203
AUGMENTED SELF-REGULATION A DISSERTATION SUBMITTED TO THE SCHOOL OF EDUCATION AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Neema Moraveji June 2012

Augmented Self-Regulation (Moraveji, 2012)

Embed Size (px)

DESCRIPTION

Neema Moraveji's Ph.D. Dissertation at Stanford University, spring 2012.

Citation preview

Page 1: Augmented Self-Regulation (Moraveji, 2012)

AUGMENTED SELF-REGULATION

A DISSERTATION

SUBMITTED TO THE SCHOOL OF EDUCATION

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Neema Moraveji

June 2012

Page 2: Augmented Self-Regulation (Moraveji, 2012)

ii

Page 3: Augmented Self-Regulation (Moraveji, 2012)

iii

Page 4: Augmented Self-Regulation (Moraveji, 2012)

iv

ABSTRACT

Research in interactive technology most often enables humans to enact more,

better, or new behaviors or to gain greater or more efficient insight into our environment.

This dissertation develops the notion that technology can also help us develop greater

self-awareness. Self-awareness, made tangible and observable by physiological self-

regulatory processes, rests at the core of the human experience by supporting

comprehension, actions, and intentions in the world around us. Without it, our health and

behavior are victims of demands from our sociocultural environments or of flippant or

self-defeating patterns of thoughts.

This dissertation explores and evaluates methods of using wearable sensors and

interactive feedback to augment human self-regulation, specifically respiratory self-

regulation, primarily during information work but also in mobile contexts. This is a

potentially powerful means of not only influencing health and behavior but also

developing an inner sense of wellness independent of one’s physiological, cognitive, or

affective state.

The over-arching contribution of this dissertation is to demonstrate the view that

computers will have a more direct and pervasive impact on human psychophysiology

than is currently practiced. Rather than strictly adapting to human physiological and

affective state, future machines will explicitly induce changes in psychophysiological

state to amplify their user’s innate self-regulatory ability and skill. In this way, machines

of the future will help humans be ‘more human’ rather than simply adding or enhancing

existing human abilities.

Page 5: Augmented Self-Regulation (Moraveji, 2012)

v

ACKNOWLEDGEMENTS

Maryam: for being the soul I choose to merge with. Mommy & Nick: for the

courage to open your heart when it’s easiest to close. Babai & Mamani: for creating our

collective soul and blueprint. Newsha, Jim, & Azilu: for standing up for me. Daddy &

Sorour: for teaching me the power of discipline. Roya, Hameed, Vala, & Ala: for

fearlessness. Jason: for big brotherhood. Fared: for little brotherhood. Sheida, Roxanne,

& Bijan: for leading the way West. Malek, Jaleh, Nanaz, Tannaz, Chris, Nikki, & Darius:

for going first and making fun a priority.

My academic and professional mentors: Roy: for blowing me away, with love.

BJ: for breaking all the rules. Jeff: for rigor. Jelena: for pushing me (and your baby!).

Sep: for setting a high bar, with compassion. Dan R.: for cherishing reflection. Merrie &

Dan: for perpetual playful inventiveness. Mary: for making happiness come first. Paul &

Hamet: for trusting.

My dear Stanford collaborators, who led the way in many instances: Kenneth

Jung, Charlton Soesanto, Abhishek Sharma, Jimmy Chion, Ben Olson, Mahmoud Saadat,

Mohammad Hekmat, Maryam Rasaee, Takehiro Hagiwara, Poorna Kirshnamoorthy, Truc

Nguyen, Jim Zheng, Huyen Tran, Yaser Khalighi, Jonathan Palley, Stephanie Habif,

Emily Goligoski, and Karen Everett.

Thank you to all those who worked to bring breath awareness to light.

Page 6: Augmented Self-Regulation (Moraveji, 2012)

vi

TABLE OF CONTENTS

ABSTRACT IV!

ACKNOWLEDGEMENTS .................................................................................. V!

TABLE OF CONTENTS .................................................................................... VI!

LIST OF TABLES ........................................................................................... XIII!

LIST OF FIGURES ......................................................................................... XIV!

CHAPTER 1!INTRODUCTION ........................................................................... 1!

1.1! The problem ................................................................................................. 2!

1.2! Proposed solution ......................................................................................... 3!

1.2.1! Thesis statement ......................................................................................... 3!

1.3! Research challenges ...................................................................................... 3!

1.4! Summary of findings .................................................................................... 5!

CHAPTER 2!RELATED WORK .......................................................................... 9!

2.1! Stress and self-regulation ............................................................................. 9!

2.1.1! Effects of stress on learning, performance, and behavior ....................... 12!

2.1.2! Self-regulation: therapeutic and developmental ...................................... 15!

2.2! Augmenting self-regulation with breath modification ............................... 19!

2.2.1! Regulating respiration to mitigate the stress response ............................ 24!

2.3! Technology-mediated respiratory self-regulation ...................................... 29!

2.3.1! Social influence on physiological behavior .............................................. 33!

2.4! Summary and implications for current research ........................................ 35!

Page 7: Augmented Self-Regulation (Moraveji, 2012)

vii

DESIGN AND VALIDATION OF A NON-CLINICAL AMBULATORY RESPIRATION SENSOR .............................................................................. 37!

2.4.1! Sensor design ........................................................................................... 39!

2.4.2! Industrial design evolution ....................................................................... 41!

2.4.2.1! Iteration 1: Wired USB ......................................................................... 41!

2.4.2.2! Iteration 2: Wireless USB ..................................................................... 42!

2.4.2.3! Iteration 3: Wireless Bluetooth ............................................................ 42!

2.5! Sensor validation ........................................................................................ 44!

