. Once information from the environment is picked up by the five senses it is integrated with information about bodily state (physiological activations and processes) and the mind (e.g. thoughts, emotions, beliefs, memories). This information is processed, evaluated, and assessed to create an integrated physiological, psychological, and behavioral response. The ANS is thus crucial to healthy, adaptive responses to stimuli as it connects our perception to the environment with activation of the systems of the body (e.g. cardiovascular, respiratory, digestive, endocrine, immune), psychological processes, and behavior. The response of the ANS can be conceptualized as a continuum of integrated physiological, psychological, behavioral responses from states of activation and mobilization to restoration and relaxation. Optimal function includes the ability to respond accurately and appropriately to actual or perceived threat and to return to states of homeostasis, stability, and restoration efficiently.

393

 Yoga and Polyvagal Theory

The two broad divisions for these integrated actions are termed the sympathetic and parasympathetic nervous systems. The sympathetic is often referred to as “fight or flight” and is responsible for the activation of physiological systems for quick, efficient response to actual or perceived danger. The parasympathetic, referred to as “rest and digest”, is responsible for the activation of physiological restoration and homeostasis in response to perceived safety. The primary conduit of the parasympathetic nervous system is the vagus nerve. As a mixed nerve, it carries both afferent (sensory) and efferent (motor) functions. It is thought that about 80 percent of the vagus nerve has this sensory function, while the remaining 20 percent is responsible for the majority of parasympathetic response (Porges, 2011). A major significance of this nerve is this shared capacity for both perception of inner physiological state as well as the initiation of the action termed the “relaxation response”. The sensory component of the vagus nerve is part of a larger process called interoception which is important for promoting healthy states of the body, mind, and behavior. Historically, interoception referred to the perception of visceral information and inner bodily states such as heart rate and digestive processes (Ceunen et al., 2016). More recently, its meaning has broadened and includes a more complex process. This includes sensory pathways (including the vagus nerve) carrying perception of physiological state and psychological processes (thoughts, emotions, memories) for an integrated motor response of the body, mind, and behavior (Ceunen et al., 2016; Craig, 2015; Porges, 2011). Healthy interoceptive processes indicates the person is able to accurately perceive, interpret, and respond to stimuli for optimal body-mind regulation and resilience (Farb et al., 2015; Haase et al., 2016). The ability for an individual to accurately match the response of the body and mind to the perceived stimuli is thought to be important for recovery and well-being in various patient populations such as pain conditions and addiction (Ceunen et al., 2016; Craig, 2015; Farb et al., 2015; Haase et al., 2016; Porges, 2011; Strigo & Craig, 2016). The vagus nerve is important to the development of healthy interoception as it relays information about viscera and bodily state to the brain. The motor component of the vagus nerve is responsible for majority of the parasympathetic activation of the body. Its widespread connections throughout the body makes it crucial to systems-wide homeostatic responses including cardiovascular, respiratory, digestive, muscular and immune (Porges, 2011). The motor function of the vagus nerve has two source nuclei in the brainstem—the dorsal motor nucleus and the nucleus ambiguus—which provide for a range of parasympathetic responses as will be described below with polyvagal theory (Porges, 2011). These extensive motor and sensory functions make the vagus nerve a crucial bidirectional pathway for the ANS to support body-mind-behavior regulation, resilience, and biopsychosocial health.

ANS Regulation, Resilience, and the Promotion of Well-being The term regulation describes the capacity to return to or maintain homeostatic, stable states of the body and mind through managing one’s responses to threat or adversity (Gard et al., 2014). Self-regulation refers to the individual having the capacity to consciously activate these homeostatic physiological states, calm and positive psychological qualities, and pro-social behaviors such as compassion, connection and empathy. The individual learns to recognize, manage, and even alter thoughts, emotions, behavior, and physiological states in response to various stimuli. This ability to self-regulate is an important contributor to improved health and wellbeing in various conditions including: irritable bowel syndrome, anxiety, neurodegenerative disorders, trauma, and pain (Muehsam et al., 2017; Taylor et al., 2010). 394

 Yoga and Polyvagal Theory

The ANS is vital to self-regulation as it includes widespread connections to bodily systems, psychological processes, and behavioral characteristics. The significance of these extensive connections is the capacity for concurrent effects on the body, mind, and behavior by affecting the underlying autonomic state. Both top-down neuro-cognitive (e.g. practices for control of attention or regulation of emotion) and bottom-up neurophysiological (e.g. practices of movement or breath control) processes can be utilized to shift autonomic state for self-regulation. This ability to regulate the ANS is associated with positive attentional and emotional regulation and control as well as appropriate or accurate responses to environmental demands (Park & Thayer, 2014; Porges, 2011; Streeter et al., 2012; Strigo & Craig, 2016; Thayer & Lane, 2000). Lower capacity for ANS regulation is associated with poor self-regulation and adverse health outcomes or conditions such as depression, anxiety, lupus, rheumatoid arthritis, trauma (Muehsam et al., 2017; Park & Thayer, 2014; Thayer & Lane, 2000; Tsuji et al., 1994). Various conditions including fibromyalgia, irritable bowel syndrome, headaches, and other chronic pain conditions include different expression of ANS dysregulation (Azam et al., 2016; Barakat et al., 2012; Koenig et al., 2016; Kolacz & Porges, 2018; Maletic & Raison, 2009; Meeus et al., 2013; Staud, 2008; Tracy et al., 2016). ANS dysregulation includes excessive, inadequate, or inaccurate sympathetic or parasympathetic responses to the demands of the environment. This altered capacity to adequately respond to stimuli can be both a contributor as well as a result of chronic health conditions. In other words, chronic conditions can lead to diminished resilience and ANS dysregulation and/or this dysregulation may be a sign of the diminished resilience resulting from the condition (Sullivan, 2019). Resilience adds a factor of efficiency and timeliness to the concept of regulation. A resilient system is one that recovers quickly from difficulty or adverse stimuli while conserving physiological and psychological resources (Resnick et al., 2011; Tugade & Fredrickson, 2004; Whitson et al., 2016). The more resilient the individual, the more likely they are to respond to stressors appropriately, adequately, and return quickly to regulated body-mind-behavioral states. The positive health outcomes of resilience includes less perceived stress, greater motivation in recovery from illness or trauma, improved management of chronic pain and dementia (Resnick et al., 2011; Tugade & Fredrickson, 2007). Polyvagal theory offers a nuanced understanding of how the ANS affects body-mind-behavioral states in a synchronistic fashion for improved regulation and resilience.

POLYVAGAL THEORY: DEEPER LOOK AT THE ANS RELATIONSHIPS BETWEEN BODY, MIND, AND BEHAVIOR The sympathetic and parasympathetic divisions of the ANS are not meant to be understood as a dichotomous system where one branch is activated at the expense of the other. They are not antagonistic and do not operate as a simple “turned on or off system”. When one activates it does not mean the other automatically “shuts off” or deactivates. Rather the branches work together to create a continuum of fine-tuned responses to the environment (Berntson et al., 1991; Porges, 2011). This complexity allows the individual to have a range of responses from mobilization to restorative or homeostatic states. The polyvagal theory, developed by Stephen W. Porges Ph.D. (Neuroscientist, University of North Carolina and Indiana University), illuminates the connected relationships of the body, mind, and behavior influenced by underlying autonomic neural states or “neural platforms”. Each neural platform activates the likelihood of particular physiological activations, psychological processes, and behavioral characteristics. Recognizing and altering these underlying neural platforms provides a mechanism for 395

 Yoga and Polyvagal Theory

the promotion of systemic body-mind-behavioral regulation, resilience, and whole- person health and well-being. As mentioned above, two different origination sites for the vagus nerve in the brainstem are detailed in polyvagal theory. The finding of these two different source nuclei is significant in explaining a range of parasympathetic responses. There is a continuum of activations that slow the system down spanning from a homeostatic relaxation response to an extreme slowing or shutting down to essential survival needs and conservation—seen in situations of immense life threat. This complexity of autonomic function spanning a range of parasympathetic activations to co-activation of the sympathetic and parasympathetic branches in differing proportions as well as the effect of these neural platforms on body-mind-behavior are key to the importance of polyvagal theory. Neuroception, coined by Dr. Porges, describes a process of subconscious perception of both inner processes (interoception) and the external environment (exteroception) that individuals use to gauge the present level of safety or danger (Porges, 2003, 2004, 2011). These interoceptive and exteroceptive inputs help to activate an autonomic neural platform that readies the person for a quick, efficient response to the environment. A key piece to neuroception is how this subconscious detection and subsequent neural platform gives rise to particular and connected physiological, psychological and behavioral responses before any conscious attention or deliberation. As will be discussed below, yoga practices empower the person to recognize these habitual tendencies and provide strategies to help the individual change the reaction and relationship to subconscious stimuli. Without this conscious attention, these neural platforms give rise to instinctive reactions of physiological state, thoughts, emotions, and behavior. Polyvagal theory describes how these divisions of the vagus nerve, the neural platforms arising from neuroception, and the activation of neural platforms influence shared physiological, psychological, behavioral characteristics. The three main autonomic neural platforms are named according to the behaviors that arise from their activation. They are described below to include the physiological foundation and effect on psychological and behavioral attributes.

Social Engagement Neural Platform This neural platform is made up of a network of nerves including: the branch of the vagus nerve originating from the nucleus ambiguous in the brainstem as well as the glossopharyngeal, spinal accessory, trigeminal, and facial nerve. This branch of the vagus is responsible for the production of respiratory sinus arrhythmia- the measure of heart rate variability used to demonstrate a parasympathetic dominant and resting state of the heart. This neural platform is activated in response to perceived safety in the environment. It is termed social engagement because it links physiological restoration (slowed heart rate, greater HRV, improved peristalsis for digestion) and expressive and receptive domains of communication (motor control of facial expression to express safety to others, control of middle ear muscles to attune to human voice distinct from other noise, control of muscles of the throat for improved vocalization pleasing to other humans). The result of this activation is a combination of physiological and psychological calm enabling the potential for positive prosocial behaviors and connection with others including compassion and eudemonic well-being (Porges & Carter, 2017; Porges, 2011, 2017; Sullivan, 2019; Sullivan et al., 2018).

396

 Yoga and Polyvagal Theory

Defensive Mobilization Neural platform This neural platform involves activation of the sympathetic nervous system with defensive behaviors often referred to as “fight or flight”. In response to perceived threat, the body’s resources are mobilized for defensive strategies and include: increased muscle tone, heart rate, release of stress related hormones such as cortisol, and inhibition of digestive processes. The brain is also activated for greater vigilance and focused attention to potential threats. Alongside these physiological activations, emotions that support these protective responses such as fear, anger, anxiety, or worry and behaviors of “fight or flight” are likely to emerge. In sum, in response to perceived threat this autonomic neural platform supports mobilization of the system to include a physiological, emotional, behavioral response to assist in the finding of safety and protection (Porges, 2011; Sullivan & Robertson, 2020).

Defensive Immobilization This neural network is thought to be activated in response to extreme danger, threat, or terror. It results in a dramatic slowing of the body’s resources to the minimum necessary for survival. The branch of the vagus nerve originating out of the dorsal motor nucleus is thought to be part of this response to extreme threat. Sometimes referred to as death-feigning it signifies a passive response to life threat and is a combined state of physiological immobilization or shutdown with emotional and behavioral states including disembodied or dissociative characteristics. These three primary neural platforms arise from the predominance of either the sympathetic mobilizing activations or one of the vagal branches precipitating a slowing down of the system along a continuum of optimal restoration to immobilization. Polyvagal theory describes two more states that arise from the combination of these three neural platforms to create a wide and nuanced array of possibilities for human behavior and response. Two more neural platforms from the combination of the above: 1. Safe Mobilization When the person simultaneously perceives safety and the need for mobilization there is a co-activation of the neural platforms of social engagement and defensive mobilization. This is seen in situations such as dance, play, exercise, or creative endeavors. The person is both mobilized for action and perceives safety so there is relative balance in the neural platforms of social engagement and mobilization. As a result the person experiences safety in the midst of activation. Safe mobilization is a key to improving regulation and resilience. The person finds a greater capacity to find safety within an activated system. This functions to create a wider tolerance to life’s stressors and stimuli. Rather than becoming overwhelmed by stimuli, the person learns to manage stressors more effectively and efficiently. Yoga practices that mobilize the physiological systems such as active postures alongside relaxed breath patterns, or calming visualizations or affirmations are examples of this state of safe mobilization. This combination of activity of the body with attention to breath, thought, emotions enables the person to find safety amidst activation. This learning in yoga practice can be extended to the individual’s life situations to mirror those that require a level of activation while maintain relative calm and safety.

397

 Yoga and Polyvagal Theory

2. Safe Immobilization Co-activation of the social engagement and immobilization neural platforms arise from a perception of safety alongside a slowing down of the system. This would be present in certain meditative states where physiological resources are slowed and the person may even experience levels of dissociation but safety is perceived (Porges, 2011). Altogether these five global states provide a continuum of autonomic response to help the individual both mobilize and find homeostatic balance across a range of situations of perceived safety or danger. Self-regulation is fostered through activation of the social engagement neural platform for shared physiological, psychological, and behavioral states of homeostasis, peacefulness and calm. Resilience can then be supported by helping the person find greater ease in moving between neural platforms and cultivating a broader range of safe mobilization to safe immobilization. This will be described in more detail below. Yoga is one mind-body practice that can help the individual learn to recognize and alter these underlying autonomic states. Yoga practices such as meditation and ethical inquiry provide top-down neurocognitive practices while movement and breath control provide bottom-up neurophysiological practices for this effect on the ANS. Integration of such practices supports autonomic regulatory states, parasympathetic activation, and an integrated autonomic, emotional and attentional effect (Chu et al., 2017; Dale et al., 2011; Mackenzie et al., 2014; Sarang & Telles, 2006; Telles et al., 2016; Tyagi & Cohen, 2016). The yoga practitioner becomes empowered in their capacity to influence physiological activations, thoughts, emotions, and behavior through the recognition and change of autonomic states. This capacity for self-regulation and resilience demonstrates a primary mechanism through which yoga positively affects physical and mental health for various patient populations (Gard et al., 2014; Muehsam et al., 2017; Schmalzl et al., 2015; Streeter et al., 2012; Sullivan et al., 2018; Taylor et al., 2010).

YOGA AND THE ANS: SHARED THEMES Yoga and Self-Regulation, Interoception and Discriminative Wisdom Essential to yoga teaching is the recognition of suffering arising from the misperception and misidentification with the fluctuating stimuli of the body, mind, and environment. Yoga teaches the person to compassionately notice the relationship and reaction to what arises and to distinguish these stimuli from a deeper recognition of unchanging awareness. While awareness experiences the rise and fall of bodily sensation (physiological activations/sensations), mental and emotional phenomena (thoughts, emotions, ego-identity), and life’s situations—it is separate from them. This realization brings about an underlying equanimity, contentment, and abiding joy that is present amidst all circumstances and fluctuations of the body, mind, and environment. Discriminative wisdom is a term to describe this insight into the distinction between changing stimuli and unchanging awareness. Nonjudgmental, compassionate observation and exploration of body-mind sensation and life circumstance enables the person to identify habitual patterns of response and behavior that contribute to suffering. The person develops new, adaptive, healthy relationships to body, mind, and environmental stimuli to alleviate suffering and support equanimity and well-being amidst all stimuli. Consequently, the person develops a greater capacity to navigate the changing stimuli and situations of life while remaining situated in an unwavering equanimity that arises from this recognition of awareness. 398

 Yoga and Polyvagal Theory

As an example, the person might become aware of how particular life circumstances, such as a work situation, create a change in the state of their body and/or mind. A certain stressor may be linked to a dysfunctional breathing pattern, tension in the body, worry or anxiety. Yoga teaches the person to both notice these patterns and to manage or alter these reactions to support healthier, more adaptive states of the body-mind. Skills are learned to inquire into the causes of body-mind disturbance and to mitigate them for improved regulation and resilience, and greater well-being. Ultimately the person learns to work with the natural stressors and changes of life with greater skill and ease. The previously described neurophysiological concepts of neuroception and interoception are helpful in understanding the practical application of discriminative wisdom in yoga. Neuroception describes the habitual activations of neural platforms based on subconscious detection of safety or danger. Yoga teaches a process of inquiry to support these subconscious reactions being brought to conscious attention and control. This recognition of habitual reactions enables the person to regulate these neural activations, support experiences of equanimity and eudemonic well-being, and to choose healthy, adaptive responses. Interoception is a key piece to developing this skill of noticing subconscious physiological activations and altering or regulating one’s response. As explained above, interoception is a complex concept that includes the recognition of bodily states, relationships to thoughts, emotions, beliefs, and resultant behavior (Craig, 2015; Farb et. al., 2015; Porges, 1993; Strigo & Craig, 2016). Yoga teaches the individual to notice and tease apart the various stimuli of the body, mind, and environment. They develop insight to support the navigation of these stimuli with greater ease and well-being. The person learns to notice what arises, to self-reflect, and self-discover so that new relationships can be fostered for improved regulation and resilience. Foundational to yoga is a learned methodology to accurately discern body, mind, and environmental stimuli; to bring subconscious or unconscious habits to conscious attention; and to build healthier and more accurate relationships to these stimuli. The person is able to build interoceptive skills to distinguish between physiological and psychological activations and their relationships with one another for the cultivation of healthier responses. This learning of interoception is key to the explanatory framework of yoga whereby the person learns to distinguish body, mind, environmental phenomena from awareness for the promotion of eudemonic well-being and the alleviation of suffering (Sullivan et al., 2018; Sullivan et al., 2018).

Yoga and Neural Platforms: The Gunas In addition to building interoceptive skills and helping neuroceptive processes become conscious, yoga offers its own perspective on how underlying substrates influence connected body-mind states and behavior. Yoga teaches that all changing phenomena (including the components of the body, mind, and environment) are made up of three qualities—the gunas. These gunas—sattva, rajas, tamas—give rise to all of the various manifestations of the body state, mental and emotional characteristics, behavior, and environmental phenomena. Sattva is the quality of lightness, buoyancy, luminosity and supports bodily states of homeostasis and relaxation; psychological qualities of peace, equanimity, and calm; and behaviors that align with compassion, connection, and one’s values and ethics (Burley, 2012; Miller, 2012; Sullivan & Robertson, 2020). Rajas is the quality of mobilization and is responsible for the wide range of activations in the body, mind, and behavior. This includes a continuum from creativity, play, and inspiration to turbulence, fight/flight responses, anxiety, fear, and ager. The third quality, tamas, encourages states of stillness along a continuum from groundedness and stability to fatigue, delusion, obscuration, and dissociation. 399

 Yoga and Polyvagal Theory

When the gunas are balanced sattva provides the clarity for discriminative wisdom and insight, including developing interoceptive skills, and inner and outer harmony in alignment with one’s values for a meaningful and purpose-filled life. Rajas provides the motivation and creativity for change and determination of right action. Tamas offers the stability and groundedness for disciplined practical action and practice. When rajas is predominant over the other two gunas activation and agitation builds to those body-mind states seen in the neural platform of defensive mobilization. This includes physiological, psychological, and behavioral states of fight/flight response. When tamas is predominant the stillness and slowing can eventually produce body-mind and behavioral states similar to the neural platform of defense immobilization such as intense fatigue, dissociation, or obscuration. Polyvagal theory offers a neurophysiological description and convergent theory to the gunas of yoga. From these underlying neural platforms (polyvagal theory) and gunas (yoga) arise concurrent changes in physiological and psychological activations and behaviors (Sullivan et al., 2018; Sullivan & Robertson, 2020). By affecting these underlying substrates the person can enact change in the body, mind, and behavior simultaneously. While not reducing one to the other, the shared characteristics of gunas and neural platforms illustrate how they support one another. When one is activated (neural platform or guna) it is likely to be reflected in its counterpart (related guna or neural platform). In other words, a guna is supportive of a particular neural platform and vice versa and affecting these substrates produces a synergistic effect on physiology, emotion, and behavior. This translatory language is important to the application and understanding of yoga for regulation and resilience as well as the promotion of whole-person health and eudemonic well-being. A more complete description of these convergences of neural platforms and gunas is described in an article by Sullivan et al. (2018) and is summarized below.

Social Engagement and Sattva Guna This neural platform and guna give rise to shared physiological states such as homeostasis and relaxation; psychological states including calmness, peacefulness, and equanimity; and behavioral states such as compassion and connection. These shared attributes demonstrate that when one is activated or predominant it is likely reciprocated by the activation or predominance of the other. When sattva is predominant with its characteristics of clarity and calmness, the social engagement neural platforms is likely activated with its components of interoceptive accuracy and sensitivity, parasympathetic response, psychological calm and equanimity, and prosocial features such as compassion and connection. Likewise when the social engagement neural platform is activated the attributes of sattva such as clarity, luminosity, equanimity are likely to emerge. This neural platform and guna are essential to the development of discriminative wisdom and self-regulatory skills, as it supports healthy interoceptivity, physiological and psychological calm, equanimity and prosocial connection, compassion. Moreover these qualities of healthy connection to body-mind states, positive relationships with others, and alignment with values may facilitate greater eudemonic well-being.

Mobilization (Safe and Defensive) and Raja Guna Rajas guna and sympathetic activation mobilize the system for action from a range of safe mobilization to defensive states in response to actual or perceived threat. These underlying substrates mobilize the body, mind and behavior for action on a continuum from safe and balanced to defensive and unbalanced 400

 Yoga and Polyvagal Theory

mobilization. In safe mobilization or balanced rajas there is activation of the systems of the body, mind and behavior alongside perceived safety. This state is seen in any creative endeavor, activity, or play where the systems are primed for action but the activation does not overwhelm the perception of safety. In defensive mobilization or a predominance of rajas the body, mind and behavior are mobilized for action in response to perceived or actual threat. When one is activated (rajas or sympathetic) the other is more likely to be activated as well. The continuum from safe mobilization/balanced rajas to defensive mobilization/predominant rajas allows the activation of the systems as appropriate to the environment.

Immobilization (Safe and Defensive) and Tama Guna This neural platform and guna provide relative stillness or immobilization on a continuum from solidity, groundedness and stability to dullness, inertia, and dissociative states. The stillness of tamas supports the activation of the neural platform of immobilization. In safe immobilization or balanced tamas there is both the perception of safety with stillness, as in meditative states. Defensive immobilization or predominant tama guna may result in states of immobilization, obscuration, dullness, lethargy, dissociation, death-feigning, inertia. Polyvagal theory and yoga provide a language for directing interventions towards the alteration of underlying neural platforms (polyvagal) or gunas (yoga) for whole-person health and well-being. This synergy between perspectives enables yoga to be understood for its positive health effects in various patient populations in alignment with its unique framework, perspective, and philosophical context.

BODY-MIND REGULATION, RESLIENCE THROUGH POLYVAGAL THEORY AND YOGA The gunas and polyvagal theory offer a novel perspective on how altering underlying substrates (neural platforms or gunas) can help support regulation and resilience for widespread health effects (Sullivan et al., 2018; Sullivan & Robertson, 2020). In this model, self-regulation can be conceptualized as the promotion of sattva guna and the social engagement neural platform. The practices of yoga can be directed in a synchronistic fashion towards the promotion of this neural platform and guna. This would include activation of the parasympathetic nervous system through both top-down neurocognitive and bottom-up neurophysiological practices. From a neurocognitive perspective practices of meditation, focused attention, mantra, affirmation, and ethical inquiry to align with one’s values would promote these calm, relaxed states of the body and mind as well as prosocial behavior. From a neurophysiological perspective breath practices such as alternate nostril breath, longer exhales, and movement such as restorative postures can facilitate these characteristics of sattva and social engagement neural platform. Furthermore, the components of eudemonic wellbeing can be supported by the activation of this neural platform and guna. Healthy interoceptive skills help the person notice and manage habitual, subconscious reactions of the body, mind, and behavior to various stimuli. The person becomes able to recognize the ways in which one’s actions (including physiological activations, psychological qualities and behavior) either supports or detracts from a process of self-realization and self-actualization. This insight empowers the person to live more authentically in alignment with one’s values for a flourishing, meaningful, and purpose-filled life. In addition, improved relationships with others are fostered from 401

 Yoga and Polyvagal Theory

this guna and neural platform. Sattva guna supports the inherent recognition of connection between all people and the social engagement neural platform strengthens both expressive and receptive components of communication. Lastly, sattva guna is described as supporting dharma of which the yamas and niyamas (ethical practices) are taught as essential to this harmonious living for a purpose-filled and meaningful life (Burley, 2012; Miller, 2012; Smith, 2009; Sullivan & Robertson, 2020). Likewise, a foundation of eudemonia is the recognition of the virtue ethics as guideposts to a purpose-filled and meaningful life (Aristotle et al., 2004). When all the practices of yoga are utilized for the promotion of sattva and social engagement neural platform, the person finds an improved capacity for regulation, perceived safety, equanimity, contentment, connection, and eudemonic well-being. This provides an anchor and foundation of safety and calm that the person can stay connected or return to even amidst difficult life circumstances.

Promoting Resilience Through Polyvagal Theory and Yoga While regulatory states of the body and mind are key to physical and mental health, they are not meant to be present at all times during all levels of activity. Resilience helps clarify the need for differing levels of activation for a healthy, adaptive response to life’s natural needs and stimuli. Polyvagal theory and yoga help to conceptualize the practical application of resilience as the capacity to work with changing neural platforms or gunas (Sullivan et al., 2018). The gunas and neural platforms are constantly shifting and changing, which contributes to the natural occurring fluctuations present in the body, mind, and environment. Yoga teaches that the gunas are always in motion—all phenomena of the body, mind, and environment are in constant flux. Rather than disregarding or ignoring this intrinsic part of life, yoga teaches the person to navigate and develop healthier relationships with fluctuating stimuli for the alleviation of suffering and realization of wellbeing. By learning to recognize, adapt, and develop healthy relationships to the rising and falling of sattva, rajas, and tamas the person learns to navigate these experiences with greater ease. In addition, the person may realize the presence of an unchanging sense of equanimity, contentment, or eudemonic well-being amidst these changing stimuli. From a perspective of polyvagal theory, as the person learns to become aware of changing neural activations they are able to move between neural platforms to re-establish parasympathetic dominant states and they may be able to find a wider tolerance for safety amidst mobilizing or immobilizing activations. Once a foundation in regulation is found through practices that support sattva and social engagement neural platform, the person can move to the work of building resilience. The person learns to recognize when they shift from states of regulation, how to return to them, to be present with the changing activations without being overwhelmed, and to manage or alter the relationships and reactions to the gunas and neural platforms. Three types of resilience are described from this perspective of polyvagal theory and yoga.

Moving Between Neural Platforms and Gunas The practices of yoga can be utilized to create a foundation in sattva and the social engagement neural platform as well as to move between neural platforms or gunas. An example is rotating between practices that foster parasympathetic and sympathetic activations. Starting with parasympathetic and sattvic activation through practices such as longer exhales, restorative or comfortable postures, and medita402

 Yoga and Polyvagal Theory

tions with a focus on peace and ease would be followed by mobilizing practices that support raja guna. These activating practices could include active postures that increase muscle tension and breath rate, or breathing practices that are activating to the system. The person would cycle back and forth between these practices that create a relaxation response and those that require physiological activation, helping the person learn to move between neural platforms or gunas with greater ease. The person learns to recognize these different neural platforms and gunas and to have confidence in their ability to return to social engagement and sattva. This skill can be translated to life as the person notices when these different activations arise and can develop proficiency in regulating back to states of homeostasis or calm after times of activation.

Widening the Range of Tolerance for Neural Platforms and Gunas Proficiency in identifying and moving between neural platforms and gunas can progress to helping the person build a greater tolerance to stimuli. The person can learn to be present to different neural platforms or gunas while still perceiving safety or ease. Rather than becoming overwhelmed by these different activations, the individual fine-tunes their response to stimuli of the body, mind, and environment. An example would be in a yoga practice where the person is physiologically mobilized with an active posture while they maintain a breath pattern to stimulate a relative parasympathetic response and/ or meditative focus for comparative calm or ease. A broader range of safe mobilization can be found so that the person can be with a greater level of activating stimuli without becoming overwhelmed. As this is translated to life the person finds they can be present to a wider range of circumstances and life situations, to a range from safe mobilization to safe immobilization. From a yoga standpoint, they can perceive the quality of sattva—clarity and calm—amidst greater turbulence from rajas or dullness/obscuration from tamas. In life this translates to the person being able to be present to and tolerate a greater variety of stimuli and intensity of body and mind fluctuations and life circumstances while maintaining a degree of ease or calm.

Equanimity within Fluctuating Gunas and Neural Platforms As a progression of resilience the individual moves from being able to return to relaxed states after activation, to the ability to be with differing levels of activation while maintaining a measure of ease, to the recognition of all stimuli as a part of awareness with equanimity or eudemonic well-being. In this type of resilience, the individual recognizes everything that arises in the body, mind, and environment as awareness. Rather than moving back and forth between activations, or finding a platform of ease from which to be with greater variations in activations, there is the potential to experience all activations as awareness. A sense of welcoming arises as each stimuli of the body, mind, and environment are seen as pointers to and inherent facets of awareness. Deep equanimity, abiding joy, and eudemonic well-being are found within and amongst all gunas and neural platforms. The person comes to experience their pain, anxiety, and worry as an aspect of themselves that does not detract from this eudemonic well-being. Their depressive or anxious thoughts, medical conditions, or pain are understood as a current characteristic being experienced by the body or mind—but they do not misidentify or over identify with them. The person recognizes that these symptoms are an expression of the body, mind, or life circumstance, but that they do not define the totality of who they are nor their experience. In this type of

403

 Yoga and Polyvagal Theory

resilience the person recognizes neural platforms and gunas as awareness and experiences abiding joy, and eudemonic well-being simultaneously with these fluctuations of the body, mind, and environment.

CONCLUSION Yoga and polyvagal theory share a perspective whereby affecting underlying substrates can help build regulation and resilience to support eudemonic well-being for improved physical, mental, and social health. The person is able to consciously activate healthy and positive physiological, psychological, and behavioral states such as eudemonic well-being through the social engagement neural platform and sattva guna. This foundation of safety, calm, and relaxation supports a process of self-inquiry for selfrealization and self-actualization. Interoceptive and neuroceptive skills are developed as subconscious reactions of the body, mind, and behavior come to conscious attention. Healthy, accurate recognition of these activations supports discriminative wisdom for transformative change. The person is empowered to discover a life of greater authenticity and alignment with one’s values for a meaningful, purpose-filled, and flourishing life—they discover the agency to consciously cultivate relaxed physiological, positive psychological, and prosocial behavioral attributes supportive of health or well-being. From these positive and healthy states of the body, mind, and behavior resilience can then be cultivated. By working with changing neural platforms or guna states through yoga practice the person learns to both notice and navigate between them with greater ease. The person is then able to widen the window of tolerance to these shifting neural platforms and gunas to find relative balance within mobilized or immobilized states. Finally, there is a potential for the person to change the relationship with activations to find eudemonic well-being amidst the neural platforms and gunas. In other words, the person learns to experience these fluctuating stimuli without over-identifying with them. Yoga practices like movement, breathwork, ethics, and meditation can be used synchronistically for both regulation and resilience. The person builds self-regulatory skills through practices that facilitate the social engagement neural platform and sattva guna. Resilience is then fostered by finding the practices that activate the other neural platforms and gunas. This convergence of polyvagal theory and yoga enables the practitioner to understand how the practices are used together to influence body, mind, behavior through affecting underlying neural platforms or gunas. The practices of yoga can then be applied in a synergistic fashion in alignment with yoga’s philosophical context for optimal well-being. This is also important as yoga becomes integrated in both research and healthcare contexts so that it is understood as a distinct, unique profession for various patient populations.

ACKNOWLEDGMENT My understanding of the material presented in this chapter has been immeasurably deepened by work with respected colleagues including Matt Erb, Stephen Porges, Steffany Moonaz, and Laura Schmalzl. Thank you to Laurie Hyland Robertson for her work in editing this chapter.

404

 Yoga and Polyvagal Theory

REFERENCES Aristotle, Thomson, J. A. K., & Tredennick, H. (2004). The nicomachean ethics. Penguin Books. Azam, M. A., Katz, J., Mohabir, V., & Ritvo, P. (2016). Individuals with tension and migraine headaches exhibit increased heart rate variability during post-stress mindfulness meditation practice but a decrease during a post-stress control condition – A randomized, controlled experiment. International Journal of Psychophysiology, 110, 66–74. doi:10.1016/j.ijpsycho.2016.10.011 PMID:27769879 Barakat, A., Vogelzangs, N., Licht, C. M. M., Geenen, R., Macfarlane, G. J., de Geus, E., ... Dekker, J. (2012). Dysregulation of the ANS is associated with pain intensity, not with the presence of chronic widespread pain. Arthritis Care and Research, 64, 1209–1216. PMID:22422576 Berntson, G. G., Cacioppo, J. T., & Quigley, K. S. (1991). Autonomic determinism: The modes of autonomic control, the doctrine of autonomic space, and the laws of autonomic constraint. Psychological Review, 98(4), 459–487. doi:10.1037/0033-295X.98.4.459 PMID:1660159 Bhavanani, A. B. (2011). Are we practicing yoga therapy or yogopathy. Yoga Therapy Today, 26–28. Bhavanani, A. B., Sullivan, M. B., Taylor, M. J., & Wheeler, A. (2019). Shared foundations for practice: The language of yoga therapy. Yoga Therapy Today, 19, 44–47. Burley, M. (2012). Classical samkhya and yoga: an indian metaphysics of experience. Routledge. Ceunen, E., Vlaeyen, J. W. S., & Van Diest, I. (2016). On the origin of interoception. Frontiers in Psychology, 7, 743. doi:10.3389/fpsyg.2016.00743 PMID:27242642 Chu, I. H., Wu, W. L., Lin, I. M., Chang, Y. K., Lin, Y. J., & Yang, P. C. (2017). Effects of yoga on heart rate variability and depressive symptoms in women: A randomized controlled trial. Journal of Alternative and Complementary Medicine (New York, N.Y.), 23(4), 310–316. doi:10.1089/acm.2016.0135 PMID:28051319 Cole, S. W. (2013). Social regulation of human gene expression: Mechanisms and implications for public health. American Journal of Public Health, 103(S1), S84–S92. doi:10.2105/AJPH.2012.301183 PMID:23927506 Cole, S. W. (2014). Human social genomics. PLOS Genetics, 10(8), e1004601. doi:10.1371/journal. pgen.1004601 PMID:25166010 Cole, S. W., Levine, M. E., Arevalo, J. M. G., Ma, J., Weir, D. R., & Crimmins, E. M. (2015). Loneliness, eudaimonia, and the human conserved transcriptional response to adversity. Psychoneuroendocrinology, 62, 11–17. doi:10.1016/j.psyneuen.2015.07.001 PMID:26246388 Craig, A. D. (2015). How do you feel?: an interoceptive moment with your neurobiological self. Princeton University Press. doi:10.1515/9781400852727 Dale, L. P., Carroll, L. E., Galen, G. C., Schein, R., Bliss, A., Mattison, A. M., & Neace, W. P. (2011). Yoga practice may buffer the deleterious effects of abuse on women’s self-concept and dysfunctional coping. Journal of Aggression, Maltreatment & Trauma, 20(1), 90–102. doi:10.1080/10926771.2011.538005

405

 Yoga and Polyvagal Theory

Dezutter, J., Casalin, S., Wachholtz, A., Luyckx, K., Hekking, J., & Vandewiele, W. (2013). Meaning in life: An important factor for the psychological well-being of chronically ill patients? Rehabilitation Psychology, 58(4), 334–341. doi:10.1037/a0034393 PMID:24295525 Dezutter, J., Luyckx, K., & Wachholtz, A. (2015). Meaning in life in chronic pain patients over time: Associations with pain experience and psychological well-being. Journal of Behavioral Medicine, 38(2), 384–396. doi:10.100710865-014-9614-1 PMID:25537924 Farb, N., Daubenmier, J., Price, C. J., Gard, T., Kerr, C., Dunn, B. D., Klein, A. C., Paulus, M. P., & Mehling, W. E. (2015). Interoception, contemplative practice, and health. Frontiers in Psychology, 6, 763. doi:10.3389/fpsyg.2015.00763 PMID:26106345 Fredrickson, B. L., Grewen, K. M., Algoe, S. B., Firestine, A. M., Arevalo, J. M. G., Ma, J., & Cole, S. W. (2015). Psychological well-being and the human conserved transcriptional response to adversity. PLoS One, 10(3), e0121839. doi:10.1371/journal.pone.0121839 PMID:25811656 Fredrickson, B. L., Grewen, K. M., Coffey, K. A., Algoe, S. B., Firestine, A. M., Arevalo, J. M. G., … Cole, S. W. (2013). A functional genomic perspective on human well-being. Proceedings of the National Academy of Sciences, 110, 13684–13689. 10.1073/pnas.1305419110 Gard, T., Noggle, J. J., Park, C. L., Vago, D. R., & Wilson, A. (2014). Potential self-regulatory mechanisms of yoga for psychological health. Frontiers in Human Neuroscience, 8, 770. doi:10.3389/ fnhum.2014.00770 PMID:25368562 Haase, L., Stewart, J. L., Youssef, B., May, A. C., Isakovic, S., Simmons, A. N., Johnson, D. C., Potterat, E. G., & Paulus, M. P. (2016). When the brain does not adequately feel the body: Links between low resilience and interoception. Biological Psychology, 113, 37–45. doi:10.1016/j.biopsycho.2015.11.004 PMID:26607442 Huta, V., & Waterman, A. S. (2014). Eudemonia and its distinction from hedonia: Developing a classification and terminology for understanding conceptual and operational definitions. Journal of Happiness Studies, 15(6), 1425–1456. doi:10.100710902-013-9485-0 Ivtzan, I., & Papantoniou, A. (2014). Yoga meets positive psychology: Examining the integration of hedonic (gratitude) and eudemonic (meaning) wellbeing in relation to the extent of yoga practice. Journal of Bodywork and Movement Therapies, 18(2), 183–189. doi:10.1016/j.jbmt.2013.11.005 PMID:24725784 Keyes, C. L., & Simoes, E. J. (2012). To flourish or not: Positive mental health and all-cause mortality. American Journal of Public Health, 102(11), 2164–2172. doi:10.2105/AJPH.2012.300918 PMID:22994191 Koenig, J., Falvay, D., Clamor, A., Wagner, J., Jarczok, M. N., Ellis, R. J., ... Thayer, J. F. (2016). Pneumogastric (vagus) nerve activity indexed by heart rate variability in chronic pain patients compared to healthy controls: A systematic review and meta-analysis. Pain Physician, 19, E55–E78. PMID:26752494 Kolacz, J., & Porges, S. W. (2018). Chronic diffuse pain and functional gastrointestinal disorders after traumatic stress: Pathophysiology through a polyvagal perspective. Frontiers in Medicine, 5, 145. doi:10.3389/fmed.2018.00145 PMID:29904631

406

 Yoga and Polyvagal Theory

Mackenzie, M. J., Carlson, L. E., Paskevich, D. M., Ekkekakis, P., Wurz, A. J., Wytsma, K., Krenz, K. A., McAuley, E., & Culos-Reed, S. N. (2014). Associations between attention, affect and cardiac activity in a single yoga session for female cancer survivors: An enactive neurophenomenology-based approach. Consciousness and Cognition, 27, 129–146. doi:10.1016/j.concog.2014.04.005 PMID:24879038 Maletic, V., & Raison, C. L. (2009). Neurobiology of depression, fibromyalgia and neuropathic pain. Frontiers in Bioscience, 14(1), 5291–5338. doi:10.2741/3598 PMID:19482616 Mallinson, J., & Singleton, M. (2017). Roots of yoga. Penguin Classics. Meeus, M., Goubert, D., De Backer, F., Struyf, F., Hermans, L., Coppieters, I., De Wandele, I., Da Silva, H., & Calders, P. (2013). Heart rate variability in patients with fibromyalgia and patients with chronic fatigue syndrome: A systematic review. Seminars in Arthritis and Rheumatism, 43(2), 279–287. doi:10.1016/j.semarthrit.2013.03.004 PMID:23838093 Miller, R. (2012). The samkhya karika. Integrative Restoration Institute. Muehsam, D., Lutgendorf, S., Mills, P. J., Rickhi, B., Chevalier, G., Bat, N., Chopra, D., & Gurfein, B. (2017). The embodied mind: A review on functional genomic and neurological correlates of mind-body therapies. Neuroscience and Biobehavioral Reviews, 73, 165–181. doi:10.1016/j.neubiorev.2016.12.027 PMID:28017838 Park, G., & Thayer, J. F. (2014). From the heart to the mind: Cardiac vagal tone modulates top-down and bottom-up visual perception and attention to emotional stimuli. Frontiers in Psychology, 5, 278. doi:10.3389/fpsyg.2014.00278 PMID:24817853 Porges, S. W. (1993). The infant’s sixth sense: Awareness and regulation of bodily processes. Zero to Three: Bulletin of the National Center for Clinical Infant Programs, 14, 12–16. Porges, S. W. (2003). Social engagement and attachment. Annals of the New York Academy of Sciences, 1008(1), 31–47. doi:10.1196/annals.1301.004 PMID:14998870 Porges, S. W. (2004). Neuroception: A subconscious system for detecting threats and safety. Zero to Three, 24, 19–24. Porges, S. W. (2011). The polyvagal theory: Neurophysiological foundations of emotions, attachment, communication, and self-regulation. W. W. Norton. Porges, S. W. (2017). The Oxford handbook of compassion science. Oxford University Press. Porges, S. W., & Carter, C. S. (2017). Complementary and integrative treatments in psychiatric practice. American Psychiatric Association Publishing. Pressman, S. D., Jenkins, B. N., & Moskowitz, J. T. (2019). Positive affect and health: What do we know and where next should we go? Annual Review of Psychology, 70(1), 627–650. doi:10.1146/annurevpsych-010418-102955 PMID:30260746 Resnick, B., Galik, E., Dorsey, S., Scheve, A., & Gutkin, S. (2011). Reliability and validity testing of the physical resilience measure. The Gerontologist, 51(5), 643–652. doi:10.1093/geront/gnr016 PMID:21402647

407

 Yoga and Polyvagal Theory

Ryff, C. D., Singer, B. H., & Dienberg Love, G. (2004). Positive health: Connecting well-being with biology. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 359(1449), 1383–1394. doi:10.1098/rstb.2004.1521 PMID:15347530 Sarang, P., & Telles, S. (2006). Effects of two yoga based relaxation techniques on heart rate variability (HRV). International Journal of Stress Management, 13(4), 460–475. doi:10.1037/1072-5245.13.4.460 Schleicher, H., Alonso, C., Shirtcliff, E. A., Muller, D., Loevinger, B. L., & Coe, C. L. (2005). In the Face of Pain: The relationship between psychological well-being and disability in women with fibromyalgia. Psychotherapy and Psychosomatics, 74(4), 231–239. doi:10.1159/000085147 PMID:15947513 Schmalzl, L., Powers, C., & Henje Blom, E. (2015). Neurophysiological and neurocognitive mechanisms underlying the effects of yoga-based practices: Towards a comprehensive theoretical framework. Frontiers in Human Neuroscience, 9, 235. doi:10.3389/fnhum.2015.00235 PMID:26005409 Smith, J. D. (Ed.). (2009). Mahābhārata. Penguin. Staud, R. (2008). Heart rate variability as a biomarker of fibromyalgia syndrome. Future Rheumatology, 3(5), 475–483. doi:10.2217/17460816.3.5.475 PMID:19890437 Stoler-Miller, B. (1998). Yoga: discipline of freedom. Bantam Books. Streeter, C. C., Gerbarg, P. L., Saper, R. B., Ciraulo, D. A., & Brown, R. P. (2012). Effects of yoga on the ANS, gamma-aminobutyric-acid, and allostasis in epilepsy, depression, and post-traumatic stress disorder. Medical Hypotheses, 78, 571–579. doi:10.1016/j.mehy.2012.01.021 PMID:22365651 Strigo, I. A., & Craig, A. D. (2016). Interoception, homeostatic emotions and sympathovagal balance. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 371(1708), 20160010. doi:10.1098/rstb.2016.0010 PMID:28080968 Sullivan, M., & Robertson, L. (2020). Understanding Yoga Therapy: Applied philosophy and science for health and well-being. Taylor & Francis Group. doi:10.4324/9780429507243 Sullivan, M. B. (2019). Yoga and science in pain care: Treating the person in pain. Jessica Kingsley Publishers. Sullivan, M. B., Erb, M., Schmalzl, L., Moonaz, S., Noggle Taylor, J., & Porges, S. W. (2018). Yoga therapy and polyvagal theory: The convergence of traditional wisdom and contemporary neuroscience for self-regulation and resilience. Frontiers in Human Neuroscience, 12, 67. doi:10.3389/fnhum.2018.00067 PMID:29535617 Sullivan, M. B., Moonaz, S., Weber, K., Taylor, J. N., & Schmalzl, L. (2018). Toward an explanatory framework for yoga therapy informed by philosophical and ethical perspectives. Alternative Therapies in Health and Medicine, 24, 38–47. PMID:29135457 Taylor, A. G., Goehler, L. E., Galper, D. I., Innes, K. E., & Bourguignon, C. (2010). Top-down and bottom-up mechanisms in mind-body medicine: Development of an integrative framework for psychophysiological research. Explore (New York, N.Y.), 6(1), 29–41. doi:10.1016/j.explore.2009.10.004 PMID:20129310

408

 Yoga and Polyvagal Theory

Telles, S., Sharma, S. K., Gupta, R. K., Bhardwaj, A. K., & Balkrishna, A. (2016). Heart rate variability in chronic low back pain patients randomized to yoga or standard care. BMC Complementary and Alternative Medicine, 16(1), 279. doi:10.118612906-016-1271-1 PMID:27514611 Thayer, J. F., & Lane, R. D. (2000). A model of neurovisceral integration in emotion regulation and dysregulation. Journal of Affective Disorders, 61(3), 201–216. doi:10.1016/S0165-0327(00)00338-4 PMID:11163422 Tracy, L. M., Ioannou, L., Baker, K. S., Gibson, S. J., Georgiou-Karistianis, N., & Giummarra, M. J. (2016). Meta-analytic evidence for decreased heart rate variability in chronic pain implicating parasympathetic nervous system dysregulation. Pain, 157(1), 7–29. doi:10.1097/j.pain.0000000000000360 PMID:26431423 Tsuji, H., Venditti, F. J. Jr, Manders, E. S., Evans, J. C., Larson, M. G., Feldman, C. L., & Levy, D. (1994). Reduced heart rate variability and mortality risk in an elderly cohort. The Framingham Heart Study. Circulation, 90(2), 878–883. doi:10.1161/01.CIR.90.2.878 PMID:8044959 Tugade, M. M., & Fredrickson, B. L. (2004). Resilient Individuals use positive emotions to bounce back from negative emotional experiences. Journal of Personality and Social Psychology, 86(2), 320–333. doi:10.1037/0022-3514.86.2.320 PMID:14769087 Tugade, M. M., & Fredrickson, B. L. (2007). Regulation of Positive Emotions: Emotion regulation strategies that promote resilience. Journal of Happiness Studies, 8(3), 311–333. doi:10.100710902-006-9015-4 Tyagi, A., & Cohen, M. (2016). Yoga and heart rate variability: A comprehensive review of the literature. International Journal of Yoga, 9(2), 97. doi:10.4103/0973-6131.183712 PMID:27512317 Whitson, H. E., Duan-Porter, W., Schmader, K. E., Morey, M. C., Cohen, H. J., & Colón-Emeric, C. S. (2016). Physical resilience in older adults: Systematic review and development of an emerging construct. The Journals of Gerontology. Series A, Biological Sciences and Medical Sciences, 71(4), 489–495. doi:10.1093/gerona/glv202 PMID:26718984

409

410

Chapter 23

Traditional Frameworks of Well-Being and Modern Science Ganesh Mohan Svastha Yoga and Ayurveda, Singapore

ABSTRACT This chapter describes how to connect traditional frameworks from yoga to modern science and systems theory in a way that is evidence-based, clinically testable, inclusive, and extensible. This way of systematization enables leveraging the evidence for yoga from across the spectrum of well-being modalities. The key in the presentation is the choice of an approach that is equally valid for both ancient and modern systems. The authors begin by examining the importance of a skill-based approach and layout how wellbeing systems can be built by mapping the skills and qualities of the organism. They take a selection of domains or categories under which skills from traditional and modern sources can be analyzed: attention, transcendence, movement, breathing, devotion, cognition, and emotion. They conclude the chapter by presenting a larger map of well-being that includes more domains arranged from a clinical perspective.