2.6! Conclusion .................................................................................................. 48!

CHAPTER 3!PERIPHERAL PACED RESPIRATION: INFLUENCING RESPIRATORY PATTERNS DURING INFORMATION WORK ............... 49!

3.1! A peripheral paced respiration interface .................................................... 50!

3.1.1! Wizard-of-Oz prototype ........................................................................... 50!

3.1.2! User interface ............................................................................................ 52!

3.1.3! Pacing respiration peripherally ................................................................ 53!

3.2! Study ........................................................................................................... 55!

3.2.1! Procedure ................................................................................................. 55!

3.3! Results ......................................................................................................... 56!

3.4! Discussion ................................................................................................... 59!

3.5! Conclusion .................................................................................................. 60!

CHAPTER 4!BREATHCAST: A STUDY OF SOCIAL INFLUENCE ON BREATH MODIFICATION .......................................................................... 62!

4.1! Study ........................................................................................................... 64!

4.1.1! Participants ............................................................................................... 64!

4.1.2! Procedure ................................................................................................. 65!

Page 8: Augmented Self-Regulation (Moraveji, 2012)

viii

4.2! Results and discussion ................................................................................ 66!

4.3! Conclusion .................................................................................................. 70!

CHAPTER 5!BREATHAWARE: CONTINUOUS INFLUENCE OF SELF-REGULATION DURING INFORMATION WORK ..................................... 71!

5.1! Design principles ........................................................................................ 72!

5.1.1! Accommodate different levels of attention .............................................. 73!

5.1.2! Sustain motivation .................................................................................... 74!

5.1.3! Demonstrate desired patterns of breath .................................................. 75!

5.1.4! Personalize feedback ................................................................................ 76!

5.1.5! Reinforce the relationship between breath and body .............................. 77!

5.1.6! Avoid exasperating stress with negative feedback ................................... 77!

5.1.7! Develop awareness at different timescales ............................................... 78!

5.1.8! Encourage internal self-assessments ........................................................ 79!

5.1.9! Consider secondary components of respiration ...................................... 80!

5.1.10!Protect the privacy of breath .................................................................... 80!

5.2! Prototype design ......................................................................................... 81!

5.2.1! Client ........................................................................................................ 81!

5.2.2! Social network .......................................................................................... 82!

5.3! Interaction design ....................................................................................... 82!

5.3.1! Breath rate (immediate) – DP3, DP7 ........................................................ 83!

5.3.2! Breath rate (daily) – DP5, DP7 .................................................................. 84!

5.3.3! Breath rate (immediate but relative) – DP3, DP7 ..................................... 85!

5.3.4! Breath rate (longitudinal) – DP7 ............................................................... 86!

5.3.5! Calm points – DP2 .................................................................................... 86!

Page 9: Augmented Self-Regulation (Moraveji, 2012)

ix

5.3.6! Daily milestones – DP2 ............................................................................. 87!

5.3.7! Encouraging messages – DP2 ................................................................... 88!

5.3.8! Check-in – DP8 ......................................................................................... 89!

5.3.9! Cautionary feedback – DP6 ...................................................................... 90!

5.3.10!Activity stream – DP2, DP10 .................................................................... 90!

5.3.11!Buddy list – DP10 ..................................................................................... 91!

5.3.12!Re-record and modify baseline – DP3, DP4 ............................................ 92!

5.4! Test deployment ......................................................................................... 92!

5.4.1! Results and Feedback ............................................................................... 93!

5.5! Discussion and implications for research ................................................... 95!

5.6! Conclusion .................................................................................................. 97!

CHAPTER 6!BREATHTRAY: CONTINUOUS RESPIRATORY FEEDBACK AND ITS EFFECTS ON COGNITIVE PERFORMANCE ............................. 99!

6.1! BreathTray design ..................................................................................... 100!

6.2! Study design ............................................................................................. 101!

6.2.1! Participants ............................................................................................. 101!

6.2.2! Serial sevens ........................................................................................... 102!

6.2.3! Problem-solving with auditory distractors ............................................. 103!

6.2.4! Procedure ............................................................................................... 104!

6.3! Results ....................................................................................................... 106!

6.3.1! BreathTray impact on breath regulation ................................................ 107!

6.3.2! Magnify or persist? .................................................................................. 108!

6.3.3! Impact on cognitive performance .......................................................... 109!

6.3.4! Qualitative feedback ............................................................................... 109!

Page 10: Augmented Self-Regulation (Moraveji, 2012)

x

6.4! Discussion ................................................................................................. 110!

6.4.1! Study limitations .................................................................................... 111!

6.5! Conclusion ................................................................................................ 112!

CHAPTER 7!BREATHWEAR: AMBULATORY INFLUENCE ON RESPIRATORY PATTERNS ...................................................................... 113!

7.1! Design opportunities and challenges ....................................................... 116!

7.1.1! Continuous state influence ..................................................................... 116!

7.1.2! Context-awareness .................................................................................. 117!

7.1.3! Sleep ....................................................................................................... 117!

7.1.4! Social ...................................................................................................... 117!

7.1.5! Annoyance .............................................................................................. 118!

7.1.6! Evolving user goals ................................................................................. 118!

7.1.7! Over-dependence ................................................................................... 118!

7.1.8! Inaccuracy ............................................................................................... 119!

7.2! Interaction design goals ............................................................................ 119!

7.2.1! Monitor ................................................................................................... 119!

7.2.2! Influence ................................................................................................. 120!

7.2.3! Customize ............................................................................................... 120!

7.3! Design iteration 1 ...................................................................................... 120!

7.5! Design iteration 2 ...................................................................................... 123!

7.5.1! Recent activity indicator ......................................................................... 124!

7.5.2! Relative breath rate ................................................................................. 125!

7.5.3! Breathbelt feedback ................................................................................ 125!

7.5.4! Additional settings .................................................................................. 125!

Page 11: Augmented Self-Regulation (Moraveji, 2012)

xi

7.6! study .......................................................................................................... 126!

7.6.1! Results .................................................................................................... 127!

7.6.2! User feedback ......................................................................................... 135!

7.7! Discussion ................................................................................................. 137!

7.8! Conclusion ................................................................................................ 139!

CHAPTER 8!CONCLUSIONS AND FUTURE WORK .................................... 140!

8.1! Summary of findings ................................................................................ 140!

8.2! Limitations ................................................................................................ 142!

8.2.1! Breath rate alone is not illustrative of autonomic activity ...................... 142!

8.2.2! The evaluator effect ................................................................................ 143!

8.2.3! Competition confound ........................................................................... 143!

8.2.4! Single data source .................................................................................. 144!

8.3! Supplementary contributions and implications ....................................... 144!

8.3.1! Autonomic interaction design ................................................................ 144!

8.3.2! ASR and incentivizing self-regulation ................................................... 145!

8.3.3! Techniques for incentivizing self-regulation ......................................... 146!

8.3.4! ASR, being, and doing ........................................................................... 147!

8.3.5! ASR and the purposeful evolution of human society ............................ 148!

8.4! near-term future work ............................................................................... 150!

8.5! Concluding remarks .................................................................................. 150!

BIBLIOGRAPHY .............................................................................................. 152!

APPENDIX A! PPR STUDY POST-SURVEY ................................................ 172!

APPENDIX B! BREATHCAST STUDY POST-SURVEY .............................. 175!

Page 12: Augmented Self-Regulation (Moraveji, 2012)

xii

APPENDIX C! BREATHTRAY STUDY MATERIALS .................................. 177!

C.1! Pre-survey ................................................................................................. 177!

C.2! Video motivating breath awareness and regulation ................................. 177!

C.3! Textual Explanation of BreathTray ........................................................... 179!

C.4! Text explanation of Serial Sevens task ..................................................... 179!

C.5! Text explanation of Problem-Solving with Audio Distractors task ......... 180!

C.6! Post-survey ................................................................................................ 180!

APPENDIX D! BREATHWEAR INSTRUCTIONS AND FEEDBACK FORM 182!

Page 13: Augmented Self-Regulation (Moraveji, 2012)

xiii

LIST OF TABLES

Table 1.1: An overview of the studies and contributions in this dissertation. .... 6!

Table 2.1: Lay descriptions of the key concepts motivating breath regulation as a means of regulating psychophysiological state. Not meant to be comprehensive. .............................................................................. 28!

Table 3.1: Known methods of sensing respiration. ............................................ 38!

Table 6.1: 10 design principles for interactive systems aiming to influence respiratory self-regulation. ................................................................. 73!

Table 7.1: Mean (and standard deviation) breath rates across both BreathTray and NoBreathTray conditions in each task and across both tasks together. .......................................................................... 106!

Page 14: Augmented Self-Regulation (Moraveji, 2012)

xiv

LIST OF FIGURES

Figure 2.1: An abbreviated diagram of the complementary functions of the sympathetic and parasympathetic components of the nervous system (PrecisionNutrition.com, 2012). The stress response is characterized by sympathetic activation and the relaxation response by parasympathetic. ............................................................ 12!

Figure 2.2:A diagram showing McEwen’s theory of allostatic load as emerging from interactions between environmental stressors, individual differences, behavioral responses, and physiological responses (McEwen, Gianaros, 2011). ................................................ 14!

Figure 3.1: The initial breathbelt sensor. (Top) The adjustable sensor band with the Arduino Uno board. (Bottom) Close-up of the stretch sensor held in place by two clips. ....................................................... 40!

Figure 3.2: The respiration sensor’s raw signal (red) is filtered (blue) and then peaks (black) are detected using well-studied signal processing techniques. The Y-axis refers to raw sensor values (not normalized). ........................................................................................ 41!

Figure 3.3: Wireless USB version of the original breathbelt, using paired XBee wireless communication widgets and a Lilypad Arduino (the circular PCB). The black plastic case on the top-most image holds 2 AAA batteries. ................................................................................. 42!

Figure 3.4: The strain gauge (black) has two hooks (top), which was modified to include buttons to snap in and out of the sensor (bottom). This was necessary because the strain gauge occasionally breaks and must be replaced. ............................................................................... 43!

Figure 3.5: The most recent breathbelt design includes 3 components: (a) adjustable band, (b) strain gauge, and (c) microprocessor Bluetooth 4 transceiver, allowing it to communicate continuously with a mobile phone. .......................................................................... 44!

Figure 3.6: The PASCO respiration sensor used to validate our strain gauge sensor. It uses a gas pressure sensor to measure how air pressure in the belt changes as the wearer breathes. ....................................... 45!

Page 15: Augmented Self-Regulation (Moraveji, 2012)

xv

Figure 3.7: X-axis is Time in seconds and the Y-axis is normalized sensor values. A 5-minute excerpt of our thoracic strain gauge breathbelt (red) compared with the PASCO sensor (blue) shows their similarity, with some noise. Each peak is the apex of an inhalation. 46!

Figure 3.8: X-axis is Time in seconds. A comparison between the sensor gauge-based breathbelt and the commercial PASCO sensor that uses highly sensitive air pressure fluctuations to measure breath rate. ..................................................................................................... 47!