INTRODUCTION The Indian subcontinent has several thousand years of history of well-being practices. Among these, yoga and ayurveda are two key systems. There is substantial insight embedded in their traditional frameworks. If that insight is to be maximally relevant in a modern well-being setting, we need to analyze those frameworks and connect them with modern science. This chapter presents a way of examining and reworking ancient frameworks of well-being from yoga, without losing their essential nature and contributions, with a view to making them more applicable in contemporary life in the background of current science.

DOI: 10.4018/978-1-7998-3254-6.ch023

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 Traditional Frameworks of Well-Being and Modern Science

BACKGROUND: CONTENT AND SCOPE OF FRAMEWORKS To analyze any framework, we need to begin with its purpose and scope. A well-known way of introducing a traditional India system thought, like Vedānta (Nikhilananda, 1931) or the Yoga Sūtra (Sastri & Sastri, 1952), or even Haṭha yoga (Svātmārāma & Aiyangar, 1972), is to define the subject, the purpose, the qualification of the practitioner, and the connections between these. In other words, we need to define what a framework consists of (the content) and in whom and where it is to be used (context and applicability). A structured approach is not restricted to modern thinking. The process of logical analysis is a foundation of ancient systems as well, including of yoga and ayurveda for well-being and self-transformation. In fact, if we are to connect ancient and modern systems of wellbeing, we need a common structured approach that can apply to both modern and ancient frameworks. In the next section, we are going to examine what that common structure can be.

STATES, SKILLS, SYSTEMS THEORY The foundation of yoga and ayurveda—arguably of all wellbeing frameworks –are properties (guṇa) and functions (karma). This derives from the Nyāya-Vaiśeṣika system that, along with Sāṅkhya, underpins ayurveda. Substrates (dravya) have properties (guṇa) and functions or actions (karma) (Sinha, 1923). Analyzing the concept underlying this presentation, we can see that properties describe the state of a system (Sharma & Dash, 1976). Functions or actions are what allow the state of the system to change (Sharma & Dash, 1976). In well-being frameworks, guṇa describes the current and potential state of well-being the individual may have. Karma describes the capacities or skills of individuals to change their state of well-being. All well-being frameworks select properties and functions according to their context or scope and explain or postulate connections between these chosen properties and functions. Depending on the scope of a framework, it may choose to include different qualities and functions. Some frameworks deliberately include a narrower subset of the mind-body continuum. Some are broader, including a large section of the mind and body. Smaller frameworks of narrower scope are nested within larger or broader frameworks. Some frameworks deal mainly with the body, and within that, some may deal with body feeling, while others may deal with body movement. Similarly, some frameworks deal mainly with thoughts, and others with the breath, and so on. We are not claiming the qualities and functions listed in the frameworks of yoga are entirely comprehensive. In fact, it is impossible for any system to list all the qualities and functions that would describe the body and mind. Consequently, it is not possible to build one framework that covers all of well-being and self-transformation. Any well-being system will cover only part of what the body and mind are capable of in function or the different states that the body and mind can exhibit as qualities. The value of traditional yoga is not purely in the lists that it proposes, but in the way in which the frameworks are structured and the foundation of a holistic approach that these frameworks reflect. There is a lot that we can learn from analyzing the way yoga structures the connections between different levels of mind and body. Apart from that, the qualities and functions that yoga uses are also well-selected. They are often the ones most useful in cultivating the capacity for self-regulation as well as other positive shifts of body and mind.

411

 Traditional Frameworks of Well-Being and Modern Science

An important phenomenon to note here is that of emergence. Properties and functions may arise from the organization of simpler units. As a simple example, components of a plane like the wings or the engine do not fly by themselves. Only when those components are assembled in a certain way does the function of flying emerge. In the human being, there are many levels of emergent function all the way from molecules to consciousness. It is important to consider which levels a framework targets. Generally, yoga frameworks target functions that emerge at higher levels of organization of the whole system, not at the level of molecules or individual cellular functions, or even body systems in isolation.

COMMON ORGANIZATIONS OF FRAMEWORKS Here are three common ways in which well-being frameworks are organized. These are all useful in practice. 1. External to internal, obvious to subtle: for example, from body to breath, from senses to mind. The eight limbs of yoga can be organized from external to internal. The first five limbs are more external as compared to the last three. And even within the first five limbs there is an organization from external to internal. The first limb (yama) begins with what we shall not do, behavioral restraints. The second limb is about what we can do to bring our mind from action to stillness (niyama). The third limb deals with the body (āsana). The fourth limb with the breath (prāṇāyāma). The fifth limb with the senses (pratyāhāra) and the last three limbs deal with the mind. 2. Hierarchical: top to bottom, bottom to top: common in modern science which builds a biological framework starting from atoms and extends it up to the mind. Yoga practices can also target loweror higher-level functions. This organization tends to emphasize causation and emergence. This present in yoga also, as the qualities and functions described under some limbs feed into emergent functions in other limbs. For example, providing sensory safety in the environment and locomotor relaxation and stability in āsana supports psychological stillness in the mind. Stillness does not emerge in the mind when there is danger in the environment (neuroception) or when there is restlessness in the body (fight and flight). 3. Landscape or spread: a presentation of the different areas to work on. Well-being frameworks may also be organized in the form of a spread or a landscape. The eight limbs of yoga can also be viewed like this. Each of the limbs has its own specific qualities and functions. The practices under each of the limbs, therefore, is not a substitute for practices that under other limbs. For example, that a person can move well does not mean that they will be able to practice meditation or mindfulness effectively. Another example: people may have skills to manage their lifestyle healthfully but that does not mean that the quality of the breathing is necessarily good. That is because each of these areas has its own specific qualities and functions. By following this principle of building a landscape

412

 Traditional Frameworks of Well-Being and Modern Science

it is possible for us to create a larger map of well-being than the eight limbs of yoga, by drawing upon related texts and disciplines.

BROADEST TRADITIONAL WELL-BEING FRAMEWORKS: GUṆAS, DOṢAS In yoga and ayurveda, the largest frameworks are the three guṇas (sattva, rajas, tamas or SRT) and doṣas (vāta, pitta, kapha or VPK). SRT apply more to the mind and VPK apply more to the body (Sharma & Dash, 1976). VPK describe the whole organism starting from the body and the qualities listed under them are more physical than psychological (Murthy, 1994). SRT describe the whole organism starting from consciousness and the mind and therefore, the qualities listed under them are more psychological than physical. Though both frameworks are comprehensive in their own way, SRT are more top-down while VPK are more bottom up. SRT are a foundation of the yoga psychology map. They contain characteristics mainly of energy, attention, emotion, thought, speech, and action (Easwaran, 2007). These can be related to neurological observations related to the polyvagal theory, cognitive function, and interconnection between parts of the brain, and more. The doṣas are presented primarily in ayurveda but also form a traditional foundation for the health benefits of practices of hatha yoga. Together with the guṇas, the doṣas form a map of body-mind functions that are diagnostic or assessment tools as well as a foundation for therapeutic interventions. The doṣas are detailed by describing their qualities and functions. The list of qualities and functions are extensible. For instance, these ten qualities (Murthy, 1994)—heavy, dull, cold, oily, smooth, dense, soft, stable, subtle, clear—with their ten opposites form a basic list which is then grouped under the three different doṣas. The major ayurvedic texts cite slightly different qualities but there is broad concurrence in the lists and their groupings into VPK. From here, we summarize some of the traditional frameworks and their connections to modern concepts, grouping them under domains or areas. We do not aim to cover all the domains in a map of well-being here, but to highlight key domains and use them to help elucidate the value of this systemsbased approach.

ATTENTIONAL FRAMEWORKS, MINDFULNESS, ABSORPTION The presentation of attention-based practice is detailed in traditional yoga. The Yoga Sūtra begins with a definition of yoga as stillness of the mind, and the commentary of Vyasa defines yoga as samādhi (absorption) (Bryant, 2009), which directly points to the central role of managing attention. The practices of absorption and mindfulness are presented specifically in the Yoga Sūtra as dhāraṇā, dhyāna, samādhi (Aranya, 2014), and smṛti (Aranya, 1980). With the spread of mindfulness practices in the modern world, this is an area of increasing research. Yet, there is substantial complexity in this research because of the numerous parameters that are part of varied meditation traditions: effort, content, emotion, spotlight etc. (Van Dam et al., 2017). Comparing the parameters used in modern mindfulness and meditation research with the traditional yoga presentations, we can identify several important states and skills. The presentation of attentional practices in the Yoga Sūtra separates the content of attention from the process. The process of absorp413

 Traditional Frameworks of Well-Being and Modern Science

tion and the skills therein fall under the heading of dhāraṇā, dhyāna, and samādhi. The possible targets of focus, or the contents of the absorptive experience, are described in both Chapter 1 and 3 of the text in numerous sūtras, leading to various classifications of meditation pathways. For example, sūtras 1.33 and 1.35-38 describe, in order, absorption in positive emotions, smell, light, desire less states, and beneficial dream experiences. Absorptive meditation states can also be classified according to the ground that they arise in—the habitual state of the mind that forms the foundation upon which the absorptive states are based (Aranya, 2014). An important correlate in modern positive psychology are flow states. Flow states should be considered rajas- or activity-dominant absorption in the view of the Yoga Sūtra as they typically involve states of greater challenge and corresponding skill leading to absorption (Csikszentmihalyi, 1991).

TRANSCENDENCE FRAMEWORKS, SĀṄKHYA, JÑĀNA YOGA The Sāṅkhya framework is a basis for both yoga and ayurveda. Sāṅkhya divides conscious experiences into layers from inside to outside—from the singular experience of existing, the “I” feeling, to the multitudinous experiences of the world through the senses. This framework provides a valuable lens into the pathway of what is termed as jñāna yoga traditionally—the yoga of insight gained through absorption or samādhi. Ultimately, the pathway of Sāṅkhya is to separate the sense of self from the experiences of the world: as the Sāṅkhya Kārikā says, “Not am, not mine, not I” (Sinha, 1915). Neurobiologically, this offers a map of controlled delinking of sense of self of from all other experiences. It is ego dissolution in a beneficial and stable way. In a therapeutic context, when a person presents with a history of trauma, we may see symptoms of dissociation (Lowenstein, 2018). That dissociation is uncontrolled and not beneficial. The pathway of Sāṅkhya in contrast, proposes a means to controlled dissolution of the bindings of the I-sense rather than an uncontrolled experience of dissociation. The map that Sāṅkhya provides, therefore, aims to layout the nature of conscious experiences, and how one may transform the I-sense through sustained absorption in progressively internal or subtle experiences. This pathway is presented in earlier works, such as the Upanishads, and is the backbone of many Eastern pathways. Sāṅkhya provides a more detailed structure than earlier systems and forms a base for later systems. Ultimately, this pathway takes the form of a progressive delinking of sense of self, from external to internal, such as, “I am not my body,” “I am not my senses,” “I am not my mind” (Aranya, 2014). When the yogi is able to complete this process, the individual is liberated from suffering, which is the stated goal of Sāṅkhya and of yoga.

MOVEMENT FRAMEWORKS, ĀSANA Haṭha yoga frameworks layout systems of body positions, movements, breath modulation, and inner body-based practices. Perhaps the most commonly known text on haṭha yoga is the Haṭha Yoga Pradīpikā. Perusing it, a reader will find only 15 āsanas described (Mohan & Mohan, 2017), unlike the large range of movements and positions that characterize modern yoga classes. Consequently, not all āsana practices that have been popularized in the 20th century and later have clear roots in traditional texts.

414

 Traditional Frameworks of Well-Being and Modern Science

However, some practices of haṭha yoga are not about exercise or fitness. In fact, this category of practices which include the mudrās and bandhas, and other inner- or subtle-body practices, are among the oldest that we can find in haṭha yoga texts, dating to the 11th century (Szántó, 2016). The Yoga Sūtra only speaks of āsanas briefly, but even in the short span of three sutras, it presents a few keys crucial to a beneficial mind-body connection. It notes that the body must find a balance in stability and comfort and that relaxing effort and connecting to the infinite is beneficial in achieving that balance (Aranya, 1980). But to find further details, we need to look to more recent figures in the field of yoga, such as the yogi Krishnamacharya in the 20th century who popularized many elements of what we see in āsana classes today: vinyāsa, krama, group fitness āsana classes and more (Mohan & Mohan, 2010). The āsana practices that are mainly about exercise and fitness are better analyzed more from a modern science perspective than from the lens of traditional frameworks. For example, strength training or stabilization exercises are detailed much better in modern science than they are in traditional texts. Conversely, modern exercise and rehabilitation systems are limited in their application of subtler body awareness and mind-body connection frameworks that yoga has explained centuries ago. What differentiates āsana from just exercise is the integration of breathing, movement, and awareness, and the intention that the āsana should lead to deeper self-transformation and control over the mind. The latter point is noted in the Haṭha Yoga Pradīpikā itself, that haṭha yoga should lead to rāja yoga (Mohan & Mohan, 2017) (meditation-based practices). The number of skills and states to be developed in āsana as both exercise and self-transformation is substantial. A list of those skills would include: awareness of the body, relaxation and release, ease and efficiency, mobilization and traction, stability and control, strength and endurance, range of movement and stretching, alignment and safe loading, movement exploration and creativity, balance and gait, speed and responsiveness, coordination and increasing complexity, positive emotion and supportive dialogue toward the body, lifestyle modification and perhaps more. The mind-body connection components of these skills are well-explored in traditional yoga teachings. An ideal approach to movement frameworks would combine the detailed knowledge of functional anatomy from modern science with the holistic mind-body frameworks of yoga, with a skill-based approach as the foundation. Numerous neurobiological pathways are involved in āsana practice, particularly the sensorimotor pathways and somatic maps in the nervous system at various levels, forming a fertile ground for researchbased elucidations.

BREATHING FRAMEWORKS, PRĀNĀYĀMA One key meeting point between haṭha yoga and rāja yoga is how physical practices can change one’s sense of self in the body. Another key meeting point is how the body may be a platform for attentional states on subtler inner experiences. Traditionally, the pathway to explore this connection has been prāṇāyāma and breathing. The Yoga Sūtra, when describing prāṇāyāma in sutra 2.50, uses the term “deśa.” This word literally means “place.” In this context, “place” refers to where the yogi experiences the prāṇāyāma—this is the key to the transformation that prāṇāyāma can bring.

415

 Traditional Frameworks of Well-Being and Modern Science

The central experience of prāṇāyāma should be an inner experience, starting from the feeling of the breath flowing into the body—to the chest, abdomen, and heart space—and progressing to deeper subtler body sensations. Further, that experience should be sattva-dominant: calm, clear, pleasant, and stable (Aranya, 1980). There are numerous descriptions of inner body experiences in the haṭha yoga texts including the Haṭha Yoga Pradīpikā and Yoga Yājñavalkya—under the broad categories of nāḍīs, prāṇa, cakras, kuṇḍalinī, marmasthāna, kandasthāna etc. Many of these practices appear esoteric from a modern lens. As a group, they are broadly aimed at shifting mind-body experience with the aim of changing sense of self and leading to deeper meditative states. In other words, these practices are aimed at ultimately leading the person to the rāja yoga or jñāna yoga pathway that commences with “I am not my body.” Consequently, they aim to bring the body to a state of experiential quiescence—partly through stopping the breath in hatha yoga—this is expressed in the Haṭha Yoga Pradīpikā as the iḍā (left) and piṅgalā (right) nāḍīs becoming lifeless (maraṇāvasthā) when the prāṇa enters the suṣumnā (central channel) (Mohan & Mohan, 2017). That is, when body sensation or awareness is limited to the central spinal pathways, the bilateral expansion of sensation to left and right diminishes and there is quiescence or temporary suspension of the sense of embodiment. This forms a foundation of the experience that, “I am not my body,” and therefore leads to the deeper practices of rāja yoga. A different but related thread in traditional prāṇāyāma descriptions is their health benefits through the lens of the three doṣas. The yoga texts describe the effects of prāṇāyāma techniques as being heating or cooling and relate these to the doṣas. In short, right-side emphasis and fast breathing are considered heating practices, while left-side emphasis and śītalī or śītkārī (inhalation through the lips) are considered cooling practices (Mohan & Mohan, 2017). This connection between breathing and the doṣas is a critical element in the therapeutic use of yoga for metabolic, autoimmune, hormonal, digestive, and immune system disorders. From a modern medical perspective, apart from the somatosensory pathways through the spinal cord, interoception of visceral sensations through the vagus nerve is vital for homeostasis and, thus, well-being. The vagus nerve is substantially afferent, informing the central nervous system about the state of the organs (Agostoni et al., 1957). Breathing, reflected through respiratory sinus arrhythmia, is a direct modulator of vagal function—both the input and output branches of the feedback loop. Prāṇāyāma, therefore, works on visceral homeostasis in multiple ways. It facilitates safe and stable access to interoception through paying attention to inner body sensations in a sattva-dominant state. This allows the information from the viscera to propagate to the central nervous system more effectively, thereby facilitating better regulation of the organ systems. Prāṇāyāma also influences vagal output by altering the parameters of breathing which in turn reflects in control of the heart. Breathing affects the nervous system in numerous ways (Bordoni et al., 2018), including movement of the brain mass, oscillation of neural networks (Zelano et al., 2016), circulatory changes, and motor coordination. Interestingly, the pressure of air flowing through the nostrils influences cerebral rhythms, even if there is no active attempt at paying attention to the breath (Piarulli et al., 2018). In summary, prāṇāyāma and breathing exert influences on health and well-being through a variety of mechanisms at many levels. Many of these connections are encapsulated in the traditional framework of connecting types of prāṇāyāma to the doṣas. There are many discoveries awaiting us in the science of breathing for well-being that could throw further light on how traditional breathing frameworks are insightful. 416

 Traditional Frameworks of Well-Being and Modern Science

DEVOTIONAL FRAMEWORKS, BHAKTI YOGA The cosmology or philosophical world view of yoga contains descriptions of Īśvara (the Divine) as well as origin of the world and the relationship between self and the world. Broadly, yoga pathways that emphasize devotion to and absorption in the Divine are termed bhakti yoga. The Bhagavad Gita contains detailed descriptions of the pathway of devotion to the Divine, also clearly emphasizing it toward the conclusion of the text (Easwaran, 2007). The Yoga Sūtra with its commentaries also presents the topic in depth in sūtras 1.23-29. It is essential to note how the Yoga Sūtra introduces the practice of absorption in Īśvaraas an option rather than an unavoidable necessity. Yoga Sūtra 1.23 which is the first reference to the Divine in the text explicitly includes the Sanskrit word “vā” which means “or”. In the conclusion of the Bhagavad Gita also, despite Krishna emphasizing devotion to the Divine, he explicitly points Arjuna toward thinking for himself and making his own choice by finding clarity rather than demanding faith without question (Easwaran, 2007). This basic point of providing options other than devotion is a foundation of traditional yoga that is crucial to its psychological applications. This is one of the major points that differentiates yoga from religions—that yoga suggests devotion to the Divine as a useful way to reach stillness of the mind, but it is not enforced in any way, nor is it the only pathway. This makes it possible for yoga pathways to accommodate the beliefs of different religions and faiths—so long as they increase the skills of being sattva-dominant—calm, clear, pleasant, stable etc. The presentation of Īśvara or the Divine in the Yoga Sūtra is detailed but careful. It stops short of recommending any specific form or religious belief. Instead, it posits Īśvara as the originator of the science and practice of yoga, the pathway to a still mind. And if each yogi can reach that state, Īśvara is what has always been in that state of transcendence or liberation from suffering. It is beyond our scope in this article to delve deeper into this topic, but we can conclude this section with an interesting observation. The role of Īśvara in yoga as the creator of the universe is more of a logical presentation rather than a belief system. The foundation of Sāṅkhya is to deconstruct the process of conscious experience. Therefore, the view of yoga tends toward the observer’s subjectivity being the foundation of the universe—in other words, the interaction of the consciousness with the mind creates the subjective universe. The experience of the external or objective universe that is shared across individuals can be considered to arise from the interaction of the consciousness of individuals with the substrate of the mind of Īśvara. That is, Īśvara does not need to create the universe per se, but to provide the possibility of it arising from the interaction with the consciousness individual. This presentation raises several interesting connections to modern thinking in quantum physics about the role of the observer in determining reality or the outcome of experiments.

COGNITIVE FRAMEWORKS, SVĀDHYĀYA, CBT The modern field of cognitive behavioral therapy (CBT) has substantial research behind it. It employs strategies to reduce maladaptive cognition about the world, self, and future (Beck, 1970). There is evidence to show that CBT is useful in a variety of psychological disorders, particularly in anxiety disorders (Hofmann et al., 2012).

417

 Traditional Frameworks of Well-Being and Modern Science

The idea of CBT is not new from the lens of yoga. The Yoga Sūtra puts forth frameworks for correcting cognition as a foundation for reducing suffering. In fact, it is the basic premise of the Yoga Sūtra that the foundation of suffering is misperceiving one’s unchanging self or consciousness (puruṣa) as the I-sense that changes with each experience (asmitā). Of course, changing this, from the yoga perspective, is not merely about fixing cognitive errors, but creating an experiential shift of which cognitive changes or reframing are a part. Consequently, we can find cognitive frameworks in multiple places in traditional yoga texts, particularly the Yoga Sūtra. Chapter 1 of the Yoga Sūtra (1.5-11) specifically presents a framework that deals with maladaptive cognition. Five classes of cognitive activities are detailed there: accurate, inaccurate, imagination or abstractions, memories, and the cognition during sleep (Aranya, 2014). This is a classification that is aimed at explaining how to bring these cognitions of the mind to stillness, which is the stated definition of yoga in the Yoga Sūtra. The first four categories in this framework can form a simple foundation for self-reflection, to look at one’s thoughts and behaviors and classify them as being accurate or inaccurate in the present, as being based in abstractions and imaginations, or in memories from the past. Most thinking is a combination of these categories and they can each contribute to suffering. The practice of svādhyāya, presented in chapter 2 of the Yoga Sūtra, is meditation with a mantra or studying texts that lead to stillness of the mind (mokṣa). This is a substantially a practice of bringing clarity to cognition, particularly using words and reflection (Aranya, 2014). In the same chapter, we also find another cognitive practice in the form of pratipakṣa-bhāvana. The yogi identifies cognition that is opposite to the positive direction that he is proceeding toward. Then, reframing them as cognitive distortions, he finds reasons to challenge their value, realizing that they are patterns he has already experienced and will only lead to further suffering (Aranya, 2014). This is very similar to some of the practices detailed in CBT centuries later, such as identifying cognitive distortions, challenging them, and changing them.

EMOTIONAL FRAMEWORKS, POSITIVE PSYCHOLOGY Cultivating positive emotional states is recommended in the Yoga Sūtra in chapter 1 and chapter 2 asana essential practice. Sūtra 1.33 lays out a straightforward framework. To make the mind calm and clear, it is recommended that yogi should practice love, compassion, joy, and equanimity, when faced with other happiness, unhappiness, immorality, and virtue in other living beings. This is a succinct presentation that is not easy to do, but valuable to mental and physical health. This is also a practice that is extremely well-known in Buddhism under the Metta Sutta (Dhamma talks, Mettā Sutta, AN 4:125), and consequently has been widely practiced over centuries and in modern times. The niyamas, second among the eight limbs of yoga, also present positive emotion practices in the form of contentment or gratitude (santoṣa) and acceptance (Īśvara-praṇidhāna). The yoga Yājñavalkya presents ten niyamas, expanding on the five presented in the Yoga Sūtra, offering more practices based on positive emotions (Mohan & Mohan, 2013). There is a large body of evidence accumulating in modern research on the broad role of positive emotions and behaviors for wellbeing (Fredrickson, 2001), as well as research on specific emotions such as altruism (Post, 2005) and gratitude (Armenta et al., 2017). This is one area where there is little difference between traditional and modern perspectives. The pathways of practicing positive emotions have not changed since the ancient times. 418

 Traditional Frameworks of Well-Being and Modern Science

A HOLISTIC MAP OF WELL-BEING Expanding upon what we have presented so far, we may build a larger map of areas or domains to work on for well-being, from a practical or clinical angle. We can use the different frameworks presented in yoga, combining them with modern science, to address skill requirements in these areas. And if we wish, we could add more areas too. 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16.

Make positive changes easy. Sustain stable and comfortable attention. Manage your inner dialogue. Cultivate emotional resilience. Move functionally throughout life. Breathe well. Balance your lifestyle and environment. Explore food as nourishment. Befriend the stress response: safety, relaxation, empowerment. Nurture beneficial relationships and boundaries. Cultivate wisdom in speech. Use touch beneficially. Find meaning in engaging and disengaging with the world. Consider your beliefs, religion, and spirituality. Explore sound, rhythm, music. Discover other arts: painting, drama, writing and more.

CONCLUSION In this chapter, we have examined how yoga and modern science can be connected. We used a systemsbased approach, starting with states and skills that are useful to consider in a discussion on well-being. This is how ancient texts from India have themselves viewed health and well-being, so this is an excellent way to begin connecting them with modern approaches. Any living organism as complicated a human being has innumerable states and skills at many levels of emergence, many of which may be crucial to monitor and shift in the quest for well-being. We need to have a way to decide which skills we work with in yoga and how to categorize them and understand their connections. Three common ways to do that are from external to internal, as a hierarchy from top to bottom and bottom to top, and as a landscape or a spread. All these and more classifications can be useful. The broadest frameworks used in traditional yoga are the three guṇas and three doṣas which mainly use a landscape or spread classification. We have taken up some of the common domains or areas that ancient yoga texts use to classify skills and states under. Specifically: attention, transcendence, movement, breathing, devotion, cognition, and emotion. Clinically, we may not choose to divide the skills we examine under these exact headings, but the same principles hold. So, at the conclusion of the chapter, we present a useful list of domains that we can look at when applying yoga for well-being, considering a combination of both ancient and modern frameworks.

419

 Traditional Frameworks of Well-Being and Modern Science

REFERENCES Agostoni, E., Chinnock, J. E., De Daly, M. B., & Murray, J. G. (1957). Functional and histological studies of the vagus nerve and its branches to the heart, lungs and abdominal viscera in the cat. The Journal of Physiology, 135(1), 182–205. doi:10.1113/jphysiol.1957.sp005703 PMID:13398974 Aranya, H. (1980). Pātañjala-Yogadarśana. Motilal Banarsidass Publishers. Aranya, H. (2014). Yoga philosophy of patañjali with bhāsvati. University of Calcutta. Armenta, C. N., Fritz, M. M., & Lyubomirsky, S. (2017). Functions of positive emotions: Gratitude as a motivator of self-improvement and positive change. Emotion Review, 9(3), 183–190. doi:10.1177/1754073916669596 Beck, A. T. (1970). Cognitive therapy: Nature and relation to behavior therapy. Behavior Therapy, 1(2), 184–200. doi:10.1016/S0005-7894(70)80030-2 PMID:27993332 Bernardi, L., Sleight, P., Bandinelli, G., Cencetti, S., Fattorini, L., Wdowczyc-Szulc, J., & Lagi, A. (2001). Effect of rosary prayer and yoga mantras on autonomic cardiovascular rhythms: Comparative study. British Medical Journal, 323(7327), 1446–1449. doi:10.1136/bmj.323.7327.1446 PMID:11751348 Bordoni, B., Purgol, S., Bizzarri, A., Modica, M., & Morabito, B. (2018). The influence of breathing on the central nervous system. Cureus, 10, e2724. doi:10.7759/cureus.2724 PMID:30083485 Bryant, E. F. (2009). The yoga sūtras of patañjali. North Point Press. Burns, D. D. (1989). The feeling good handbook. Harper Collins Publishers. Csikszentmihalyi, M. (1991). Flow. Harper Collins Publishers. Dhammatalks. (n.d.). Goodwill MettāSutta (AN 4:125). https://www.dhammatalks.org/suttas/AN/ AN4_125.html Easwaran, E. (2007). The Bhagavad Gita (Classics of Indian Spirituality). Nilgiri Press. Fredrickson, B. L. (2001). The role of positive emotions in positive psychology: The broaden-and-build theory of positive emotions. The American Psychologist, 56(3), 218–226. doi:10.1037/0003-066X.56.3.218 PMID:11315248 Hofmann, S. G., Asnaani, A., Vonk, I. J., Sawyer, A. T., & Fang, A. (2012). The efficacy of cognitive behavioral therapy: A review of meta-analyses. Cognitive Therapy and Research, 36(5), 427–440. doi:10.100710608-012-9476-1 PMID:23459093 Loewenstein, R. J. (2018). Dissociation debates: Everything you know is wrong. Dialogues in Clinical Neuroscience, 20(3), 229–242. doi:10.31887/DCNS.2018.20.3/rloewenstein PMID:30581293 Mohan, A. G., & Mohan, G. (2010). Krishnamacharya. Shambhala Publications. Mohan, A. G., & Mohan, G. (2013). Yoga Yājñavalkya. Svastha Yoga. Mohan, A. G., & Mohan, G. (2017). Haṭha Yoga Pradīpikā. Svastha Yoga.

420

 Traditional Frameworks of Well-Being and Modern Science

Murthy, S. (1994). Vāgbhata’s Aṣṭāṅga Hṛdayam. Krishnadas Academy. Nikhilananda, S. (1931). Vedantasara of Sadananda. Advaita Ashrama. Piarulli, A., Zaccaro, A., Laurino, M., Menicucci, D., De Vito, A., Bruschini, L., Berrettini, S., Bergamasco, M., Laureys, S., & Gemignani, A. (2018). Ultra-slow mechanical stimulation of olfactory epithelium modulates consciousness by slowing cerebral rhythms in humans. Scientific Reports, 8(1), 1–17. doi:10.103841598-018-24924-9 PMID:29700421 Post, S. G. (2005). Altruism, happiness, and health: It’s good to be good. International Journal of Behavioral Medicine, 12(2), 66–77. doi:10.120715327558ijbm1202_4 PMID:15901215 Sastri, P. S. R., & Sastri, S. R. K. (1952). Pātañjala-yogasūtra-bhāṣya-vivaraṇam of śaṅkara bhagavatpāda. Government Oriental Manuscripts Library. Sharma, R. K., & Dash, V. B. (1976). Agniveśa’s caraka saṃhitā. Chowkhamba Sanskrit Series Office. Sinha, N. (1915). The samkhya philosophy. The Panini Office. Sinha, N. (1923). The Vaiśeṣika sūtras of kāṇāda. The Panini Office. Svātmārāma, S., & Aiyangar, M. S. (1972). The Hathayogapradipika of Svatmarama: with the commentary Jyotsna of Brahmananda. Adyar Library and Research Centre. Szántó, P.-D. (2016). A Brief Introduction to the Amrtasiddhi. SOAS University of London. Van Dam, N. T., van Vugt, M. K., Vago, D. R., Schmalzl, L., Saron, C. D., Olendzki, A., Meissner, T., Lazar, S. W., Kerr, C. E., Gorchov, J., Fox, K. C. R., Field, B. A., Britton, W. B., BrefczynskiLewis, J. A., & Meyer, D. E. (2018). Mind the hype: A critical evaluation and prescriptive agenda for research on mindfulness and meditation. Perspectives on Psychological Science, 13(1), 36–61. doi:10.1177/1745691617709589 PMID:29016274 Zelano, C., Jiang, H., Zhou, G., Arora, N., Schuele, S., Rosenow, J., & Gottfried, J. A. (2016). Nasal respiration entrains human limbic oscillations and modulates cognitive function. The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 36(49), 12448–12467. doi:10.1523/ JNEUROSCI.2586-16.2016 PMID:27927961

421

422

Chapter 24

The Science and Art of Contemporary Breathwork and Yoga Jim Morningstar Transformations Incorporated, USA Jessica Dibb Inspiration Consciousness School, USA

ABSTRACT This chapter provides evidence for the benefit of bringing yoga and its breathing practices into a working relationship with contemporary breathwork techniques for the betterment of both in certain circumstances. The history and developmental overlap of both are explored as well as their applications in the field of physical heath, psychotherapy, and spiritual growth. The benefits of both faster than normal and slower than normal breathing techniques are reviewed for the treatment of anxiety, depression, and addictions as well as the neurological correlates for breathwork and trauma recovery. A pilot study utilizing the combination of the two disciplines is reported.

INTRODUCTION Since the 1970s there has been an explosion in the blending of ancient breathing techniques for health and spiritual awareness with contemporary growth and therapeutic practices. The result has been the evolution of powerful and incisive healing and consciousness changing modalities. This has given birth to the field of contemporary breathwork which has promoted significant advances in medical, psychological and spiritual domains. Breathwork is one of the newly emerging tools of complementary medicine which has the daunting yet highly engaging task of blending the old and the new for the purpose of creating something better than either in isolation. It presents a new paradigm of whole brain learning which is at the intersection of science and art. Our life is a dance of the two cerebral hemispheres: right/left, intuitive/logical. This DOI: 10.4018/978-1-7998-3254-6.ch024

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 The Science and Art of Contemporary Breathwork and Yoga

understanding is not just a poetic viewpoint, but necessary to a full comprehension of how we operate. We can no longer live in world of Newtonian physics, black and white thinking. We learn with our whole body, organically not just with the cerebral cortex. Breathwork bridges both brain hemispheres and accesses functions of paleo, meso and neo cortex, thus reaches a more inclusive level functioning. A 3,000 year old tradition, yoga, is now regarded in the Western world as a holistic approach to health, and is classified by the National Institutes of Health in the United States as a form of Complementary and Alternative Medicine (Woodyard, 2011). Yoga is a form of mind-body fitness that involves a combination of muscular activity and an internally directed mindful focus on awareness of the self, the breath, and energy (Collins, 1998). The time honored tradition of yoga and the developing field of contemporary breathwork have developed independently with both similarities and differences in the use of the breath as evidenced in the various schools of both disciplines, e.g., nostril versus mouth breathing, use of breath retention, slower than normal vs. faster than normal breathing, etc. This chapter focuses on the overlaps and potential complementary benefits of their use together, rather than on the theories which appear to differ or purport to be mutually exclusive. How yoga and contemporary advances in breathwork are complementing and enhancing each other is the objective of this chapter.

Slower Than Normal Breathing Breath awareness becomes the foundation for practices which I call maintenance breathing skills that have been so influential for humans through millennia in bringing increased peace of mind and healing to the body. This is exemplified by pranayama in the yoga tradition and translated into contemporary forms such as mindfulness and coherent breathing. Coherent Breathing, a form of maintenance breathing, for example, entails slowing the breath to 5 to 6 breaths per minute for 4 minutes or more. The physiological effects are documented by Heart Math research (Heart Math Institute, 2001) for regulating the parasympathetic nervous system and calming the mind. It is the practical and immediately experienced effects of breath awareness and modulation that can open minds to the benefits of more regularly implemented practices. This can also prepare people for more major lifestyle renovation, e.g., opening awareness to their dietary and exercise habits.

Natural Breathing Ultimately our goal in breath awareness is not to be consciously controlling our breathing every moment of the day. Breath modulation techniques are used to counter the holding patterns which impede easy flowing natural breathing and to practice maintaining that state. Breath awareness simply makes us more conscious of when the breath is not flowing easy and naturally so we can pay attention and address the signs of distress. The goal of breath awareness and breath modulation is to experience the joy of natural breathing and being alive.

Faster Than Normal Breathing Therapeutic breathwork (Morningstar, 2017) is form of breathing that adds to the repertoire of breath awareness and maintenance breathing, by including and teaching the skills of sympathetic nervous system regulation. As a form of healing, increasing the breathing cycle to 30-70 breaths per minute in 423

 The Science and Art of Contemporary Breathwork and Yoga

a safe supportive setting for the purpose of harmonizing one’s emotional system allows past holding patterns and emotional blockage to be relieved more directly than by years of talking about symptoms (talk therapies). Talking often does not reach the unconscious memory which has stored emotional reactivity through trauma which gets activated by environmental stimuli and which produces feeling and behavioral responses that do not fit current circumstances. In extreme cases they are experienced as flashbacks which put an individual into another time and space that feels out of their control. The elegance of this kind of breathing technique is that it helps to reprogram traumatic responses by going directly to the limbic system where they are stored (somatic approach), rather than by trying to enter by the doorway of the prefrontal cortex (verbal approach) whose pathways to the experience have been blocked by the body’s emergency control systems. When people learn to use this type of breathing technique consciously, they regain the confidence in their own abilities to handle settings and circumstances which were previously paralyzing to them. This could be as simple as having dinner with their family or as dynamic as performing before a large audience.

HISTORY OF YOGA AND BREATHWORK TRADITIONS Traditional and contemporary uses of breathwork in the healing arts include breathing techniques such as yoga and Taoist techniques which have been used in the East for centuries, not just for health maintenance but for diagnosis and cure. Off-shoots have been used in Western medicine for childbirth (the Lamaze technique, https://www.lamaze.org), pain control (hypnosis, autosuggestion), asthma relief, panic attacks, anxiety and a variety of mental health disorders. In India, the tradition of consciously using modulated breath in various breathing techniques is as long-established as it is in China (Minett, 2004). Here, too, it is used in healing and healthcare, as well as to provide a spiritual path. The method for maintaining and improving a person’s physical and spiritual wellbeing is known as yoga, while the medical aspect is utilized in ayurvedic medicine. The first evidence of yoga practices has been linked with the pre-Aryan Harappa culture which existed in the Indus Valley between 3000-1500 B.C. In this tradition, prana, which is similar to Qi in China, stimulates growth in all living organisms at the level of the smallest cells. Everything that is living requires prana in order to exist. Prana exists everywhere, and can be absorbed by the body via the value of the food we eat, via our skin, but above all through breathing. Prana is not the same as oxygen; rather it is prana that gives oxygen it’s life-giving quality and is cultivated by a variety of breathing practices or pranayama which employ both slower and faster than normal breathing.

DEVELOPMENT OF CONTEMPORARY BREATHWORK Applications of contemporary breathwork are found in remedial medicine, physical and occupational therapy. Breathwork has been effectively employed to release deep tissue trauma and re-access the shutdown of entire systems under autonomic control that have not been touched by other medical intervention.

424

 The Science and Art of Contemporary Breathwork and Yoga

APPLICATIONS OF CURRENT RESEARCH TO YOGA AND BREATHWORK Research is now proving what yogis have known all along: breathwork can deliver powerful mind and body benefits. Mindfulness and mindfulness meditation use the practices of breath awareness, and sometimes breath modulation, to enhance focus and to regulate and harmonize physical, emotional and mental functioning. Below are some of the findings of this research (Levine, 2015, 2018).

Breath for Detoxification and Energy When we inhale, oxygen is absorbed into the bloodstream and is transported to every cell of the body. When we exhale, CO2 (the body’s waste) is removed from the body. As Patton Sarley writes in his article for Kripalu: “Cellular respiration, the source of all our physical energy and expression in the world, is dependent on a constant flow of oxygen being delivered to each cell and carbon dioxide being taken away” (O’Brien, 2015).

The Brain on Meditation Mindfulness Practice greatly reduces stress and improves our ability to focus and, it turns out, that can translate into higher test scores and a better memory. For a study published in the journal, Psychological Science, a group of researchers at the University of California, Santa Barbara, enrolled 48 undergraduates in an experiment (Mrazek et al., 2013). Half of the subjects practiced mindfulness four days a week for two weeks, while the other half took part in a healthy nutrition program. Researchers found that in the group that meditated, the students’ working memory actually improved, while their habitual mind wandering started to decrease. They also had higher GRE verbal scores, which jumped on average from 460 to 520 in only 14 days. “We had already found that mind wandering underlies performance on a variety of tests, including working memory capacity and intelligence,” UC Santa Barbara graduate student Michael D. Mrazek told The New York Times. In short, the meditation group was better able to focus on ideas and remember facts without getting distracted as easily (Mrazek et al., 2013).

The Healing Power of the Breath In their book, Simple Techniques to Reduce Stress and Anxiety, Enhance Concentration, and Balance Your Emotions, Drs. Richard P. Brown and Patricia L. Gerbarg (2012) provide a drug-free alternative that works through a range of simple breathing techniques drawn from yoga, Buddhist meditation, the Chinese practice of qigong, Orthodox Christian monks, and other sources. These methods have been scientifically shown to be effective in alleviating specific stress and mood challenges such as anxiety, insomnia, post-traumatic stress disorder, and many others. The authors explain how breathing practices activate communication pathways between the mind and the body, positively impacting the brain and calming the stress response.

425

 The Science and Art of Contemporary Breathwork and Yoga

Breath for pH Balance and Anti-inflammation Discussions about pH and inflammation have been prominent in health and wellness lately. Deep breathing reduces inflammation and has alkalizing effects on the body, balancing the pH. Conversely, the effects of shallow breathing are a buildup of CO2, creating acidic conditions. Deep breathing may be the most healing action you can take.

Breath for Stress and Chronic Conditions Yoga is recognized as a form of mind-body medicine that integrates an individual’s physical, mental and spiritual components to improve aspects of health, particularly stress related illnesses (O’Brien, 2015). It has been shown that many people move through life in a near-constant state of “fight/flight” activation with the stress hormone cortisol pumping through their bodies. Chronic stress is a root cause of countless chronic conditions. Short, shallow breathing is appropriate for fight/flight, as the body prepares to survive a conflict. But it’s not appropriate for everyday life. Deepening the breath, and elongating the exhale, can shift the sympathetic nervous systems and activate the parasympathetic nervous systems. This simple act recalibrates the stress response and addresses or contributes to the prevention of many serious chronic conditions. An abundance of oxygen may be the easiest and biggest boost to one’s health and wellness regime. Natural breathing patterns can be restored by bringing attention to the breath throughout the day and including pranayama in yoga practice.

FURTHER EVIDENCE-BASED BENEFITS OF YOGIC BREATHING PRACTICES (LEVINE, 2018) Happiness and Emotional Stability Manipulating the breath can alter the emotional state, accounting for as much as a 40 percent variance in feelings of anger, fear, joy, and sadness, according to findings in the journal, Cognition & Emotion. The breathing instructions used to evoke joy in the study were “Breathe and exhale slowly and deeply through the nose.”

Weight Loss Yogic breathing practices increase levels of leptin, a hormone produced by fat tissue that signals the brain to inhibit hunger (Telles et al., 2010).

Better Exercise Stamina A cardiologist at the University of Pavia, Italy, compared a group of mountaineers who practiced slow breathing for one hour a day for two years before attempting to climb Mount Everest to a group who did not (Bernardi et al., 2006). The breathing group reached the summit without needing the supplemental

426

 The Science and Art of Contemporary Breathwork and Yoga

oxygen the other group did, and their blood and exhalation samples showed they were using 70 percent of the surface area of their lungs, an amount that maximizes the O2 taken in.

Longer Life Just one session of relaxing practices like meditation, yoga, and chanting influenced the expression of genes in both short-term and long-term practitioners, according to a Harvard study. Blood samples taken before and after the breathing practices indicated a post-practice increase in genetic material involved in improving metabolism and a suppression of genetic pathways linked with inflammation. Since chronic inflammation has also been associated with such deadly diseases as Alzheimer’s, depression, cancer, and heart disease, it’s probably fair to say that better breathing may not only change your life but may also save it.