Figure 4.1: The user interface of the web-based Wizard-of-Oz interface that study administrators would use to control the visualization of the desktop-based feedback of a user in real-time. The panel included elements that were not yet in use (e.g. breath regularity). ................. 51!

Figure 4.2: The user interface of the client was used by the study administrator to select the current peripheral pacing method. ......... 52!

Figure 4.3: The peripheral paced respiration feedback used an animated, semi-translucent grey bar stretching across the screen. Vertical arrows on the left indicate the full range of motion. ......................... 54!

Figure 4.4: (Top) Mean breath rate for the No Feedback and PPR conditions with standard error bars. (Bottom) Mean breath rate during the PPR condition when PPR was on and off. ......................................... 57!

Figure 4.5: Breath rate for one participant in both no feedback (top) and PPR (bottom) conditions. Bold (orange) areas indicate where PPR occurred .............................................................................................. 58!

Figure 5.1: Breathcast works by intermittently animating a semi-transparent bar across the bottom third of the user’s screen. The inset shows how profile icons of other Breathcast users are discreetly displayed on the bar. The vertical arrow on the lower right illustrates the range of bar movement. In asynchronous mode, the bar is blue to aid differentiation. ........................................................ 63!

Figure 5.2: Mean breath rates for participants in each condition. ..................... 67!

Figure 5.3: Mean breath rate for each condition. BL=baseline, WBL=working baseline, A=asynchronous, S=synchronous. ............ 68!

Figure 5.4: Breath rate of a user with a resting rate of 19.5bpm. PPR occurrences are orange. The working baseline condition (top) saw

Page 16: Augmented Self-Regulation (Moraveji, 2012)

xvi

the breath rate climb upwards. With synchronous feedback (middle), it decreases noticeably during PPR. Asynchronous feedback (bottom) saw rates drop little and continue to climb overall. ................................................................................................ 69!

Figure 6.1: The BreathTray and its drop-down menu in two states: sensor disconnected (left) and connected (right). The user’s current breath rate is displayed on the user’s system tray. ............................. 83!

Figure 6.2: ‘Today’s Highs and Lows’ shows three desktop screenshots from each category with relevant metadata (breath rate, time of day). Here, the user’s breath rate was highest when working in email and on a presentation. It was lowest when viewing their calendar and viewing a document. .................................................................... 85!

Figure 6.3: The user’s Breath.fm profile for an imaginary user, ‘KKP’. The top shows their overall data including name, last activity update, total calm points accumulated, mean baseline, and mean BPM. The area below shows their activity stream with event notifications updated in real-time: positive and negative reinforcement messages and milestone images. ........................................................ 86!

Figure 6.4: Calm point milestones. The desktop of a user who achieved the 80-point milestone. The inspiring images are always randomized as an attempt to create anticipation for the different milestones. ..... 87!

Figure 6.5: The prototype system showing two types of notifications in the lower-right corner of the screen: (a) Left, positive feedback gives the user a congratulatory message and a duration for which they were breathing relatively calmly. (b) Right, a cautionary message tells users how long they have been breathing relatively fast. ........... 89!

Figure 6.6: When other users have recently logged in and had data sent to the web repository, the drop-down menu also doubles as the location of the buddy list. Usernames, current point values, and last recorded breath rates are displayed. ............................................ 92!

Figure 7.1: The BreathTray shows 4 components: calm points, breath rate, percent relative to resting rate, and whether they are above (red) or below (blue) their resting breath rate. ......................................... 100!

Figure 7.2: The ‘Serial Sevens’ task adapted to a web-based interface. A starting number was shown (top) and numbers disappeared when participants typed and pressed Enter (bottom). ............................... 102!

Page 17: Augmented Self-Regulation (Moraveji, 2012)

xvii

Figure 7.3: The math problem in the ‘Problem-Solving with Audio Distractors’ task. Users were to choose the correct expression using the radio buttons and press the “answer” button. ................. 103!

Figure 7.4: The study setup with the USB-connected wearable respiration sensor (left), computer, headphones, and external mouse. The web browser was maximized to fill the screen for all conditions. ........... 104!

Figure 7.5: Mean breath rates in each task, split by condition and also aggregated across all condition. The difference between BreathTray and No BreathTray is significantly different in the Serial Sevens task. ............................................................................ 107!

Figure 8.1: The initial prototype. Dashboard (left) shows real-time feedback and Settings (right) allows the user to make changes to their breath rate baseline and to manually send data back to the research team via email. ................................................................... 121!

Figure 8.2: Push notifications on an iOS5-based mobile phone based on initial prototype design. Each notification has a type (positive or negative/cautionary) and duration that the system detected the user was in that state. For example, the user had been breathing above their resting rate for 15.2min at left, at or below their resting rate for 12.7min on right. ..................................................... 122!

Figure 8.3: The second iteration of the Breathwear client interface, which includes a recent activity indicator (left, top, in green) and additional configuration options in the Settings screen (not shown). The center image shows when the user is hovering around their baseline (hence the yellow) and the sensor is connected (hence the ‘Time Connected’ indicator). During a high breath rate state, the band is red (not shown). The baseline here is set to 15bpm (the default). ................................................................ 124!

Figure 8.4: Distribution of breath rates of each participant. X-axis is “Breath rate in Breaths per Min”. Y-axis is ‘Frequency in Seconds’. Clockwise from top-left, users 1, 2, 4, 5, and 3. .............. 128!

Figure 8.5: Frequency and type of push notifications received by participant. ‘Calm’ and ‘Zen’ are two types of positive notifications. This graph shows that there was no discernable trend around type or frequency of push notifications but that one can characterize an individual’s respiratory patterns to some degree using this visualization. ......................................................... 130!