Breathwork and Yogic Practices in Psychotherapy Psychological applications of breathwork include the healing of emotional traumas not reached by cognitive therapies, chronic tension not relieved but only masked by medications, negative and limiting beliefs, and habitual behaviors held in place by negative conditioning. Yoga and regulated breathing are being used as therapeutic tools in therapeutic practice as reported in the American Psychological Monitor on Psychology (Novotney, 2009): There is a growing body of research documenting yoga’s psychological benefits. Several recent studies suggest that yoga may help strengthen social attachments, reduce stress and relieve anxiety, depression and insomnia. Researchers are also starting to claim some success in using yoga and yoga-based treatments to help active-duty military and veterans with post-traumatic stress disorder. Sat Bir Khalsa, Ph.D., is a neuroscientist and professor of medicine at Harvard Medical School at Brigham and Women’s Hospital in Boston who studies yoga’s effects on depression and insomnia. Khalsa, who has practiced yoga for more than 35 years, says several studies in his 2004 comprehensive review of yoga’s use as a therapeutic intervention, published in the Indian Journal of Physiology and Pharmacology (Vol. 48, No. 3), show that yoga targets unmanaged stress, a main component of chronic disorders such as anxiety, depression, obesity, diabetes and insomnia. It does this, he says, by reducing the stress response, which includes the activity of the sympathetic nervous system and the levels of the stress hormone cortisol. The practice enhances resilience and improves mind-body awareness, which can help people adjust their behaviors based on the feelings they’re experiencing in their bodies, according to Khalsa. In a 2007 study in the Journal of Alternative and Complementary Medicine (Vol. 13, No. 4), researchers at Boston University School of Medicine and McLean Hospital used magnetic resonance imaging to compare levels of the neurotransmitter gamma-aminobutyric acid (GABA) before and after two types of activities: an hour of yoga and an hour of reading a book. The yoga group showed a 27 percent increase in GABA levels, which evidence suggests may counteract anxiety and other psychiatric disorders. GABA levels of the reading group remained unchanged.

427

 The Science and Art of Contemporary Breathwork and Yoga

And in what is becoming one of the most widely applied yoga-based trauma treatments, clinical psychologist Richard Miller, PhD, has developed a nine-week, twice-weekly integrative restoration program based on the ancient practice of yoga Nidra. In 2006, the Department of Defense began testing iRest with active-duty soldiers returning from Iraq and Afghanistan who were experiencing PTSD. At the end of the program, participants reported a reduction in insomnia, depression, anxiety and fear, improved interpersonal relations and an increased sense of control over their lives. Since then, iRest classes have been established at VA facilities in Miami, Chicago and Washington, D.C. Miller has also helped develop similar programs for veterans, homeless people and those with chemical dependencies and chronic pain.

Breathwork and Yogic Meditation Meditation Reduces Anxiety — Including Social Anxiety A lot of people start meditating for its benefits in stress reduction, and there’s lots of good evidence to support this rationale. There’s a whole newer sub-genre of meditation, mentioned earlier, called Mindfulness-Based Stress Reduction (MBSR), developed by Jon Kabat-Zinn at the University of Massachusetts’ Center for Mindfulness (now available around the world), that aims to reduce a person’s stress level, physically and mentally. Studies have shown its benefits in reducing anxiety, even years after the initial 8-week course (Miller et al., 1995). Research has also shown that mindfulness meditation, in contrast to attending to the breath only, can reduce anxiety – and that these changes seem to be mediated through the brain regions associated with those self-referential (“me-centered”) thoughts (Zeidan et al., 2014). Mindfulness meditation has also been shown to help people with social anxiety disorder: a Stanford University team found that MBSR brought about changes in brain regions involved in attention, as well as relief from symptoms of social anxiety (Goldin et al., 2013).

Short Meditation Breaks Can Help Children in School For developing brains, meditation has as much as or perhaps even more promise than it has for adults. There’s been increasing interest from educators and researchers in bringing meditation and yoga to school kids, who are dealing with the usual stressors inside school, and often times additional stress and trauma outside school. Some schools have started implementing meditation into their daily schedules, and with good effect: One district in San Francisco started a twice daily meditation program in some of its highrisk schools – and saw suspensions decrease, and GPAs and attendance increase (McFadden et al., 2014). Studies have confirmed the cognitive and emotional benefits of meditation for schoolchildren (Mrazek et al., 2013), but more work will probably need to be done before it gains more widespread acceptance.

Trauma Recovery Therapeutic Breathwork and Healing Trauma Therapeutic Breathwork™ adds a highly significant element not utilized in any other forms of trauma treatment, which is the activating breath. In Therapeutic Breathwork, clients are coached to engage in another way of breathing that mildly activates the limbic system (sympathetic nervous system) through increasing the breath volume in a safe setting. This allows formerly implicit-only or partially integrated 428

 The Science and Art of Contemporary Breathwork and Yoga

stimuli to resurface and be more fully integrated with safe outcomes. This is facilitated through the assistance of the “observer” breath worker and the “participant” client into an integrated story—often reinforced with the use of mental suggestion (affirmations or “inner child” visualizations that signal safety to the body). This reprograms the chaotic emotional responses and negative story the mind reached (or was told) to explain the disjointed reactions to life experiences. Since traumatic overwhelming stimuli can happen at any developmental stage in varying degrees, the negative stories are prime to be created in broad themes reflective of the developmental challenges of that stage. 1. 2. 3. 4. 5. 6.

Prenatal/after birth: “I’m not safe” (fragmentation) Year 1: “I am not enough or don’t have enough” (abandonment/ deprivation) Year 2: “I’m not in control” (overpowering or seductive manipulation/ abuse) Years 3 and 4: “I can’t express my feelings” (emotional suppression/ guilt and shame) Years 4 and 5: “I can’t express my gender” (gender judgment/ confusion/ alienation) Years 5 and 6: “I can’t have fulfilling love” (intimacy block/ betrayal)

Of course, traumas around these themes can happen or be reinforced at any period of a person’s life. Healing involves relinking our emotions, sensations, and pictures with more resourceful outcomes than being frozen and using self-talk that is continuously negative (e.g., “I am not lovable”). Therapeutic Breathwork precipitously engages the sympathetic nervous system to help reprogram what has been frozen in the fight-or-flight mechanism of the vagal system. As such, fears arise and are breathed through consciously and are integrated. Energy thwarted by trauma and unavailable for more creative endeavors than being on continuous “red alert” or hyper-vigilant, is thereby recovered. It does not anesthetize fears that might be appropriate for the person’s wellbeing (e.g., wearing a seatbelt). But it helps put the original fears into a workable context in order to be understood and allow the energy of the fear to be integrated into more conscious use. Pathways that had been cut off to the prefrontal cortex are now activated in resourceful ways (social engagement system or mammalian level of the vagus system), and clients “make sense” to themselves and can make productive choices based on present needs rather than operating in the trance state of past trauma (Morningstar, 2017).

Depression, Anxiety and Addictions Breathwork in the Healing of Anxiety, Depression and Addictions Anxiety is characterized by a diffuse fear that somatically is measured by a generalized constriction in the musculature, especially around the breathing system. The fear generated by not being able to breathe is some of the most intense and immediate fears wired in the human organism. So, a generalized tension around the breathing mechanism signalizing danger via the brainstem is not directly suppressible by higher cortical functioning. In other words, we can’t just talk ourselves out of it. There has to be physiological signals to the midbrain that the danger has passed in order for the system to relax and recover. This is regulated by the polyvagal nerve which controls and responds directly to breathing rate and amplitude. Regulating one’s breathing also changes the chemical composition of the blood and sends inhibitory messages to the brain regarding the fear response. In order to maintain this inhibitory state and for the 429

 The Science and Art of Contemporary Breathwork and Yoga

system to reboot the polyvagal nerve must be regulated. This has two phases. First is the breakthrough of the sympathetic emergency reaction which can be accomplished with adaptive faster than normal breathing and sympathetic system regulation. The second phase requires on-going maintenance, which can be accomplished through periods of slower than normal breathing. This often takes several cycles of coaching in order for the individual to learn self-regulation, dependent on the severity of the original holding pattern or trauma (Ehrmann, 2011). The anxious person’s breathing is marked by shorter, shallower breathing. Restriction could come from holding in the abdominal area and a concentration of breath in the upper chest with clavicles moving upwards on the inhale, or the opposite of frozenness in the chest and/or reverse breathing, i.e., belly going in on the inhale and out on the exhale. Their experience of diffuse fear is not just in their head. Even if it started with a psychological fear, the body’s resultant constriction is to avoid a real or imagined harm. This puts the organism in a vigilant, ready for fight or flight, posture that restricts breathing and reinforces the experience of imminent danger that can become habitual and self-perpetuating. There are a variety of these holding patterns that develop in early life and predispose individuals to an anxiety prone existence. A more thorough examination of six major body themes and breathing styles can be reviewed in Break Through with Breathwork (Morningstar, 2017). Underlying anxiety is often a precipitating factor in turning to substances, legal or illegal, for relief. It has long been known that pain is intensified by the anxiety that accompanies a threatening situation. Fear causes the body to tense itself, a spontaneous reaction to the need to fight or run away. So, what we have is a two-stage process: (1) fear giving rise to tension, and (2) tension increasing pain. Using the breath to limit tension will help to quiet both the fear and the pain. Depression presents another facet of breathing dysregulation and is characterized by overly suppressed breathing for an extended period leading to an altered brain chemistry, which keeps the system in a different form of emergency shutdown (Ehrmann, 2011). In this case a suppressive holding pattern in the breathing mechanism slows down the metabolism and other bodily functions to preserve resources during periods of deprivation or being trapped. The experience of “an elephant sitting on one’s chest” is common, as well as the psychological sense of hopelessness and helplessness. The depressed person’s breathing is labored, minimal and can appear to be a struggle to maintain survival. In the animal kingdom these states have been induced by putting animals in conditions of painful stimulation from which they could not escape. Their breathing and energetic functions become depressed until the situation is alleviated, at which time the organism can recover and eventually return to normal functioning. Humans can psychologically create trapped conditions in which they see no escape and give the same emergency instructions to their nervous system and bodily functions. Breathwork can be an important adjunct in the recovery process to jumpstart the energy system which then must be maintained by providing an altered psychological and emotional framework. The adroit use of faster than normal breathing, psychological reframing and eventually slower than normal maintenance can reduce the recovery time significantly and avert the dependency on pharmaceutical intervention to bring the system back to healthy self-regulation and enhanced growth. Other research has demonstrated the positive effect of breath modulation on anxiety and depression (Brown et al., 2005a, 2005b; Brown & Gerbarg, 2012). Addiction, in part, is an attempt on an individual’s part to maintain a state of feeling good when chronically stressed by using a stimulus that becomes associated with that “feeling good” state and be430

 The Science and Art of Contemporary Breathwork and Yoga

coming overly attached to it. This is controlled by a reward center in the brain which produces dopamine when stimulated. Once the associative connection has been made between the addictive substance and this center, repeated stimulation is reinforced independent of higher cortical mechanisms of control or suppression. Animals in the laboratory, when give the opportunity to voluntarily stimulate this appetitive center by pressing a bar connected to electrodes in their brain, will continue this activity above any other appetitive need like food and water till death (Nettle, 2006). This center is that strong in its drive for satisfaction. Humans, once the strong association has been made to that center, have a hard time “willing” themselves to break the connection and to parse out more balanced and healthy forms of stimulation that integrate the needs of the entire organism and address the original stress which they had not resolved with a more balanced approach. In a sense the period of addiction highlighted this unanswered challenge and exacerbated it to the point, sometimes life or death, where it had to be addressed and other resources had to be brought to bear in reaching healthier solutions. Breathing patterns of people with strong addictions have been noted to be similar to those with anxiety, but often more strongly entrenched. When a person’s addictive substance has been ingested, the holding pattern temporarily is released, and they feel appeased. The pleasure of letting go of the institutionalized tension they carry is in itself part of the addition cycle. Breathwork has been shown to provide that breakthrough experience which helps disconnect the addictive substance as the only source of stress reduction and helps in the process of rebuilding a healthier balanced energetic approach to life. Breathwork is equipped to do more than shock the system out of the addictive cycle, but to give self-regulatory tools on the physical, emotional, mental and spiritual levels - all of which must be integrated into a new life paradigm for the recovery to be a renewal and a growth to a more integrated level of functioning and fulfillment. When any of these aspects are not integrated into the recovery, relapse will continue until it is. There have been research studies on the use of regular conscious connected breathing programs (faster than normal breathing) in alcohol recovery with over 50 percent sustained sobriety (Rajski, 2002).

Quality of Life in Seniors In a study, one hundred thirty-five subjects, generally healthy men and women aged 65-85 years, were recruited . Participants were randomized to 6 months of Hatha yoga class, walking exercise class, or wait-list control. Subjects assigned to classes also were asked to practice at home. The yoga intervention produced improvements in physical measures (e.g., timed 1-legged standing, forward flexibility) as well as a number of quality-of-life measures related to sense of well-being and energy and fatigue when compared to exercise and wait-list control groups (Oken et al., 2006).

Spiritual Practices Employing Breathwork The diaphragm has been called the “spiritual muscle” in certain ancient practices. Breath control has been a central practice of most spiritual/mystical traditions for thousands of years (Minett, 2004). The breath is known as the “rainbow bridge” from the physical to the spiritual realms. Simple attention to a steady flow of breath can put one in an altered state, making one more sensitive to who one is beyond the physical. A prominent researcher in the area of spiritual traditions, Ken Wilber (2006) claims that meditation (facilitated by slow regulated breathing) is the only practice which has scientifically documented results on its positive influence on spiritual growth. The Russian mystic G. I. Gurdjieff is known to have said that without mastering breathing, nothing can be mastered. 431

 The Science and Art of Contemporary Breathwork and Yoga

As simplistic as this may sound, how long can you go just paying attention to your inhale and exhale? Most people cannot go for more than a few seconds without thoughts or sensations distracting them. If followed to its origins, the breath leads to our creative source, the spirit which initiates the “breath of life” in one’s body. In many languages spirit and breath are linked by the same words. e.g. nephesh in Hebrew, atman in Sanskrit, psyche in Greek. A very common phenomenon in therapeutic breathwork, after physical, emotional and mental holdings have been released, is to have openings to one’s spiritual realm or higher self that transport the breather to transcendent states (Grof, 2000). That breath is a powerful agent of healing and growth is well established in cultures around the world. That each of us has the ability to access this power “right under our nose” is being discovered and documented in many volumes (Morningstar, 1994).

BREATHING IN TRADITIONAL YOGA PRACTICE Max Strom, yoga teacher and author of A Life Worth Breathing (2012), calls it a classic Cinderella story: Pranayama is often overlooked while the beautiful sister, asana, is the guest of honor in yoga studios. But give breathing a chance, and you’ll realize it’s the true queen, Strom says. He lists five transformative breathing techniques that are foundational to yoga practice: 1. Basic Breath Awareness Begin by noticing where you already are with your breath, says Bo Forbes, PsyD, clinical psychologist and integrative yoga therapist. Do you know when and why your breath is shallow, or what makes it speed up? “This is really valuable information in creating stress resilience,” she says. Plus, just becoming aware of your breath tends to slow it down (Forbes, 2011). 2. Ujjayi Pranayama (Victorious Breath or Ocean Breath) This classic pranayama practice, known for its soft, soothing sound similar to breaking ocean waves, can further enhance the relaxation response of slow breathing, says Patricia Gerbarg, MD, assistant clinical professor of psychiatry at New York Medical College and co-author of The Healing Power of the Breath (Brown & Gerbarg, 2012). Her theory is that the vibrations in the larynx stimulate sensory receptors that signal the vagus nerve to induce a calming effect. 3. Nadi Shodhana Pranayama (Alternate-Nostril Breathing) This practice of alternating between the right and left nostrils as you inhale and exhale “unblocks and purifies the nadis, which in yogic belief are energy passages that carry life force and cosmic energy through the body,” Cole says. While there is no clear scientific evidence to support these effects, one pilot study found that within seven days of practicing this technique, overactive nervous systems were essentially rebalanced and a study of 90 people with high blood pressure found Nadi shodhana lowered blood pressure and improved mental focus (Levine, 2015). 4. Kumbhaka Pranayama (Breath Retention) 432

 The Science and Art of Contemporary Breathwork and Yoga

If you inhale fully and then wait 10 seconds, you will be able to inhale a bit more, Strom says. Why? Holding your breath increases pressure inside the lungs and give them time to fully expand, increasing their capacity. As a result, the blood that then travels to the heart, brain, and muscles will be more oxygenated. 5. Kapalabhati Pranayama (Breath of Fire or Skull-Shining Breath) This rapid breathing technique is energizing and activates the sympathetic nervous system. In a study using EEG electrodes to measure brain activity, researchers found that kapalabhati pranayama increased the speed of decision-making in a test requiring focus.

The Use of Breath in Traditional Asana Practice While priorities may differ between styles and teachers, when to inhale and exhale during asana is a fairly standardized practice element. Here, Cole (5) offers three simple guidelines for pairing breath with types of poses:

When Bending Forward, Exhale When you exhale, the lungs empty, making the torso more compact, so there is less physical mass between your upper and lower body as they move toward each other. The heart rate also slows on the exhalation, making it less activating than an inhalation and inducing a relaxation response. Since forward bends are typically quieting postures, this breathing rule enhances the energetic effects of the pose and the depth of the fold.

When Lifting or Opening the Chest, Inhale In a heart-opening backbend, for instance, you increase the space in your chest cavity, giving the lungs, rib cage, and diaphragm more room to fill with air. And heart rate speeds up on an inhalation, increasing alertness and pumping more blood to muscles. Plus, “Deep inhalation requires muscular effort that contributes to its activating effect,” Cole says. Poses that lift and open the chest are often the practice’s energizing components, so synchronizing them with inhalations takes optimum advantage of the breath’s effects on the body.

When Twisting, Exhale In twists, the inhalation accompanies the preparation phase of the pose (lengthening the spine, etc.), and the exhalation is paired with the twisting action. Posturally, that’s because as your lungs empty there’s more physical space available for your rib cage to rotate further. But twists are also touted for their detoxifying effects, and the exhalation is the breath’s cleansing mechanism for expelling CO2.

433

 The Science and Art of Contemporary Breathwork and Yoga

RESEARCH ON HOW YOGA AND CONTEMPORARY BREATHWORK HAVE SUCCESSFULLY BEEN EMPLOYED TOGETHER Recent exploratory studies first reported here have investigated the use of contemporary slower and faster than normal breathing techniques with traditional yoga asanas. The theory is that trained practitioners in both disciplines could effectively facilitate directed breathing techniques while a student was holding a pose to support the focus of their intent and energy movement, especially in the areas of emotional release. In a pilot study in the United States two protocols were used to document the effects of slower and faster than normal breathing techniques during a yoga class and the use of asanas during group breathwork employing slower and faster than normal breathing. Practitioners certified as both yoga teachers and breathworkers were used and their observations recorded along with the subjects’ self-reports. The full protocols of two pilot studies are reported in the Appendix. In both protocols, the practitioner who conducted the session had over 6 years of experience as a yoga instructor as well as 3 years as a breathworker. The hypotheses were that: 1. Slower and faster than normal breathing techniques with asanas would help the subject in their ability to regulate their emotional state. 2. Slower and faster than normal breathing techniques would help the subject in their ability to release holding patterns and/or trauma in their body. 3. Slower and faster than normal breathing techniques with asanas would not negatively affect subject’s flexibility or ability to hold a pose.

Protocol I: Yoga Session with Breathing Exercises Protocol I focused on utilizing both slower-than-normal and faster-than-normal breathing in a yoga session using various yoga asanas (postures). The sessions went for about 75 minutes which included: 1 to 3 asanas and at least one slower-than-normal form of breathing along and faster-than-normal breathing with each posture. Protocol I practitioner, Sara Bassindale, reported: “With both the individual and group sessions I personally felt that using more “restorative” poses felt more in alignment with using the different breathing techniques. I should also say, that is typically more my “yoga teaching style”, having more mindful, supportive poses. I chose a couple that had the face, body facing the earth, then a couple heart/ hip opening with front of body facing up specifically to see how people felt/ reacted to this contracted/ inward vs. the open/ outward. Some feedback was more challenging with the inward and more freeing/ connected with the outward. Everyone was very open to this experiment, and those who weren’t familiar with the conscious connected breath had interest in working with it more.” The participants who included comments on their feedback forms noted: H.M.: I felt the energy going through me more at some points of this session more than a normal breathing session. E.S.: The breathing practices go well with a yin-style yoga; the effect of breathwork on the emotional state is profound in letting go and accepting; that in turn allows the physical to soften. Thank you. A.B.: Supine butterfly more open – expansive; child’s – more contracting. 434

 The Science and Art of Contemporary Breathwork and Yoga

L.R.: During and throughout the practice my ability to drop deeper and deeper into the slower-thannormal breathing was quite apparent. As for trauma and/or holding patterns being released, I feel it won’t be revealed to me immediately and am anxious how the rest of the day/week will be affected by this practice. A.Y.: I thought this practice would be a bit scary (especially the faster breathing). But it was very calming. A.Z.: I found the slow breathing quite easy, but I was surprised at how difficult the fast breathing was and the amount of concentration it took. Every participant indicated that they would recommend combining combinations of breathing practices and yoga asanas for the purpose of healing and growth except for one who responded with a “?”.

Protocol II: Breathwork Sessions Incorporating Yoga Asanas Protocol II focused on utilizing various breathing techniques with various yoga asanas (postures). In this session the practitioner, Sara Egert, employed 4 asanas and 2 forms of breathing with each posture: 1. 5 minutes of each pose: mountain, cobra, child’s pose, corpse with a slower-than-normal (inhale through nose; hold for 2-3 seconds; exhale through mouth with pursed lips). 2. 10 minutes of faster-than-normal breathwork in each pose: mountain, cobra, child’s pose and ending in corpse pose for about 20 minutes. The participants who included comments on their feedback forms noted: M.G.: I loved this session! H.H.: I felt the energy going through me more at some points of this session more than a normal breathing session. M.M.: I absolutely loved this workshop & found incredible release from this, energetically & physically. I would definitely take this again. Would be nice to have journal time after. L.M.: The combination was incredible. I would absolutely recommend it to anyone and everyone. T.P.: I enjoyed the session very much. K.S.: I realized I have done breathwork before during pelvic floor physical therapy. As a social worker, I have done similar breathwork in sessions, but adding the yoga is great. I enjoyed this. I look forward to going to another similar session. Every participant indicated that they would recommend combining combinations of breathing practices and yoga asanas for the purpose of healing and growth. At the conclusion of both Protocol I and Protocol II sessions, the participants rated the following aspects on a scale of 1 to 7 from “Greatly Reduced” to “Greatly Increased.” The practitioners also rated their observations of the participants on this scale: (1) My flexibility was… (2) My ability to hold a posture was… (3) My ability to regulate my emotional state was…

435

 The Science and Art of Contemporary Breathwork and Yoga

(4) My ability to use faster- and slower-than breathing to release holding patterns and/or trauma in my body was… The graphs below show the ratings of subjects and practitioners from the two protocols. Preliminary indications are suggestive of a favorable outcome. 1) Subjects self-report and practitioner’s observations showed an average increase in their ability to regulate their emotional state. 2) Subjects self-report and practitioner’s observations showed an average increase in their ability to release holding patterns and/or trauma in their body. 3) Subjects self-report and practitioner’s observations did not show an average decrease in subjects flexibility or ability to hold a pose. These are simply beginning attempts to pair two powerful tools in the world of healing and growth: yoga and contemporary breathwork.

FUTURE RESEARCH DIRECTIONS Further controlled research with greater numbers will be needed to be more conclusive concerning the suggestive results. Nonetheless the positive reports of both practitioners and the preponderance of subjects are encouraging of further investigation.

CONCLUSION Breath awareness and modulation have been an integral part of the practice of yoga for thousands of years and continues to be one of the pillars of modern yoga teaching and practice. The more recent revival of interest in breathwork in and of itself as a healing and growth agent in health related disciplines, has opened the doors to both the application of contemporary breathwork research to the practice of yoga, but also has added impetus for the use of yoga as an adjunct to evolving areas of medicine, psychotherapy and spiritual disciplines. This synchronous blending of research and practice holds much promise for the science and applications of yoga and contemporary breathwork together.

436

 The Science and Art of Contemporary Breathwork and Yoga

Table 1. Volunteer and practitioner feedback on Protocol I and Protocol II Protocol I - Yoga session exploring breathwork techniques with asanas N=7 Volunteers’ Feedback

Greatly Reduced

Moderately Reduced

Somewhat Reduced

My flexibility was

1

My ability to hold a posture was

1

Unaffected

1

Somewhat Increased

Moderately Increased

Greatly Increased

weighted average

median

6

5.57

5

4.86

5

3

2

My ability to regulate my emotional state was

2

4

1

5.86

6

My ability to use faster-than and slower-than breathing in order to release holding patterns and/or trauma in my body was

5

1

1

5.43

5

Somewhat Increased

Moderately Increased

Greatly Increased

weighted average

median

1

6

5.86

5

3

1

4.43

5

4

2

1

5.57

5

4

1

1

5.29

5

Unaffected

Somewhat Increased

Moderately Increased

Greatly Increased

weighted average

median

1

5

5

Protocol I - Yoga session exploring breathwork techniques with asanas N=7 Practitioner Feedback

Greatly Reduced

Moderately Reduced

Somewhat Reduced

Unaffected

2

1

My flexibility was My ability to hold a posture was My ability to regulate my emotional state was My ability to use faster-than and slower-than breathing in order to release holding patterns and/or trauma in my body was

1

Protocol II - Breathwork session using yoga asanas before and/or during session N=11 Volunteers’ Feedback

Greatly Reduced

Moderately Reduced

Somewhat Reduced

My flexibility was My ability to hold a posture was

1

My ability to regulate my emotional state was

2

My ability to use faster-than and slower-than breathing in order to release holding patterns and/or trauma in my body was

5.36

5.5

6

4

6.09

5.5

2

1

4

2

5.00

6

1

4

3

3

5.73

6

Somewhat Increased

Moderately Increased

Greatly Increased

weighted average

median

5

5

1

5.09

5

Protocol II - Breathwork session using yoga asanas before and/or during session N=11 Practitioner Feedback

Greatly Reduced

Moderately Reduced

Somewhat Reduced

Unaffected

My flexibility was

11

My ability to hold a posture was

10

My ability to regulate my emotional state was

1

5

4.91

5

My ability to use faster-than and slower-than breathing in order to release holding patterns and/or trauma in my body was

1

5

4.91

5

ADDENDUM: Optimizing Well-Being through Self-Regulating Breathwork By Jessica Dibb It has been a wondrous life-long adventure for me to learn about, study, and practice so many varied breathing techniques. For over 30 years I have done yogic breathing practices such as ujjayi, bhastrika,

437

 The Science and Art of Contemporary Breathwork and Yoga

alternate nose breathing, kapalabhati, bumblebee breath, and many more. I have experienced coherent breathing, therapeutic breathwork, holotropic breathwork, transformational breathwork, rebirthing, vipassana, qigong breathing, t’ai chi breathing, Buteyko, tumo breath, tonglen practice, and more. Training people in Integrative Breathwork for over 25 years has been a privilege and taught me a lot about what people need for well-being and self-actualization. With a deep respect for each breath practice and opening to whatever the stated benefits of the practice were, I have also carefully observed whatever other benefits I experience. From simple relaxation techniques to deep soul opening modalities, I feel the beauty and understand the efficacy of different breathing rates, mouth and nose variations, concentration practices, parasympathetic/sympathetic coherence practices, cortisol regulation practices, trauma reduction and resolution practices, practices that cultivate desirable human qualities such as compassion, human potential practices, and spiritual practices. It’s been an honor to do over 10,000 individual and group breathwork sessions with people. With all the science that I’ve learned about breathing and breathwork, and all the practical experience and therapeutic efficacy that I’ve witnessed, I have come to the conclusion that the future of breathwork and breathwork training is to develop practitioners who can help support breathers in finding their own optimal rate, depth, and focus of each breath for well-being in any moment. For instance, while there is significant research and experiential evidence that five to six breaths per minute creates optimal emotional and neural regulation (Elliot & Edmonson, 2008), people who have done enough Breathwork to already produce significant physiological, emotional, and cognitive benefits, as well as experience awakened non-ordinary states of consciousness, may begin to find that as they are practicing slow, deep breathing, they are able to breathe slower and slower - even down to one or two breaths per minute or less. This should never be forced, but may arise organically as the body is adapting and adjusting to transformation physiologically, emotionally, cognitively and spiritually. Especially towards the end of a breathing session, someone’s breathing may slow down significantly, and they may describe feelings of wholeness, integration, and well-being, as well as bliss (from upcoming book, Breathwork in Psychotherapy, Jessica Dibb). As humans, we want to have a formula, and formulas are important for many understandings and advancements in human life and evolution. Physiologically the more autonomic functions such as the formation of red blood cells and gastric motility tend to follow a set pattern. Yet, no two individuals are alike. Obvious differences are things such as height, weight, gender, muscularity, and genetics. Variability and differentiation in embedded cortical patterns is created from each person’s bio-individuality (Davidson, 2014) and also unique experiences of attachment styles in childhood, internalized psychological object relations, birth order, adverse childhood experiences, trauma from birth through the present, quality of childhood nutrition, environmental factors during formative years such as exposure to chemical toxins and heavy metal, the doshas of Vata, Pitta, or Kapha, temperament, neurotransmitter levels (which seem to have some correlation to temperament) (Bond, 2001), and even the way each person has condensed memories to create a personal narrative. Epigenetic factors that affect breathing and optimal well-being and self-actualization would, therefore, be highly individualized and not universally measurable. Researchers have noted 80 percent of brain cells are glia, free agents, not predetermined or committed as part of our DNA (Jabr, 2012). Unlike neurons, glial cells are not committed to neural pathways, so they are more affected by bio-individuality and epigenetics. While there are obvious and specific research-validated benefits from specific breathing rates found in techniques such as coherent breathing, kapalabhati breathing, and bhastrika breathing, I have observed a more efficacious possibility for individual health and self-actualization. If we can work with opening 438

 The Science and Art of Contemporary Breathwork and Yoga

up the physiological constrictions and patterns that inhibit breathing fully in a person, encourage deep diaphragmatic deep full-bodied breathing, and support the meeting of life force while they are consciously breathing, each breath will find a unique rate that attends to physiological balance, autonomic sympathetic and parasympathetic balance, trauma resolution, neural regulation, emotional regulation, and tapping into individual and collective potential and spiritual experience. If we continue to midwife this self-regulating breath rate over the course of a breathwork or breathing practice session, much as we would midwife a birth, the person’s inner intelligence will choose what they need to experience in each moment in physical healing, emotional healing, cognitive awareness, and spiritual experience. Whenever we are focusing on something, it will spike the blood pressure slightly, even if only a little. Since it takes focus for a person to affect a certain breath rate it will raise the blood pressure a bit, even if the practice itself is about parasympathetic activation and overall lowers blood pressure. Whether the breather is instructed in relaxing breaths such as alternate nostril breathing, or bumble bee breath; activating breathing such as kapalabhati, bhastrika, or holotropic breathwork; or regulated breathing such as ujjayi or coherent breathing that balances activation and relaxation, we can say that the fact that there is attention on the rate may be slightly dissociative from the person’s optimal and authentic bioindividual physical coherence, emotional regulation, and spiritual awakening. So, I am excited about the possibility of breathwork becoming a truly bio-individual experience. The necessary training for such breathworkers of the future would be like post-graduate level training. The Therapist/Breathworker would need to: • • • • • • • • • •

learn many types of breathwork understand the science of breathing and breathwork understand the physiological, emotional, cognitive, and spiritual impacts of breathwork be skilled at facilitating focusing be skilled in facilitating the development of desirable qualities of being, such as compassion, empowerment, inner peace, joy, loving kindness be knowledgeable and skilled at facilitating human emotions such as rage, fear and grief have some medical knowledge about basic challenges such as lung, heart and neurological conditions have competent knowledge in emotional issues involving significant brain chemistry alterations such as ADHD, anxiety, schizophrenia, bipolar, etc. understand attachment styles and object relations dynamics have deep training in accompanying people with inner soul work and spiritual states

They would also need to have done deep work with themselves in all these areas. Thus, they would be like a skilled midwife of well-being, human potential, and spiritual awakening. They would be proficient in technique, yet know that individual attunement and artistry is the most powerful method.

Case Study 1 using Self-regulating Integrative Breathwork Diagnosis: Suspected Prolonged Psychotic episode and/or Schneiderian first-rank symptoms - possibly induced by low levels of vitamin B12 Jonathan (a pseudonym) was one of the most infamous drug addicts in the city of one million people where he lived. He had a reputation amongst other hardcore addicts for being able to swallow whole 439

 The Science and Art of Contemporary Breathwork and Yoga

handfuls of amphetamines and go on functioning, or drink several bottles of alcohol and still walk around or go skiing - doing slaloms no less! The intensity of his first integrative Breathwork session was so impactful to him, it superseded his drug experiences. I journeyed with Jonathan over the next decade as he got and stayed mostly sober with three relapses, all three of which lasted a few days to a couple of weeks. The second to the last one he almost died from internal bleeding. However, it was the last one that created an experience for him that set the stage for a powerful psychological transformation he would have in the future. During the third relapse he was so intoxicated after a long period of abstinence that he entered a full-blown drug and alcohol induced psychosis of which the doctors, and even I, were afraid he might not completely recover. I have seen peoples’ brains be permanently altered/damaged by that one step too far, that one time too many. We were deeply concerned. Someone with alcohol induced psychosis can revert to schizophrenia, especially men (Niemi-Pynttäri et al., 2013). He was admitted to the hospital, and they allowed me to stay with him for 10 hours while we kept corralling his wild ramblings back to sound psychodynamic communication, body awareness (as much as we could through his dissociation), and deep conscious breaths. Deeper breaths have the advantage of detoxing a person faster because of the increased oxygenation. This also means that the DTs can be accelerated, this is why grounding in the body with good therapeutic communication is essential, as well as medical support with IV hydration and supportive medications. I was pretty sure that detoxing him as quickly as possible, with Breathwork, would give him the best chance of recovering all of his brain coherence. It was harrowing, but he got there. Of course, I can’t say for sure that he would not have come out of it intact anyway, without the conscious breathing. What I can definitely say is that there was significant medical and intuitive concern that this time he had gone too far to avoid some kind of brain/ personality permanent harm. He recovered though, and quickly. Yet that was only the precursor to an even more amazing unfolding of the power of the mind-body connection with Jonathan. About 3 years later when he had been consistently sober, and in fact was a sponsor for at least 20 people in AA, Jonathan began to exhibit alarming emotional symptoms reflective of psychosis. He started muttering to himself in all kinds of situations, including in public, and appeared to be having conversations with angels or entities, listening to them for advice, and following their dictates. At 1 a.m. in the morning one of these “guides” told him to swallow a small plastic toy so he would be okay, which he did, necessitating a visit to the ER. At this time his wife told him that she thought she wanted to divorce him and take their young son to live with her. Jonathan was so distressed that he begged her to attend a Breathwork course with him called Awakening that I facilitate. His impulse control was quite poor by this point. He would start talking to her in the middle of class quite loudly. Sometimes he got up and followed me around the room as I stood in the center teaching. I had to ask assistants to lovingly walk him to the side of the room and work with him. During the first three classes of Awakening, we did standardized breathing practices, such as ujjayi and coherent breathing, however, we had not yet done any integrative breathing or self-regulating breathwork. He was barely sleeping, hardly eating, and delusional. His Psychiatrist was unclear what to do. He was not an imminent harm to himself or others, and he did not want to be hospitalized. We were all convinced that he was using drugs and drinking again. His symptoms were pretty much the same as whenever he relapsed, and his addiction progressed. This time he kept insisting that he was not using, yet we were so sure he was. He surrendered to three tests for alcohol and drugs, and at least two of those were when he couldn’t possibly prepare. We confronted him in a doctor’s office, and he had to give urine and get blood drawn right there. This occurred over a 5-day period and every single test came back negative for drugs or alcohol, blood and urine both. I had known and loved Jonathan for almost 15 years, and I was 440

 The Science and Art of Contemporary Breathwork and Yoga

deeply concerned. His behavior was bizarre and erratic, and his sentences were often incoherent. He also was exhausted beyond description and said it felt like he had no energy. For the whole 2 to 3 weeks he could barely stand up straight. So, I brought Jonathan in for an individual integrative Breathwork session. He talked about some things in his life, including his wife possibly leaving, and never feeling that he could get his mother’s approval. I said “Jonathan, have you ever felt truly unconditionally loved?” He immediately said “Yes, I have felt loved like that.” and he seemed quite sure about it upon further inquiry. With nothing to lose I laid him down on the mat and gently guided him into deep integrative breathing. It was hard for him to focus, but we kept at it one breath at a time. We used Integrative Breathwork with self-regulating breathwork principles. About 12 minutes into it he suddenly opened his eyes and said with incredible clarity “You know what? I don’t think I ever have felt unconditionally loved.” And then his eyes cleared from the distant psychotic state they had been in and I saw Jonathan. That was the end of his psychosis. It completely stopped at that moment. We then processed his feelings about that while he was breathing, and he came into a state of loving himself tenderly, and being able to receive my love and other peoples’ love more deeply than he ever had. Now comes the really truly interesting part. All along I had been concerned that he had a severe vitamin B12 deficiency. In one sampling of 280 patients in a psychiatric hospital in Uganda 28.6 percent of the patients had blood serum levels of B12 less than 240 that correlated to higher incidence of schizophrenia or longer duration of psychiatric illness in general (Ssonko et al., 2014). In a clinical and research report, 19 psychiatric patients were identified with low B12 which is correlated with psychiatric manifestations of depression, apathy, irritability, dementia, catatonia, delirium, and hallucinations. A large number of the cases were diagnosed with psychosis. The majority of the patients who had psychosis had prominent Schneiderian first-rank symptoms including “thought alienation phenomenon, commenting and third person auditory hallucinations, delusions of persecution and reference, and passivity phenomenon”. The report also notes that most of the patients had no major physical symptoms that we would associate commonly with vitamin B12 deficiency (Jayaram et al., 2013). So, before the Breathwork session I had requested that he get tested for vitamin B12 and vitamin D levels. The day after the Breathwork session he remained coherent, lucid, not delusional, and behaving normally. He still complained of unusual and complete exhaustion. The next day his blood tests came back, and his vitamin D was 7 when low-normal is 20 or 30 (which many medical doctors consider insufficient for optimal health), and his B12 was 115 when low-normal is 200 (which many medical doctors consider insufficient for optimal health). So, it is very possible that the entire episode could have been triggered by extremely low vitamin B12 and Vitamin D levels. Once he started receiving B12 injections and took high-potency Vitamin D his energy was restored. Yet the interesting, maybe amazing, phenomenon here is that his psychosis completely stopped during his Breathwork session - before he got treated chemically - and it held the whole four days until he had a B12 injection, after several weeks of psychotic affect.

Case Study 2 using Self-regulating Integrative Breathwork Diagnosis: Asthma Daniella (a pseudonym) was a bright, educated, unique 30-year old woman who was very sensitive to her own and others’ feelings and had a great passion for animals, learning, the arts, politics, people, family, and life. She was born into a family of 3 brothers, and a father who dominated her very inside and out beautiful mother. Her mother was unable to ever feel truly loved by her somewhat coarse and 441

 The Science and Art of Contemporary Breathwork and Yoga

certainly philandering husband. Daniella had delighted in contact with the natural world, as well as contact with great ideas and creativity, most of her life, and she suffered greatly from a sense of isolation and feeling unmet, particularly as a teenager. She described “writhing” on her bedroom floor with a silent scream for hours at a time feeling the sense of isolation from her brothers, father, and her ineffectual mother. At the age of 19 she developed an acute asthmatic attack and was rushed to the hospital. An oxygen tent became her home for several days. She remembered crying, wailing, and begging for someone to put their hands in and touch her. The sense of isolation she had during childhood was now manifested outwardly. She didn’t remember anybody trying to touch her during all of those days, except to do medical procedures. Daniella had been attending yoga classes that used simple pranayama techniques, as well as learning about healthy breathing through her training as a Shiatsu Practitioner. Still, she had asthma. In her first integrative breathwork session, we used self-regulating breathwork principles. In this session as the memory of her hospitalization emerged, she began shaking and writhing. The whole set of sensations arose from that time, the trauma of feeling that she would not be able to live because her breathing was so compromised and that she would die untouched - truly a metaphor for her life up to that point. Because of her asthma she could not breathe deep breaths in the Breathwork session without coughing. Her breath was becoming more and more shallow and constricted. I knew she was experiencing the real feeling that her lungs were closing, and she might be at severe risk. It is always an art to decide when touch is corrective and produces evolution for the client, and when touching is soothing but not ultimately transformative. In this case it felt like for her to be touched would be a totally new experience that would lead to significant insight and a “felt sense” of the realness and value of her body. I began to gently touch different parts of her arms, torso, belly, legs, feet, face, hands, spine. Just as her breath was becoming almost frozen, like a spasm that would not allow her to breathe, she describes suddenly feeling the goodness of her lungs and the warmth of them. And all at once, against all odds in her perception, her lungs began to relax and open, and breathe with life force. She went on to do scores of other Breathwork sessions and healed her asthma within the first year. It is now thirty years later, and she still does not own or use an inhaler or any medication. Let’s touch on one more fascinating topic - the apparent release of drugs, alcohol, and anesthesia during some Breathwork sessions. Though there is no known scientific evidence to back this up, anecdotally Breathers experience strong sensations that feel like the reliving and releasing of chemicals, drugs, alcohol and toxins. We know scientifically that drugs and toxins can get stored in body tissue and cells. Though it cannot be said for sure, it appears from the following two stories, and many others like it, that Breathwork will eventually get down to such deep layers of soma and psyche that it can clean out residual drugs, alcohol, and toxins. Robert (pseudonym) had been an alcoholic since he first began drinking when he was 13 and felt his emotional pain disappear. He was 39 when he walked into an AA meeting out of desperation and began what is now over 35 years of recovery. Along with the alcohol path of 26 years, he had included fairly regular use of amphetamines and cocaine, enough that he had a permanent cocaine drip for years in the back of his nose and throat. The disappearance of the drip after sobriety was a delightful, pleasurable experience for Robert that he was really grateful for. As his journey continued, he wanted to grow and found his way to yoga, meditation, and eventually Integrative Breathwork using self-regulating principles. Each Breathwork session felt liberating to him, even as he was unpacking memories of physical abuse from his father and sexual abuse from a Boy Scout master. Each session resulted in his life feeling more fulfilled, with the depression he had carried with him for many years slowly and steadily lifting. 442

 The Science and Art of Contemporary Breathwork and Yoga

Spiritual experiences of connection with a higher power were becoming more frequent. He was able to take care of his personal and professional relationships without retreating and hiding. Life had become really good when, one day, Robert showed up for what was likely his 30th Breathwork session. He was feeling filled with life force and optimism. Imagine his surprise when in the middle of the session he began to experience a very distinct cocaine drip in the back of his throat, though he had not used cocaine for over four years, nor had a cocaine drip. In the orientation of Group 5 Breathwork he stayed present with this sensation and allowed himself to cry, to shake, and to process the images and memories that were arising in his mind’s eye. He was calm at the end of the session and went home with a spirit of curiosity about what was happening. For about five days he had a small cocaine drip and then it was gone, never to return. Marjorie (pseudonym) was about two years into her Breathwork journey and had done about 35 sessions of Integrative Breathwork using self-regulating principles, when she had a remarkable experience. She began reliving her birth while her adult awareness was able to speak about everything she was experiencing. At one point she cried out that she could smell a gas that did not smell good and she started to fade out. She kept fighting it to come into consciousness and insisted there was the presence of what she called, “something like ether.” I would have been skeptical about this claim, except that right from the beginning I was assailed with an odor of chloroform that was so powerful it was everything I could do to keep my head from falling to my chest. I had a duty to stay present to her, yet my body kept trying to pass out and the smell was very strong. Forcing myself to jump up, I threw open two windows in the room and still had to put my head right into the window and breathe the cold winter air to prevent from falling asleep. This whole incident took about ten minutes and finally the smell subsided. Subsequent to that session, Marjorie found discovered from her mother that, indeed, her mother had been given chloroform at the birth which had taken place in the first half of the 1900’s. Marjorie also reported that it felt like she was “seeing life in a new way,” and that her “brain was more awake.” The Breathworker of the future, trained to help the client self-regulate their breathing for optimal well-being and coherence of soma, psyche and soul will be a skilled midwife of well-being, human potential, and spiritual awakening. They will be proficient in many breathing techniques, yet know that individual attunement and artistry is the most powerful method.

REFERENCES Atkinson, N. L., & Permuth-Levine, R. (2009). Benefits, barriers, and cues to action of yoga practice: A focus group approach. American Journal of Health Behavior, 33(1), 3–14. doi:10.5993/AJHB.33.1.1 PMID:18844516 Bernardi, L., Schneider, A., Pomidori, L., Paolucci, E., & Cogo, A. (2006). Hypoxic ventilatory response in successful extreme high altitude climbers. The European Respiratory Journal, 27(1), 165–171. doi: 10.1183/09031936.06.00015805 PMID:16387950 Bond, A. J. (2001). Neurotransmitters, temperament and social functioning. European Neuropsychopharmacology, 11(4), 261–274. doi:10.1016/S0924-977X(01)00094-3 PMID:11532380 Brown, R., & Gerbarg, P. (2012). The healing power of the breath: Simple techniques to reduce stress and anxiety, enhance concentration and balance your emotions. Shambhala Publications.