Page 18: Augmented Self-Regulation (Moraveji, 2012)

xviii

Figure 8.6: Calm points per hour granted to each study participant. This graph shows how User #2, who had a great deal of meditation experience, was rewarded a great deal more calm points per hour than the other participants. .............................................................. 131!

Figure 8.7: A line graph produced by one participant with data she labeled herself, “meditating”, “surfing the web”, and “reading”. ................. 132!

Figure 8.8: A line graph showing the relationship between a user’s breath rate punctuated by the different types of push notifications (green=calm, blue=zen, red=stress). The line indicating the user’s breath rate is by default gray and then colored according to the duration of the state detected by the subsequent push notification.134!

Page 19: Augmented Self-Regulation (Moraveji, 2012)

1

CHAPTER 1 INTRODUCTION

People across a range of ages and occupations engaged in computer-based work

or study frequently suffer from task interruptions (Iqbal & Horvitz, 2007), difficulty

sustaining attention (Holt & Andrews, 1989), and chronic stress (Mark, Gudith, Klocke,

2008). This can seem like merely an inconvenience to productivity but consider that the

impact of high stress reactivity extends to decreasing cognitive performance (Luine,

Villages, Martinex, McEwen, 1994; McEwen & Sapolsky, 1995), behavioral problems in

children (Obradovi!, Boyce, 2012), and ultimately damage to the brain (Sapolsky, 1996).

The problem of susceptibility to chronic stress can be approached as a problem of

inadequate self-regulation rather than only exasperating environmental stressors. As such,

this dissertation examines how technology can augment one’s self-regulation processes

rather than analyzing exactly how different environmental stressors occur and could be

mitigated. Specifically, the research agenda is to augment respiratory self-regulation, a

common technique used to help one learn to self-modulate one’s own

psychophysiological state. The goal of this agenda is to usher in new tools, technique,

and design examples for amplifying, augmenting, and ultimately strengthening one’s

Page 20: Augmented Self-Regulation (Moraveji, 2012)

2

innate self-regulatory system. Our approach leverages the recent emergence and unique

affordances of wearable sensors and personal mobile devices that make continuous

monitoring and feedback possible. Thus, the research here represents the first known

study of digital augmentation of continuous respiratory behavior.

To date, technology-mediated respiration influence has required the user to stop

and disengage from their current task and focus full attention on modifying their

respiratory state. This is often done during clinical sessions separate from the work

context creating the stress. However, humans can “regulate their respiration rates in a

relatively short time period” (Ley, 1999), making breath regulation a viable treatment for

sporadic and subtle stressors such as those that may be encountered in everyday settings

such as multi-tasking on a laptop computer, transitioning between tasks, and face-to-face

meetings. Indeed, continuous influence of respiratory patterns represents a fundamentally

different approach than existing modal solutions, which are subject to inconveniences and

compliance issues. Building off of research that studies how tools augment our

intellectual capacities (Pea, 1985; Pea, Gomez, 1992; Pea, 2004), this approach relies on

an understanding of psychophysiology, wearable sensors, interaction design, and

persuasion (Fogg, 2002; Cialdini, 2008).

1.1 THE PROBLEM

Human psychophysiological self-regulation is a crucial component to modern life

given the numerous demands on our attention and arousal. Such consistent arousal causes

mild but chronic stress (McEwen & Sapolsky, 1995), which can lead to poor academic

performance (Cassady & Johnson, 2001), burnout (Etzion, 1984), and even loss of brain

plasticity (McEwen & Gianaros, 2010). While tools grow to extend and amplify our

Page 21: Augmented Self-Regulation (Moraveji, 2012)

3

cognitive and communicative abilities (Pea, 1985), less research supports our innate self-

regulative abilities, particularly in the domain of stress. Learning to self-regulate in the

face of stressors is an invaluable tool that is only rarely given due focus in traditional

education systems (Pope, 2001; Obradovic, Boyce, 2012).

1.2 PROPOSED SOLUTION

Many non-technological techniques are now common to help people monitor and

adjust to a flood of demands (e.g., conscious breathing, yoga, nature retreats). Many of

these incorporate respiration monitoring and modification as a means of modulating

arousal and reducing tension. They often hinge upon one’s ability to disengage from their

task, monitor their own respiration, and adjust it to a subjectively appropriate pattern.

Both (1) monitoring and (2) pacing one’s respiration are viable places where technology

can augment innate ability. Crucially, both monitoring and pacing can in principle be

done in parallel to existing information work (something we will investigate in Chapter

6), reducing problems of compliance and disruption that emerge when requiring one to

stop a task. This dissertation examines interactive techniques to influence respiratory

self-regulation during existing tasks. Our focus begins with influence breathing during

stationary work at a laptop computer and is then extended to mobile settings.

1.2.1 THESIS STATEMENT

Technological tools can be an effective means of influencing respiratory change and self-regulation during, and without negatively impacting, cognitive work.

1.3 RESEARCH CHALLENGES

There are several research challenges involved in demonstrating the above claim.

First, how can one’s respiration be influenced during meaningful computer work without

Page 22: Augmented Self-Regulation (Moraveji, 2012)

4

significantly distracting the user given normal patterns of use (i.e., without requiring the

use of headphones or distracting nearby workers), while ensuring the user knows the end

points of both inhalation and exhalation?

Second, given the real-time nature of physiological behavior, what might be the

role of social feedback on motivating peripheral respiratory change during computer

work?

Third, what is the design space of user interaction techniques to encourage the

development of one’s own self-regulation system as opposed to system-generated explicit

respiration-pacing (which, ostensibly, off-loads from the user to the system both

respiration monitoring and pacing)?