443

 The Science and Art of Contemporary Breathwork and Yoga

Brown, R., Gerbarg, P., & Sudarshan, K. (2005a). Yogic breathing in the treatment of stress, anxiety, and depression: Part 1 – Neurophysiologic Model. Journal of Alternative and Complementary Medicine (New York, N.Y.), 11, 189–201. doi:10.1089/acm.2005.11.189 PMID:15750381 Brown, R., Gerbarg, P., & Sudarshan, K. (2005b). Yogic breathing in the treatment of stress, anxiety, and depression: Part 2 – Applications and Guidelines. Journal of Alternative and Complementary Medicine (New York, N.Y.), 11, 711–717. doi:10.1089/acm.2005.11.711 PMID:16131297 Collins, C. (1998). Yoga: Intuition, preventive medicine, and treatment. Journal of Obstetric, Gynecologic, and Neonatal Nursing, 27(5), 563–568. doi:10.1111/j.1552-6909.1998.tb02623.x PMID:9773368 Davidson, R. J. (2014). One of a kind: The neurobiology of individuality. Cerebrum, 2014, 8. PMID:26000075 Ehrmann, W. (2011). The healing role of breathwork for depression based on the polyvagal theory. Breath & Inspiration, 55, 9–11. Elliott, S., & Edmonson, D. (2008). Coherent Breathing: The definitive method - theory and practice. Coherence Press. Forbes, B. (2011). Yoga for Emotional Balance: Simple Practices to Relieve Anxiety and Depression. Shambhala Publications. Goldin, P., Ziv, M., Jazaieri, H., Hahn, K., & Gross, J. (2013). MBSR vs aerobic exercise in social anxiety: fMRI of emotion regulation of negative self-beliefs. Social Cognitive and Affective Neuroscience, 8(1), 65–72. doi:10.1093can/nss054 PMID:22586252 Grof, S. (2000). Psychology of the future lessons from modern consciousness research. State University of New York Press. Heart Math Institute. (2001). Science of the Heart: Volume 1 (1993-2001) exploring the role of the heart in human performance. Retrieved from https://www.heartmath.org/resources/downloads/science-ofthe-heart/ Jabr, F. (2012). Know your neurons: What is the ratio of glia to neurons in the brain? Scientific American Blog Network. Retrieved from https://blogs.scientificamerican.com/brainwaves/know-your-neuronswhat-is-the-ratio-of-glia-to-neurons-in-the-brain/ Jayaram, N., Rao, M. G., Narasimha, A., Raveendranathan, D., Varambally, S., Venkatasubramanian, G., & Gangadhar, B. N. (2013). Vitamin B12 Levels and Psychiatric Symptomatology: A Case Series. The Journal of Neuropsychiatry and Clinical Neurosciences, 3. PMID:23686033 Lalande, L., Bambling, M., King, R., & Roger, L. (2012). Breathwork: An additional treatment option for depression and anxiety? Journal of Contemporary Psychotherapy, 42(2), 113–119. doi:10.100710879011-9180-6 Levine, J. (2018). The science of breathing. Retrieved from https://www.yogajournal.com/yoga-101/ science-breathing

444

 The Science and Art of Contemporary Breathwork and Yoga

McFadden, C., Sandler, T., & Fieldstadt, E. (2014). San Francisco schools transformed by the power of meditation. Retrieved from https://www.nbcnews.com/nightly-news/san-francisco-schools-transformedpower-meditation-n276301 Miller, J., Fletcher, K., & Kabat-Zinn, J. (1995). Three-year follow-up and clinical implications of a mindfulness meditation-based stress reduction intervention in the treatment of anxiety disorders. General Hospital Psychiatry, 17(3), 192–200. doi:10.1016/0163-8343(95)00025-M PMID:7649463 Minett, G. (2004). Exhale: An overview of breathwork. Floris Books. Morningstar, J. (1994). Breathing in light and love: Your call to breath and body mastery. Transformations. Morningstar, J. (2017). Break Through with Breathwork: jump-starting personal growth in counseling and the healing arts. North Atlantic Books. Mrazek, M., Franklin, M., Phillips, D., Baird, B., & Schooler, J. (2013). Mindfulness training improves working memory capacity and GRE performance while reducing mind wandering. Psychological Science, 24(5), 776–781. doi:10.1177/0956797612459659 PMID:23538911 Nettle, D. (2006). Happiness: The science behind your smile. Oxford University Press. Niemi-Pynttäri, J., Sund, R., Putkonen, H., Vorma, H., Wahlbeck, K., & Pirkola, S. (2013). Substanceinduced psychoses converting into schizophrenia: A register-based study of 18,478 Finnish inpatient cases. The Journal of Clinical Psychiatry, 74(1), e94–e99. doi:10.4088/JCP.12m07822 PMID:23419236 Novotney, A. (2009). Yoga as a practice tool. America Psychological Association. Monitor on Psychology, 40. O’Brien, E. (2015). The art & science of breath. Retrieved from https://www.fivepillarsyoga.com/theart-science-of-breath/ Oken, B. S., Zajdel, D., Kishiyama, S., Flegal, K., Dehen, C., Haas, M., ... Leyva, J. (2006). Randomized, controlled, six-month trial of yoga in healthy seniors: Effects on cognition and quality of life. Alternative Therapies in Health and Medicine, 12, 40–47. PMID:16454146 Rajski, P. (2002). Alcoholism and rebirthing. International Journal for the Advancement of Counseling, 24(2), 123–136. doi:10.1023/A:1020963817447 Ssonko, M., Ddungu, H., & Musisi, S. (2014). Low serum vitamin B12 levels among psychiatric patients admitted in Butabika mental hospital in Uganda. BMC Research Notes, 7(1), 90. doi:10.1186/17560500-7-90 PMID:24533701 Strom, M. (2012). A life worth breathing: A yoga master’s handbook of strength, grace, and healing. Skyhorse Publishing Inc. Telles, S., Naveen, K., Balkrishna, A., & Kumar, S. (2010). Short term health impact of a yoga and diet change program on obesity. Medical Science Monitor, 16(1), CR35–CR40. PMID:20037492 Wilber, K. (2006). Integral Spirituality: A startling new role for religion in the modern and postmodern world. Integral Books.

445

 The Science and Art of Contemporary Breathwork and Yoga

Woodyard, C. (2011). Exploring the therapeutic effects of yoga and its ability to increase quality of life. International Journal of Yoga, 4(2), 49–54. doi:10.4103/0973-6131.85485 PMID:22022122 Zeidan, F., Martucci, K., Kraft, R., McHaffie, J., & Coghill, R. (2014). Neural correlates of mindfulness meditation-related anxiety relief. Social Cognitive and Affective Neuroscience, 9(6), 751–759. doi:10.1093can/nst041 PMID:23615765

446

The Science and Art of Contemporary Breathwork and Yoga

APPENDIX 1

Transformations – School of Integrative Psychology Therapeutic Breathwork and YogaAsana Research Project Project Protocols Overview I.

Yoga Sessions with Breathing Exercises

Experimentation with a slower-than-normal and a faster-than-normal breathing technique with specific asanas Purpose: to document changes in a) b) c) d)

Flexibility Endurance Emotional states of calm/anxiety Holding patterns/trauma release

Method: • •

• • • •

Recruit volunteers and explain project Rate level of familiarity with yoga and with breathwork Beginner (1-3 months) Medium (3 months - 5 years) Expert (5 years +) Initiate group or individual session Do series of postures and breathing exercises Administer rating scale to practitioner and volunteer Report results II. Breathwork Sessions Incorporating Yoga Asanas

Experimentation with using specific yoga asanas before and/or during a faster-than-normal breathing session Purpose: To document usefulness of asanas during breathwork to a) Focus attention on the breath b) Focus attention on parts of body needing release c) Support process of releasing holding patterns and/or trauma Method:

447

The Science and Art of Contemporary Breathwork and Yoga

• • • • • •

Recruit volunteers and explain project Rate level of familiarity with breathwork and with yoga Initiate group or individual session Conduct breathwork session with specific asanas before and/or during session Administer rating scale to practitioner and volunteer Report results

Transformations – School of Integrative Psychology Therapeutic Breathwork and Yoga Asana Research Project Project Protocols Overview I. Sample Protocols for Using Various Breathing Techniques with Yoga Asanas 1. Prep volunteer(s) with purpose of session: to document changes in flexibility, endurance, emotional state, and holding pattern release using slower- and faster-than-normal breathing techniques during yoga postures. Research done in conjunction with Transformations Incorporated and the School of Integrative Psychology, Director Jim Morningstar, PhD 2. Sign release form to use data and accept self-responsibility 3. Conduct yoga session (30-90 minutes) employing 4-6 asanas and 2-3 forms of breathing with each posture (see asana options handout) a) Explain, demonstrate, and practice breathing techniques before session (i) Traditional Pranayama – slower-than-normal (optional, if volunteer is already adept) (ii) Slower-than-normal Stress Release Full inhale through nose, Hold for 2-3 seconds, Exhale through mouth with pursed lips (4-5 minutes) (iii) Faster-than-normal Conscious Connected Breathing – Inhale and exhale through mouth (can switch to nose if necessary) without pause for 100 breaths b) Solicit from volunteer what they would like to release from their body, mind, spirit during session and what they would like to gain in their body, mind, spirit during the session c) Conduct session d) Complete practitioner and volunteer feedback forms and return to Jim Morningstar, Ph.D.

Transformations – School of Integrative Psychology Therapeutic Breathwork and Yoga Asana Research Project Project Protocols Overview II. Sample Protocol for A (Faster-Than-Normal) Breathing Session Incorporating Yoga Asanas

448

The Science and Art of Contemporary Breathwork and Yoga

1. Prep volunteer(s) with purpose of session: to document changes in flexibility, endurance, emotional state, and holding pattern release using slower- and faster-than-normal breathing techniques during yoga postures. Research done in conjunction with Transformations Incorporated and the School of Integrative Psychology, Director Jim Morningstar, Ph.D. 2. Sign release form to use data and accept self-responsibility 3. Conduct breathwork session (30-90 minutes) employing 1-3 asanas and at least one-than-normal and one faster-than-normal form of breathing with each posture (see asana options handout) a) Explain, demonstrate, and practice breathing techniques before session (i)Traditional Pranayama – slower-than-normal (optional, if volunteer is already adept) (ii) Slower-than-normal Stress Release Full inhale through nose, Hold for 2-3 seconds, Exhale through mouth with pursed lips (4-5 minutes) (iii) Faster-than-normal Conscious Connected Breathing – Inhale and exhale through mouth (can switch to nose if necessary) without pause for 100 breaths b) Solicit from volunteer what they would like to release from their body, mind, spirit during session and what they would like to gain in their body, mind, spirit during the session c) Conduct session d) Complete practitioner and volunteer feedback forms and return to Jim Morningstar, Ph.D.

Transformations – School of Integrative Psychology Therapeutic Breathwork and YogaAsana Research Project Asana Options

Standing Poses Tadasana (Mountain pose) Virabhadrasana II (Warrior pose II) Utthita trikonasana (Extended triangle pose) Sitting Poses Sukhasana (Easy pose) Upavistha konasana (Seated wide-angle pose) Ardha matsyendrasana (Half lord of the fishes pose) Kneeling Poses Balasana (Child’s pose) Ustrasana (Camel pose)

449

The Science and Art of Contemporary Breathwork and Yoga

Simhasana (Lion pose) Supine Poses Savasana (Corpse pose) Setu Bandhasana (Bridge pose) Matsyasana (Fish pose) Prone Poses Bhujangasana (Cobra pose) Dhanurasana (Bow pose) Salabhasana (Locust pose) Arm Support Poses Adho mukha svanasana (Downward-facing dog pose) Chaturanga dandasana (Four-limbed stick pose) Chatus pada pitham (Four-footed tabletop pose)

450

451

Chapter 25

An Exploration of Influence of Duration on Physiological Effects of Asanas Debadutta Subudhi Indian Institute of Technology, Madras, India M. Manivannan Indian Institute of Technology, Madras, India

ABSTRACT Different body postures that can be maintained for a certain duration with awareness are called yogaasanas. These asanas matter flexibility, coordination, and strength, while the breathing practices and meditation sharpens the mind for better awareness and reduce anxiety and thus adds quality into life. Other beneficial effects might involve a reduction of stress, blood pressure, and improvements in resilience, mood, and metabolic regulation. The asanas performed regularly for a short duration in hours is well studied in the literature. However, when performed for a long duration continuously for several hours (40-hour yogathon), without food and sleep has a significant effect in regulating homeostasis. The homeostasis is accessed through cardio-respiratory and galvanic skin response changes. The study shows the physiological changes after the yogathon and compares it with effects on physiology due to short term yoga. It also emphasizes on the reduction of dependency on food, because of energy compensation through yoga-asanas.

INTRODUCTION Physiological and Psychological Effects of Asanas Asana (physical posture) is described as the body posture attuning stability along with ease (Bhavanani et al., 2018). Asanas have physiological significance in regulating homeostasis and autonomic balance through massage of internal organs that clears the blockages in the energy channels of the body and the DOI: 10.4018/978-1-7998-3254-6.ch025

Copyright © 2021, IGI Global. Copying or distributing in print or electronic forms without written permission of IGI Global is prohibited.

 An Exploration of Influence of Duration on Physiological Effects of Asanas

body energy system becomes more balanced (Woodyard et al., 2011). The range of motion of human body parts reduces as it gets aged. A set of asanas brings a full range of motion of all joints thereby enhancing flexibility with time (Petric et al., 2014). Combination of yoga-asanas with relaxing postures such as savasana, balasana improves the cardiopulmonary system, diminishing O2 consumption (Telles et al., 2000). Immediate effect of 30 minutes yoga-asanas brings the mind to a calm state with reduction of GSR (De et al., 2019). The duration of performance of asanas has a significant effect on the physiology. Depending on the duration asanas it can be classified into two types such as short-term and long term. Short-term yogaasana is defined as yoga-asanas carried out for hours (< 2 hours) in one or two sessions per day on a weekly or monthly basis. Whereas the long term yoga-asana is carried for more than a month up to 6 months. Most of the literature is based on the above two categories. However, the authors in this chapter include another classification called ‘Yogathon’. Yogathon is defined as performing asanas more than 12 hours, continuously at a stretch. There are no studies known regarding how it affects physiology and psychology of the subjects. This chapter analyses the physiological and psychological effects on health after performing 40 hours of yoga-asana without food and sleep. The physiological aspect is assessed from the variability of cardiac and pulmonary systems called cardio-respiratory variability (CRV). The psychological aspect is assessed from the galvanic skin response (GSR). CRV consists of heart rate variability (HRV), pulse rate variability (PRV), and breath rate variability (BRV). The variability depicts the characteristics of the autonomic nervous system (ANS). The ANS has sympathetic and parasympathetic divisions which control the internal organs. The authors consider the cardio-pulmonary system to describe changes in the sympathetic nervous system (SNS) and the parasympathetic nervous system (PSNS). The yogathon is generalized for the individuals, practicing asanas regularly for at least 2 to 3 years and yogathon should be performed under guidance.

OBJECTIVES The chapter focuses on evaluating physiological and psychological effects due to yogathon for 40 hours with fasting and without sleep. The results of yogathon are compared with that of short-term yoga-asanas, obtained from existing literature. Specifically, the chapter focuses on: a. Changes in CRV after yogathon. b. Changes in CRV due to no intake of solid foods (Water and juices are allowed as prescribed by the committee). c. Changes in CRV without sleep for 40 hours, however subjects are resting in Corpse-pose, childpose, lotus-pose, reclining-bound-angle-pose, easy-pose and various reclining pose. The Changes in HRV, PRV, BRV due to no intake of solid foods and without sleep are obtained from the literature. The authors also show how asanas help the self in managing metabolic energy consumption by interventions other than food and sleep.

452

 An Exploration of Influence of Duration on Physiological Effects of Asanas

BACKGROUND This section elaborates CRV and its components along with their relationship to physiology. Variability is an indication of adaptability of the cardiorespiratory system to metabolic activities. Generally, a higher variability implies a better health.

Heart Rate Variability (HRV) The electrical activity of the heart is reflected through ECG and the interval between their peaks is known as the inter-beat interval (IBI). The changes in these intervals is known as HRV. The heart is the prime pumping station in the human body which accounts for the need of different organ systems such as nervous, circulatory, muscular, skeletal, endocrine, lymphatic, respiratory, digestive, excretory and reproductive. Each of the organ systems has controls from ANS such as SNS and PSNS. The variation in the heart rate is due to the control mechanism of ANS along with feedback from baroreceptor and chemoreceptor present in heart, cardio-respiratory coupling, metabolic rate. All these factors cause the heart rate to change continuously measured as HRV. The autonomous nervous system has bi-control mechanisms such as internal and external. Internal controls are heart rate, digestion and pupil dilatation whereas external control is for bringing the body to a stable state against external changes, which is termed as the homeostatic control. The ANS controls through the sympathetic, parasympathetic and enteric nervous system. The neurotransmitters in sympathetic division cause energy expenditure for different motor activities and heart rate or pulse rate rises. The neurotransmitters in parasympathetic division cause energy restoration through rest and recovery along with the revival of normal heart rate. The neurotransmitters in the enteric division are responsible for normal digestion function. The sympathetic stimulation inhibits gastrointestinal secretion and motor activity causing the narrowing of gastrointestinal sphincters. Whereas parasympathetic stimuli assist digestive activities. Therefore, most of the yoga sessions are performed in an empty belly. The major components in HRV are respiratory sinus arrhythmia (RSA), thermoregulation, and baroreceptor reflex. RSA is defined as a rise in heart rate due to respiration which occurs in the high frequency (HF) band of HRV which is 0.25 Hz (Giddens et al., 1985). The thermoregulation is an important parameter for keeping core temperature at 370C. Heat is continuously generated in the body due to variation in basal metabolic rate as well as in peripheral vasomotor tone, therefore circulation from the heart helps in keeping core temperature constant apart from other mechanisms present in the human body. The thermoregulation resembles a very low frequency band of RR-interval (0.0033 - 0.04 Hz). The baroreceptors in the heart are there to sense the pressure change in the heart due to admission and release of blood and the frequency of baroreceptor reflex is about 0.12 Hz (Hyndman et al., 1971; Kitney et al., 1975). HRV assesses numerous pathological states such as ventricular arrhythmias, hypertension, sleep apnea, fetal status, assessment of diabetic patients, monitoring patients with a heart transplant. It is used to assess changes in ANS during meditation, yoga, pranayama (breathing practice), physical endurance tests (Marathons in running, swimming, cycling etc.). It is also used as a biomarker for healthiness. The authors of this chapter used HRV to analyze the effects of Yogathon on ANS control. The HRV is obtained from the change of RR-interval after filtering the ECG signal or from the PP-interval of PPG signal which is referred to as Pulse Rate Variability (PRV). The pulse signal is obtained usually from the index finger. HRV from RR interval captures the electrical activity of the heart whereas HRV from 453

 An Exploration of Influence of Duration on Physiological Effects of Asanas

PP-interval captures the mechanical activity of the cardio-respiratory system. There is a good correlation between PRV and HRV for all time, frequency and nonlinear variables while subjects are at rest (Selvaraj et al., 2008; Shi et al., 2008) or during night sleep (Hayano et al., 2005), while the agreement differs during exercise (Gil et al., 2010; Lu et al., 2008). The change is due to mechanical coupling between respiration and the thoracic vascular system (Schäfer et al., 2013). Hence the authors consider both ECG and PPG to comment on HRV after 40 hours of Yogathon.

Breath Rate Variability (BRV) The breath signal is obtained from PPG by filtering it in the 0.1-0.5 Hz frequency ban (Soni et al., 2019). The change in breath rate over time is known as breath rate variability (BRV). The respiration combines the signal from the chemoreceptors present in the heart and brain. Therefore, change in breathing cycle accompanies a relative change in concentration of oxygen, carbon dioxide, the pH level of blood. The change in breath rate also describes psychophysiological states of the human being. The yoga postures coupled with breathing patterns stimulate vagal nerves which raise thalamic Vagal-Gamma-Aminobutyric Acid (GABA) levels for 15 subjects after 12-week yoga intervention (Streeter et al., 2018). The changes in these chemical concentrations alter breath continuously during yogathon. The significant increase in GABA receptors (p = 0.09) imparts in improved psychophysiological states (Telles et al., 2012). The increase in GABA at the medullary ventral surface causes respiratory depression, hypotension, and bradycardia (Yamada et al., 1982). The change in breath-breath interval (BBI) constitutes BRV. The authors consider breath as an important index as the changes in respiration pattern cause differences between PRV and HRV as it influences the variation in pulse as well as heart rate. The retention of asanas for a certain duration with different variations causes novice breathing patterns which subjects control voluntarily. Moreover, performing the asanas for a long duration without food allows the body to extract energy required for metabolic activity from the breath which is not explored when the duration is in hours. This aspect adds uniqueness to yogathon than other marathon events which cause the body to exhaust.

Cardio-respiratory Variability (CRV) The cardio-respiratory coupling occurs peripherally as well as centrally (Koepchen et al., 1995). PRV responds to peripheral cardiac change whereas HRV responds to central cardiac change. BRV is the connecting link between HRV and PRV. Therefore, cardiorespiratory variability constitutes HRV, PRV and BRV. Moreover, the electrical activity of the heart is reflected through ECG whereas the mechanical activity of the heart is reflected through PPG. The interval between their peaks is called the inter-beat interval (IBI). The change in these intervals over the recording period is known as HRV from ECG and PRV from PPG. HRV and PRV are the same in time and frequency during the resting condition (Selvaraj et al., 2008; Shi et al., 2018). However, HRV and PRV are not the same during and after any asanas (Gil et al., 2010). The breath signal is obtained from PPG by filtering it in the 0.1-0.5 Hz frequency band. The change in breath rate over time is known as breath rate variability (BRV). The HRV, PRV, BRV together constitute cardiorespiratory variability. The authors introduce CRV to signify the ability of the cardiorespiratory system to adapt subsequent changes during metabolic activity. In general, a higher CRV is an index for good health. Yogathon endures the cardio-respiratory system, and the impact on the functionality is evaluated through CRV. 454

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Fasting Human civilization carries out fasting and starvation since the beginning of evolution, due to life circumstances such as seasonal changes, imprisoning, war and as a part of religious activity (Ekadashi, Ramadan, Lenten season or politically motivated hunger strike (Johnstone et al., 2007). Minnesota Starvation Study details the biological changes of long-term starvation in the 1940s for the 1st time (Ancel et al., 1950). Currently fasting has become a medical and scientific topic. Yoshinori Ohsumi (Japanese cell biologist) won the Nobel Prize in Medicine in 2016 for his research on autophagy which recycles and renews cells in body and fasting activates autophagy (Mizushima et al., 2014). The effect of long term fasting on HRV is studied by many researchers. In India long term fasting is encouraged since it is embedded in culture religiously. The long term fasting lasts for days whereas intermittent fasting runs for hours. The results show a reduction in SDNN, RMSSD with induced parasympathetic inhibition and sympathetic activation. Therefore, HRV reduces after long term fasting. Whereas, the GSR remains unchanged. A study by Mazurak et al. (2013) shows body mass index (BMI), diurnal cortisol profile and HRV significantly lowers after 2 days of fasting considering 16 young healthy female volunteers. Seven women [age 22·4 ± 3·3 years, body mass index (BMI) 21·5 ± 2·0 kg/m2 (mean ± SD)] participated in a study of fasting with one subject taking leptin admission. Leptin is used to reverse the effect of a decline in SNS activity. The study is for the validation of the hypothesis that short-term fasting would increase SNS activity and decrease vagal tone in humans occurring independently of leptin. The same subjects have gone through three plausible conditions: a baseline control study on isocaloric feeding; 72-h fasting with placebo, which reduces serum leptin levels by approximately 80 percent; and 72-h fasting with the injection of r-metHuLeptin, that restores the fasting-induced decline in leptin levels. Both leptin and placebo injection show an increase in heart rate with a reduction in SDNN and RMSSD. The study also shows the retrieval of both sympathetic and parasympathetic control of ANS on the heart due to reduction in LF and HF power (Chan et al., 2007). The GSR remains unchanged for 5 subjects from two groups, fasting for 4 days (Kollar et al., 1964). None of the studies on fasting, either short time or long time, consider physiological changes while performing yoga-asanas.