Fourth, are such feedback-based attempts to augment one’s self-regulation more

effective than simply being motivated to breathe calmly? Moreover, does the available

feedback take attention away from the user’s task to such a degree that cognitive

performance is compromised?

Fifth, how can such techniques at motivating change be expanded to apply across

contexts in a mobile setting?

These challenges can be expressed in the form of the following research

questions:

• Q1: Is it feasible to augment respiratory patterns of information workers as they

are engaged in meaningful information work?

• Q2: How does synchronous social feedback compare with asynchronous feedback

in peripheral paced respiration?

Page 23: Augmented Self-Regulation (Moraveji, 2012)

5

• Q3: How can sociotechnical systems be designed so as to motivate respiratory

change without prompting it explicitly?

• Q4: How does peripheral feedback influence respiration and does that feedback

negatively influence cognitive performance?

• Q5: How must these techniques be adapted to be effective in a mobile setting?

1.4 SUMMARY OF FINDINGS

Table 1.1 presents an overview of the studies in this dissertation in order to guide

the reader’s understanding and prepare them for more in-depth consideration ahead.

Page 24: Augmented Self-Regulation (Moraveji, 2012)

6

Chapter Study Name Contribution(s)

2 Literature review

Identification and motivation of respiration rate as an effective data source for monitoring and feedback to regulate user state.

3 Breathbelt Design and validation of a non-invasive, ambulatory, non-clinical respiration sensor to detect breath rate continuously.

4 Peripheral paced respiration (PPR)

(1) Desktop user interface design technique for visually pacing the user’s respiration in parallel to their existing tasks. (2) Showing that PPR reduces breath rate without subjectively disrupting users.

5 Breathcast (1) Application of synchronous feedback to physiological behavior (i.e., respiration). (2) Method of using social feedback atop the PPR bar magnifies the effect of the bar more when the feedback is thought to be synchronous as compared to asynchronous.

6 Breathaware (1) Identification of 10 design principles for systems that aim to influence respiratory behavior. (2) Design of several desktop techniques for influencing breath rate without using explicit pacing.

7 BreathTray (1) Desktop users self-determined when to use peripheral feedback to significantly regulate breath rate. (2) Peripheral real-time feedback changes breathing without compromising cognitive task performance. (3) The two findings above only apply during single tasks; users were unable to use the feedback while multi-tasking.

8 Breathwear (1) Iterative design of continuous monitoring and influence in a mobile context. (2) Analysis of results from a longitudinal study showing how the system describes respiration patterns over time and adapts notifications to individual behavior.

Table 1.1: An overview of the studies and contributions in this dissertation.

The studies in this dissertation rest upon a thorough review of the literature of the

psychophysiology of stress and calm (Chapter 2). It is in this review the reader is

introduced to both the illustrative and actionable properties of human respiration.

Scholars in medicine frequently have used, and continue to use, respiration to understand

the state of the body and mind as well as to influence it. Technology researchers and

Page 25: Augmented Self-Regulation (Moraveji, 2012)

7

designers have since used these properties to develop interventions that train or modify

one’s respiration to affect their emotional state, cognitive state, and general health.

The literature review goes forth to describe how technological interventions to

influence the breath have taken primarily a training approach, requiring users to

disengage from their existing tasks and focus their attention on the important task of

retraining their respiration patterns. This small requirement is large enough to keep

compliance down and continues to make breath modification uncommon outside

contemplative practices. As such, the first intervention study is one of influencing

respiration patterns in parallel to existing tasks at the computer during information work

(Chapter 3; Q1), requiring us to build and evaluate a non-invasive and wearable

respiration sensor (0).

Once assured it was possible to influence breath regulation as users engage in

information tasks, we turn our attention to the problem of motivating engagement with

such a system. For this we turn to social psychology, which has found that synchronous

physical behavior increases connectedness between parties. We applied this notion to a

physiological behavior – respiration – to see if synchronous feedback motivates self-

regulation more effectively than asynchronous behavior (Chapter 4; Q2).

Maintaining our focus on influencing respiration during information work, we

turn our attention to designing methods of influencing respiration without explicit pacing.

This requires considerable attention and is, like all interventions, subject to the novelty

effect. We designed and implemented a system that utilizes motivational cues, real-time

feedback, and operant conditioning to influence respiration in a desktop operating system

environment (Chapter 5; Q3).

Page 26: Augmented Self-Regulation (Moraveji, 2012)

8

Inserting physiological feedback directly into the operating system could have

adverse effects on productivity and performance, which is in opposition to our goal of

augmenting respiration while complementing existing tasks. Hence, we evaluated the

effect of peripheral, real-time feedback and motivational cues on cognitive performance

in a controlled task. The interesting results of the study shed light on when and how to

incorporate feedback and how it affects task performance for different types of tasks

(Chapter 6; Q4).

We took insights from designing for the desktop and applied them to a mobile

device to realize the goal of continuous respiration augmentation (Chapter 7; Q5). This

transition presents a new set of challenges and exciting opportunities for truly distributing

one’s self-regulatory processes into technological tools. Finally, Chapter 9 concludes the

dissertation with study limitations, directions for future work, and a discussion of

implications for the field.

Page 27: Augmented Self-Regulation (Moraveji, 2012)

9

CHAPTER 2 RELATED WORK

This chapter discusses the theories and experimental findings of research

conducted to better understand how stressful and calm states relate to human performance

and how interactive technologies can influence self-regulation based on breath

modification. The three sections of this chapter cover (2.1) the stress response and self-

regulation, (2.2) breath modification for self-regulation, and (2.3) technology-mediated

breath modification. A summary (2.4) concludes the chapter.

2.1 STRESS AND SELF-REGULATION

Stress is a causal or correlational factor to the six leading causes of death: heart

disease, cancer, lung ailments, accidents, cirrhosis of the liver, and suicide (APA, 2011).

Chronic stress negatively impacts one’s health, physiology, and ability to learn and

perform (Lupien et al., 2009; Sapolsky, 1996; Selye, 1998/1936; Sherwood, 2006).

First applied to humans from the world of structural engineering in 1936

(reproduced, Selye, 1998), the word ‘stress’ refers to distress, which differs from

eustress, ‘positive’ stress, or general stimulation (Selye, 1975). For years, Selye tried to

Page 28: Augmented Self-Regulation (Moraveji, 2012)

10

identify the specific environmental conditions (physical stressors such as lack of food and

extreme ambient temperatures) that gave rise to a specific physiological ‘stress response’

(e.g., enlargement of the adrenal glands, gastric ulceration, etc.). Ultimately, he decided

the determinants of stress were non-specific – because he could not identify the physical

determinants.

Psychologist John Mason (1968) later studied the psychological determinants that

could produce the stress response measured in hormone levels. He put participants in

different situations (e.g., parachuting, air-traffic control) and created a table summarizing

the difference in hormone levels before and after these situations. He found three

determinants characterize stressful stimuli: interpreted as unfamiliar, unpredictable, and a

lack of control. Later, Dickerson & Kenney (2002) identified a fourth determinant using

meta-analysis: feeling a social evaluative threat.

A stressor is a source causing stress (Selye, 1975), either subjective or

environmental. Stress “is a highly individual experience that does not depend on a

particular event such as time pressure, but rather, it depends on specific psychological

determinants that trigger a stress response” (Lupien et al., 2007). As such, stressors can

be absolute (objective and universal to all humans – e.g., physical danger) or implied

(subjective to the person in question – e.g., giving a presentation) (Lazarus, 1966).

When a situation is interpreted as stressful, a series of reactions begins in the

hypothalamus, causing the stress hormones glucocorticoids (cortisol in humans) and

catecholamines to be secreted by the kidneys. This gives rise to the so-called ‘fight-or-

flight’ response of the sympathetic nervous system characterized by increase in heart rate

and blood pressure. In effect, these hormones instruct the body to “increase the

Page 29: Augmented Self-Regulation (Moraveji, 2012)

11

availability of energy substrates in different parts of the body, and allow for optimal

adaptations to changing demands of the environment” (Lupien et al., 2007; p. 211).

While this is a natural means of responding to environmental demands, prolonged

secretion of these hormones is damaging to the body and brain including impairing tissue

repair, suppressing immune function (Lupien et al., 2007) and, ultimately, atrophy of the

hypothalamus (Sapolsky, 1996). The stress response is natural and has evolved to protect

the body. It is prolonged or repeated triggering or the inability to return to a neutral state

that finally causes damage as the body is preoccupied with preparing to react rather than

healing itself (see Figure 2.1).

The stress response is counteracted by the relaxation response, a coordinated

physiological response characterized by decreased arousal, diminished heart rate,

respiratory rate, and blood pressure, in association with a state of “well-being” (Benson,

1975). This response triggers parasympathetic activity while reducing sympathetic

activity (see Figure 2.1). The relaxation response helps mitigate the negative effects of

stress on the body and mind in an effort to return the body to homeostasis (Lang, 2000;

Syrjala et al., 1992; Vasterling, 1993).

Porges’ polyvagal theory (2001) provides a rich explanation of stress reactivity

than one based solely on cardiovascular responses. It states that the human stress

response system has evolved such that more primitive responses (e.g., feigning death,

rage) occur when more evolved ones (e.g., social communication, self-soothing) have

failed. This theory explains behavioral, not only physiological, responses to stress. It aims

to describe the structure and function of the vagus nerve, which influences the heart and

other organs during parasympathetic activation.

Page 30: Augmented Self-Regulation (Moraveji, 2012)

12

Figure 2.1: An abbreviated diagram of the complementary functions of the sympathetic and parasympathetic components of the nervous system

(PrecisionNutrition.com, 2012). The stress response is characterized by sympathetic activation and the relaxation response by parasympathetic.

2.1.1 EFFECTS OF STRESS ON LEARNING, PERFORMANCE, AND BEHAVIOR

Even mild chronic stress has deleterious effects on an individual’s capacity to

learn and be productive. While common interpretations of stress are often associated with

significant negative life events, so-called ‘burnout’ (in both school and workplace) results

from “psychological strain that is especially connected with chronic day-to-day stresses

and is marked by a state of physical, emotional, and mental exhaustion” (Etzion, 1984, p.

616). As such, even mild stressors have accumulated effects if an individual does not

Page 31: Augmented Self-Regulation (Moraveji, 2012)

13

effectively adapt to them. It is clear how this occurs so frequently given that any “mental

activity appears to invariably be associated with some degree of sympathetic activation,

whose degree depends on the amount of stress involved in performing the mental task”

(Bernardi,Porta, Spicuzza, Sleight, 2005).

In a review, Marin et al (2011) illustrates the link between worker burnout and

cognitive dysfunction. These effects interact with psychological wellbeing when “the

strain of chronic workplace distress overwhelms the psychological capacities of the

individual and puts them at greater risk of psychiatric manifestations such as burnout and

depression” (Marin et al., 2011, p. 3). Such burnout has been linked to decreases in

cognitive performance and the inability to sustain attention (van der Linden et al., 2005).