Sleeplessness Sleep is the basic part of life and generally, it takes 6-8 hour per day. Sleep mediates recovery from physical and mental stress with the rejuvenation of the body and mind. It helps in growth, repair and preservation of physical functions (Schmidt et al., 2014). Sleep affects HRV and blood pressure, respiratory rate, body temperature, metabolic and immunity (Altevogt et al., 2006). The HRV reduces because of sleep deprivation (SD) as shown in many studies. The skin conductance decreases during onset of sleep and ascends up during waking state. SD for 5 subjects for 123 hours shows a reduction trend in palm conductance, respiratory sinus arrhythmia (RSA) due to long wave activity (Change in heart rate occurs for 2 or more respirations) as the day progresses (AX et al., 1961). A 9 minutes of ECG before and after sleep onset in normal subjects (12) shows a reduction in VLF and LF power of RR-interval (RRI) and an elevation in HF- power of RRI significantly (p < 0.05). The heart rate declines with an increase in RRI after the sleep onset (SO). The transition from before and after SO accompanies a waving decrease in RRI variability and respiratory frequency variability (BRV) (Shinar et al., 2006). Twenty young subjects, while undergoing a sleep deprivation period of 60 hours show a reduced heart rate, increased LF as well as HF power significantly and a reduced variability (Vaara et 455

 An Exploration of Influence of Duration on Physiological Effects of Asanas

al., 2009). SD for 36 - 40 hour in healthy subjects causes an increase in HR with higher normalized LF power and a lower normalized HF power (Zhong et al., 2005; Sauvet et al., 2009). Normalized LF blood pressure variability is significantly higher throughout the SD duration. Baroreflex sensitivity (BRS) significantly declines after SD (Zhong et al., 2005). Hence sleeplessness causes reduction in HRV with reduced VLF power, HF-power and an increase in LF power of RRI. None of the studies on changes in SDs consider physiological changes while performing yoga asanas. The GSR decreases to about 4 GSRs/min during Stage-1 and Stage-2 of sleep than that of waking state (12 GSRs/min) whereas basal skin resistance increases smoothly during sleep and drops down in waking state (Johns et al., 1969).

Combined Effect of Fasting and Sleeplessness The combined effect of fasting, sleeplessness and any amount of physical stress might amplify the effects of the various parameters that the authors discussed. However, the performance of continuous longduration yoga-asanas without food and sleep unfolds novice energy management of the body. Its effect on the cardiovascular system is not studied before. Also a little is known about asanas compensating for metabolic energy consumption while performing for longer duration continuously in the absence of primitive energy sources such as food and sleep. This chapter includes effects on CRV and GSR due to yogathon. It also presents a comparative study between short term yoga-asanas and yogathon in terms of HRV and GSR.

EXPERIMENTAL METHODOLOGY Yogathon includes analysis of RRI, PPI, BBI and GSR for better understanding of interdependence among each other to maintain homeostasis. However, the literature on short term yoga is primarily focusing on RRI. The yogathon involves prolonged practice of asanas for 40 hours, therefore the physiological changes are imparted by fasting and sleep deprivation. As the focus of this chapter is to study the physiological and psychological effects of asanas alone, without food and sleep, this section describes existing studies on the physiological effects solely due to fasting and sleeplessness without asanas.

Experimental Protocol The current study of yogathon constitutes randomized clinical trial of 7 out of 120 female participants of age (37 ± 14), weight (60 ± 8), height (160 ± 6), Body Mass Index (BMI) (23 ± 3) on International Yoga day by Mahamaharishi Foundation. The yogathon session starts from morning 6.00 am (20-62018) and continues until 10.00 pm next day (21-6-2018) at Indira Gandhi Indoor Stadium (Upalam, Puducherry). A total of 1696 yoga-asanas are performed cyclically up to 40 hours along with resting poses in-between. The inclusion criteria for the subjects include the experience of 3 to 5 year in yoga. The subjects should be with normal health conditions without any medications. The vital parameters such as Electrocardiogram (ECG), Photo plethysmography (PPG), galvanic skin response (GSR) are recorded, before and after performing the yoga-asanas using Biopac (MP-36) for data collection, at 200Hz in the supine position. PPG is recorded from the forefinger of the left hand and GSR is recorded by placing two Ag-AgCl electrodes on the right-hand palm. The ambient temperature is maintained at 25 ± 20C. Inter-beat Intervals from ECG and PPG are measured and variability among 456

 An Exploration of Influence of Duration on Physiological Effects of Asanas

these intervals of ECG and PPG are considered as HRV and PRV. The authors consider time domain, frequency domain and nonlinear measures of ECG and PPG along with GSR. The authors consider a mean change of greater than 10 percent as significant variation (Soni et al., 2019). The current study shows variabilities in heart rate, pulse rate and breath rate. The standard deviation in RR interval (SDRR), root mean square of successive difference (RMSSD), the maximum variation in heart rate (HRMV) are used to quantify HRV in time (Robert et al.,1992), similarly it follows for pulse rate and breath rate. Variability is also quantified from energy under different frequency bands (VLF, LF, HF) and nonlinear changes in RR-interval (RRI), pulse to pulse interval (PPI), breath to breath interval (BBI). The energy is obtained from the Lomb-Scargle Periodogram of RRI, PPI, BBI and nonlinear changes are obtained from Poincare’s plot. Poincare’s plot is the dispersion of the RR interval with the successive RR-interval. Figure 1 indicates different intervals for a typical participant. The literature on short term yoga is primarily focusing around RRI, however, analysis of RRI, PPI and BBI provide a better understanding of interdependence among each other to maintain homeostasis. Figure 1. RR-interval (RRI), Pulse Interval (PPI), Breath interval (BBI) for a participant

Experimental Data Analysis Time Domain Measures The authors consider SDRR, RMSSD, HRMV as time domain measures for evaluation (Kleiger et al., 1992).

Standard Deviation of RR Intervals(SDRR) SDRR is the standard deviation of all RR intervals. The difference in SDRR is due to relative changes in VLF and LF bands. Therefore, it is associated with thermoregulation and sympathetic tone of the autonomic nervous system (ANS) (Shaffer et al., 2017). A similar approach is followed for PPI and BBI.

457

 An Exploration of Influence of Duration on Physiological Effects of Asanas

SDDR =

2 1 N RRI i − mean(RRI )) (1) ∑ i =1 ( N −1

Root Mean Square of Successive Differences (RMSSD) RMSSD is the square root of the mean of the square of the successive differences between adjacent intervals. It signifies vagally mediated changes or the HF-power in HRV (Shaffer et al., 2017). A similar approach is followed for PPI and BBI. RMSSD =

2 1 * ∑ iN=1 (RRI i +1 − RRI i ) (2) N −1

Heart Rate Maximum Variation (HRMV) HRMV is the difference between the highest and lowest rates during each signal cycle. The maximum heart rate is generally mediated by respiratory sinus arrhythmia (RSA) in the low-frequency domain (0.04- 0.15 Hz) for a BR between 9 to 25 breath/min. Therefore, it is a measure of the impact of RSA on HR. Depending on BR, RSA may shift to any frequency zone (VLF, LF and HF) of HRV. Similarly, for breath rate and pulse rate, the authors defined BRMV and PRMV. A similar approach is followed for PPI and BBI. HRMV = Max heart rate − Minheart rate (3)

Frequency Domain Measures The oscillations of ECG, PPG, and Breath Signal have very-low-frequency (VLF-0.0033-0.04 Hz), low-frequency (LF-0.04-0.15), and high-frequency (HF-0.15-0.4 Hz) bands. The VLF band corresponds to long-term regulation such as thermoregulation, the renin-angiotensin system (Akselrod et al., 1981; McCraty et al., 2015). VLF rhythm is caused due to stimulation of afferent sensory neurons in the heart which is responsible for controlling heart function itself and is produced by the heart itself and fundamental to good health (Kember et al., 2000, 2001). LF-power and HF power correspond to a sympathetic and parasympathetic branch of ANS (Shaffer et al., 2017). The intervals of signals are resampled to 200 Hz before performing Lomb-Scargle power (LSP) estimation as it enhances accuracy in energy content under different frequency ranges than Fourier Transform as Lomb Scargle method is independent of baseline drift of signal-interval (Saini et al., 2008). The LSP of real-valued data sequence X of length N for arbitrary times ‘t’ is computed as described in (Ruf et al., 1990).

Nonlinear Measures Poincar´e Parameters: The rhythmic temporal interval change in time-varying signals is depicted by plotting the preceding interval to the current interval known as Poincar´e plot. This shows a two-dimensional variation along the line of equity (LOE) (slope = 1) and perpendicular to LOE. The dispersion is found in the scatter plot of the ith interval with interval, where m is the lag of the dispersion. A lag is defined as order of the interval in the time series signal. The dispersion along the perpendicular to the LOE shows an increase or decrease in rate, defined as short term variability (SD1). The dispersion along the LOE shows long term variability (SD2) (Piskorski et al., 2007). The term SD stands for standard deviation

458

 An Exploration of Influence of Duration on Physiological Effects of Asanas

and SD1, SD2 are respective standard deviations in the Cartesian plane about LOE. The points lying on LOE signify that the current interval and preceding interval are the same (No change in rate). The points lying above LOE signify that the current interval is less than the preceding interval (decreasing rate). The points lying below LOE signify that the preceding interval is less than the current interval (increasing rate). Therefore, distribution of preceding intervals with respect to current intervals will have biaxial standard deviation. The standard deviations are the measure of variability. The chapter considers ECG, PPG and breathing signals, therefore a higher variability in each of the signals after yogathon is a measure of functionality of cardio-respiratory system. The lagged Poincare analysis shows propagation of variability as the successive difference between intervals increases. Each lag is associated with different SD1 and SD2. The range of SD1 and SD2 (ms) lie within minimum (2) and maximum lag (16) signifies steadiness of the intervals over a particular time segment. I 1 = I (1, 2, 3,...................(L − m )) (4) I 2 = I (m + 1, m + 2, m + 3................L ) (5)

where m is the lag (Order of consecutive difference in the peaks of time series signal) and L is the length of the time series signal. ‘I’ is the interval in any time series signal. SD1 = var(I 1 − I 2 ) / 2 (6) SD2 = var(I 1 + I 2 ) / 2 (7)

Galvanic Skin Response Galvanic skin response refers to the activity of eccrine sweat glands present in the hypodermis and is innervated by cholinergic fibers whose discharge depends on core body temperature whenever recorded from sites other than the palm and foot. During regular exercise or performing asanas the core body temperature increases, hence the sweat secretion becomes greater at the forehead (2.90 mg · cm−2 · min−1; ± 0.19) than the hand (1.49 mg · cm−2 · min−1; ± 0.27) as observed in 10 subjects (Machado et al., 2008). While sweating on dorsal sites such as palm shows no significant difference (Machado et al., 2008). The emotional state which deals with premotor cortex and amygdala (part of the limbic system) (Tranel et al., 1994; Asahina et al., 2003) is observed at the palm and sole sites that contains higher density of sweat glands (644 /cm2) (Wilke et al., 2004). Hence sweat gland activity on palm resembles the emotional stability of the subjects. Therefore, the authors consider the palm as an active site for recording GSR. GSR is recorded on the palm site using Biopac at a sampling frequency of 200 Hz for 6 minutes. The skin conductance response (SCR) is obtained after the deconvolution of GSR with the Bateman function, which has two components such as tonic driver (dc-component) superimposed with phasic driver (ac-component). The GSR signal is analyzed using the Ledalab toolbox software package in MATLAB (Amba et al., 2017). The phasic response describes the eccrine gland function through SCR. Sympathetic innervation causes an increase in heart and pulse rate with an increase in the number of SCR due to innervation of postganglionic sympathetic fibers into the eccrine sweat glands (Wilke et al., 2007). The SCR is extracted by performing continuous deconvolution analysis (CDA) within a response window (wrw) defined after the occurrence of the events. The parameters observed are as follows: CDA-nSCR- Number of significant SCR within the response window (wrw). CDA-latency - Response Latency of 1st significant SCR wrw.

459

 An Exploration of Influence of Duration on Physiological Effects of Asanas

CDA-ampsum- Sum of all SCR amplitudes. CDA-SCR- Average of Phasic Driver wrw. CDA-ISCR- Area (time integral) of phasic driver wrw. CDA-phasic Max- Maximum value of phasic activity wrw.

Results and Discussion Time Domain Results The time-domain measures for short term yoga show an elevated SDRR, RMSSD, HRMV with reduced heart rate and breath rate. However, long-term yoga causes an increase in heart rate as well as pulse rate by 6.68 percent and 7.81 percent respectively due to a significant reduction in breath rate. As the rise in heart rate is not significant, HRMV also reduces by 33.37 percent significantly than PRMV and BRMV which show a relatively minimal RSA. It is evident that at higher breathing frequency (> 6.7 bpm) the amplitude of RSA drops. The amplitude of RSA is measured in decibel (dB). The drop in magnitude of RSA is the roll-off rate (19.4 dB/decade) (Hirsch et al., 1981). The breathing frequency before and after Yogathon is 24.81 and 21.34 respectively which signifies that the participants have lower RSA. The bar chart in Figure 2, Figure 3, Figure 4 indicates the relative changes in time-domain parameters. Figure 2. Heart rate, Pulse Rate, Breath rate for RRI, PPI and BBI. RRiy, PPiy, BBiy- represents parameters after yogathon

Table 1. accounts for changes in time domain measures after Yogathon

Frequency Domain Results The box-plot in Figure 5, Figure 6, Figure 7 indicate spectrogram results before and after long-term yogaasana for RRI, PPI and BBI respectively. VLF increases by 22.24 percent, 16.83 percent, 32.25 percent respectively for RRI, PPI and BBI showing overall enhancement in cardio-respiratory functionality. LF percentage decreases for RRI, PPI and BBI with minimal changes in HF.

460

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Figure 3. Rate of Maximum Variation(RMV) for RRI, PPI and BBI. RRiy, PPiy, BBiy- represents parameters after yogathon

The reduction in LF-power resembles lower sympathetic flow from ANS and the HF-power change is minimal due to reduced RSA and it sometimes reflects cardiac vagal tone (Grossman et al., 2007). Table 2 shows the percentage change in spectral power after Yogathon.

Figure 4. Average value of RMSSD for RRI, PPI and BBI. RRiy, PPiy, BBiy- represents parameters after yogathon

Non-linear Measure Results Poincare parameters such as SD1 and SD2 for short and long term variability are used to assess HRV by fitting ellipses along the LOE. SD1 and SD2 increase by 23.83 percent and 42.13 percent respectively after yogathon. The Poincare’s plot of seven subjects before and after yogathon is shown in Figure 8 and Figure 9 respectively. The lagged Poincare plots are shown in Figure 10 and Figure 11 considering lag (successive difference in the signal interval (m)) up to 16 before yogathon and after yogathon respectively.

461

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Table 1. BY - Before Yogathon and AY - After Yogathon. The ‘+’ and ‘-’ sign is for percentage increase and decrease in average value in time domain parameters of ECG, PPG and Breath Cycles Parameters

BY

Percentage Change

AY

Heart Rate (bpm)

86.3342

92.1021

6.68

Pulse Rate (bpm)

85.5103

92.1936

7.81

Breath Rate (bpm)

24.8108

21.3459

-13.96

hrmv (bpm)

211.8721

141.1622

-33.37

prmv (bpm)

60.274

59.6948

-0.96

brmv (bpm)

20.2149

19.633

-2.87

sdrri (sec)

0.0456

0.0648

42.1

sdppi (sec)

0.0769

0.0672

-12.61

sdbbi (sec)

0.8828

0.8192

-7.2

rmssd (sec)

0.0436

0.0539

23.62

rmssd-p (sec)

0.0793

0.065

-18.03

rmssd-b (sec)

1.1982

1.0022

-16.35

Figure 5. Box plot for Area under VLF (aVLF), LF (aLF), HF (aHF), VHF (aVHF) of RRI before and after long-term yoga (represented with aVLFy, aLFy, aHFy, aVHFy).

In yogathon the range of change in SD1 and SD2 with lag (2-16) happens in a broad range of 25 ms for SD1 and SD2 before yogathon and 12 ms after yogathon, averaged around 7 subjects. Therefore, yogathon not only elevates the variability but also makes the rate change stable. It reflects a reduction in deviation of RR-interval signifying steadiness in heart rate due to continuous practice. In other words, the systemic flow resistance becomes uniform. Table 3 shows the changes in the percentage of SD1 and SD2 of RRI, PPI, BBI after Yogathon.

462

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Figure 6. Box plot for Area under VLF (aVLF), LF (aLF), HF (aHF), VHF (aVHF) of RRI before and after long-term yoga (represented with aVLFy, aLFy, aHFy, aVHFy).

Figure 7. Box plot for Area under VLF (aVLF), LF (aLF), HF (aHF), VHF (aVHF) of BBI before and after long-term yoga (represented with aVLFy, aLFy, aHFy, aVHFy).

Table 2. The ‘+’ and ‘-’ sign is for percentage increase and decrease of change in frequency domain parameters of ECG, PPG, and Breath signal before and after Yogathon Percentage Change in Power

RRI

PPI

BBI

VLF

22.24

16.83

32.25

LF

-14.56

-26.3

-17.97

HF

3.3

-5.03

-3.13

463

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Figure 8. Poincare plot for seven subjects before yogathon. The yellow dots show corresponding values of SD1 and SD2

Figure 9. Poincare plot for seven subjects after yogathon. The yellow dots show corresponding values of SD1 and SD2

Galvanic Skin Response Results Galvanic skin response quantifies the emotional state of a person, a lesser GSR response results due to a more depressed state of mind (Greenfield et al., 1963). The average skin conductance decreases by 65.98 percent after yogathon. This signifies a relaxed state of mind. During active state the number of skin conductance responses (nSCR) tends to be more. However nSCR does not change significantly. This defines the property of an asana which is ‘sthiram’ and ‘sukham’. That is an asana itself provides rest without much expenditure of metabolic energy. Table 4 shows the percentage change in GSR parameters after yogathon.

464

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Figure 10. Poincare Plot for Lags (2- 16) for a participant before yogathon

Figure 11. Poincare Plot for Lags (2- 16) for a participant after yogathon

Table 3. The ‘+’ and ‘-’ sign is for percentage increase and decrease of change in Variability of ECG, PPG and Breath signal before and after Yogathon Percentage Change in Variability

RRI

PPI

BBI

SD1

23.83

-18.03

-16.37

SD2

42.13

-11.08

-4.32

465

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Table 4. The ‘+’ and ‘-’ sign is for percentage increase and decrease of change in GSR-parameters before and after Yogathon. GSR-Parameters

BY

AY

Percentage change

CDA-nSCR

101

104.5

3.46

CDA-latency (Sec)

1.15

0.5

-56.72

CDA-ampsum (µS)

342.78

140.59

-58.98

CDA-SCR (µS)

39.204

13.33

-65.98

CDA-ISCR (µS.Sec)

1.74E+03

631.8723

-63.68

CDA-phasic (µS)

29.72

14.46

-51.35

Comparison of Yogathon with Short Term Yoga Comparison of Time Domain Results RSA drops down with the magnitude of physical work (Grossman et al., 2004), as yogathon involves endured physical formations, hence responsible for a lesser RSA and reduced HRMV than that of short term yoga practices (6.8 percent reduction). In this case, the average breath rate reduces from 24.8 to 21.3 bpm which in turn elevates the heart rate by 6.68 percent. The SDRR and RMSSD show a higher variability both in case of short and yogathon practice.

Comparison of Frequency Domain Results The literature study shows an increased VLF power, reduced LF power and increased HF power after short term yoga-asana practice. However, each asana causes differences in the power under various frequency zones (Pitale et al., 2014). Following results from (Pitale et al., 2014) show relative changes in LF, HF power during 7 different yoga asanas. The literature shows a similar pattern for both yogathon and short term yoga-asana in case of VLF and LF, however the magnitude of percentage change of VLF is higher. It contributes to the span of a session which is more for yogathon.

Comparison of Non Linear Measure and GSR Results The area of the ellipse is the measure of total variability, Pitale et al. (2014) shows the relative difference between Standing-pose and Seated-forward-bend-pose. However, the SD1 and SD2 both enhance after short term yoga practice (Pushpanathan et al., 2014). The difference between short and long-term practice comes with a variation of SD1 and SD2 as Lag (successive difference in the signal interval (m)) increases. For short term practice, there is a stepwise enhancement in SD1 with stepwise reduction of SD2 as lag increases from 1 to 10 in the case of yogaasana session of 2 hours (Hernando et al., 2019). However, in yogathon, the change in SD1 and SD2 with lag (1-16) happens in a narrow range of 12 ms for SDs. Both the short and yogathon shows an elevated skin impedance or reduced skin conductance reflecting a relaxed state of mind. All the relative difference between short term yoga-asana and yogathon considering HRV and GSR is tabulated below in Table 5:

466

 An Exploration of Influence of Duration on Physiological Effects of Asanas

Table 5. Comparison of Time Domain Measures, Frequency Domain Measures, Poincare parameters (SD1, SD2) and average SCR between short and long-term Yoga practice. s-significance, ns- not significant Time Domain Measures

Short Duration (From Literature)

Yogathon

Heart Rate (bpm)

Decreases (s) by 8.55% (Markil et al., 2012)

Increases (ns) by 6.68%

Breath Rate (bpm)

Decreases (s) by 51.85% from maximum during Ploughpose to minimum during Corpse-pose (Ray et al., 2011). Decreases (s) by 24.6% during 15 min chair yoga session (Melville et al., 2012)

Decreases (s) by 13.96% at the end of session.

HRMV (bpm)

For female participants HRMV reduced by 6.8% (s) from warm up to cool down phase (Tsopanidou et al., 2018). No results for PRMV and BRMV.

Decreases (s) by 33.37% in HRMV. (There is no significant change BRMV and PRMV)

SDRRI (sec)

Increases by 25.32% (s) between pre and post intervention of Yoga for a month (Vinay et al., 2016) for RRI

Increases by 42.1% (s) for RRI and decreases by 12.6% and 7.2% for PPI and BBI respectively.

RMSSD (sec)

Increases by 16.36% (s) between pre and post intervention of Yoga for a month (Vinay et al., 2016)

Increases by 23.62% (s) for RRI and decreases by 18.03% and 16.35% for PPI and BBI respectively.

VLF power (ms2)

Increases (s) by 8.86% between pre and post intervention of Yoga for a month (Vinay et al., 2016), for RRI.

Increases (s) by 22.24%, 16.83%, 32.25%for RRI, PPI, BBI respectively.

LF power (ms2)

Decreases (s) by 22.64% between pre and post intervention of Yoga for a month (Vinay et al., 2016), decreases during 1 hour yoga for 12 weeks (Krishna et al., 2014). Lesser low-frequency component (LF) (5.52%, P