In a review article, McEwen and Sapolsky (1995) describe how stress hormones

impair cognitive function and result in a loss of neurons in the hippocampus in two time

scales: short-term and long-term. Newcomer et al (1999) simulated realistic elevated

stress levels synthetically and found decreases in verbal declarative memory. Luine et al

(1994) similarly induced chronic stress and found it impaired spatial memory but that the

effects were reversible.

Allostatic load (McEwen, 1999) is “the physiological consequence of chronic

exposure to fluctuating or heightened neural or neuroendocrine response that results from

repeated or chronic stress” (Taylor, 2006) (see Figure 2.2 for a visual explanation). It

can accrue as a result of four processes: 1) frequent activation of the stress response, 2)

failure to habituate to repeated stressors of the same kind, 3) failure to shut off the stress

response appropriately, and 4) inadequate reaction to the stress response.

Page 32: Augmented Self-Regulation (Moraveji, 2012)

14

Figure 2.2:A diagram showing McEwen’s theory of allostatic load as emerging from interactions between environmental stressors, individual differences, behavioral responses, and physiological responses (McEwen, Gianaros, 2011).

The deleterious effect of stress on memory has been differentiated from the

effects of emotionally arousing stimuli alone (Lupien et al., 2007). Stress, in effect,

‘distracts’ the brain from focusing on the intended material and impacting retention. This

is clearly a high-priority problem given the sociocultural dynamics of a typical classroom

or workplace.

The Nobel Prize in Medicine for 2009 was awarded to Elizabeth Blackburn for

the discovery of how chromosome-protecting telomeres are damaged by chronic stress

(Epel et al, 2004). Members of the same research team found that three months of

intensive meditation were able to reduce the damage and essentially reduce cell aging

(Epel et al., 2009).

Academic test anxiety is another active area of research at the intersection of

stress and cognitive performance. Cassady and Johnson (2002) found increased

performance on the SAT correlated with lower anxiety on a psychometric instrument

meant to assess test anxiety.

Page 33: Augmented Self-Regulation (Moraveji, 2012)

15

Stress and individual reactivity to stress has been identified as a key factor

affecting school readiness and socioemotional behavior (Obradovic et al., 2010). In

children, there are two sociobehavioral patterns of physiological (autonomic) reactivity to

stress: “under-arousal during the resting state or in response to challenging stimuli tends

to be related to externalizing symptoms, while overarousal is associated with

internalizing symptoms” (Obradovic & Boyce, 2009; p. 301). Even low self-esteem, a

predictor of stress reactivity (Pruessner et al., 2004), has been linked to stress due to

associations between chronic stress and reduced hippocampal volume (Pruessner et al.,

2005).

The effects of chronic stress on mental health are apparent and subject to ongoing

study. Prolonged stress is a major risk factor for depression (Mazure & Maciejewski,

2003) and exposure to traumatic stress can cause post-traumatic stress disorder (PTSD)

(Southwick et al., 2005). These are largely subjective or implied, internal stressors

perhaps long after the actual environmental stressor occurred. If stress does not cause

mental health disorders, it likely exacerbates it – e.g., with schizophrenia (Dohrenwend et

al., 2005) and bipolar disorder (Hammen & Gitlin, 1997).

2.1.2 SELF-REGULATION: THERAPEUTIC AND DEVELOPMENTAL

A review by leading neuroendocrinologists describes the effects of stress at each

stage of life, from prenatal to old age (Lupien et al., 2009). Writing about a study on rats,

however, the authors note “it is interesting to note that in contrast to the effects of chronic

or severe stress on the brain and behaviour earlier in life, which are long-lasting, effects

of adulthood stress — even chronic stress — are reversed after a few weeks of non-

stress” (p. 438). Thus, while study of the negative impact of stress continues to be

Page 34: Augmented Self-Regulation (Moraveji, 2012)

16

important, some recent research programs have focused on approaches for recovery and

self-regulation against stress and their efficiency and effectiveness. This work suggests

the importance of increasing, or supporting one’s internal process of increasing, states of

“non-stress,” especially in young people, as studies show chronic stress in youth likely

becomes a learned pattern of response and causes permanent damage.

Porges, one of the leading figures in the study of self-regulatory (autoregulatory)

processes, notes “self-regulation is a difficult process to operationalize” (Porges et al.,

1994). It means different things in different contexts – but these contexts impact one

another and cannot be treated completely independently. The body and mind are inter-

connected as are affective and cognitive processes. Porges relates self-regulation to “self-

soothing” (p. 182). The physiological body, too, can be seen as self-regulating or to self-

soothe itself. Self-regulation in this context refers to the adaptive process by which one’s

body and mind respond to environmental stimuli to maintain homeostasis and deliver the

organism back to a neutral (non-stress) state. Though often subconscious and tacit, self-

regulation is also a learned behavior, being impacted by our choices, sociocultural norms,

and environmental influences.

Self-regulative ability is particularly important in modern society and non-warring

nations because “absolute stressors [such as physical threats to our safety] are rare”

(Lupien et al., 2007). It is implied, self-created stressors that we must respond to most

consistently. These include giving a presentation, facing stereotype threat (Steele, 1995)

(the concern or anxiety of conforming to a negative stereotype about one of one’s social

groups), taking an exam, approaching authority figures, and doing homework. Because

stressors are so often implied, states of non-stress cannot be created solely by controlling

Page 35: Augmented Self-Regulation (Moraveji, 2012)

17

the environment. Our psychophysiological state emerges from an interaction between

environment and individual’s self-regulative ability. As a result, human societies have

developed multiple approaches to developing this ability, leading to psychological

(reducing cognitive or emotional stress and anxiety; confidence and relatedness),

physiological (developing calm habits in heart, muscle, and respiration), and behavioral

(calming rituals and practices) .

It is important to distinguish between therapeutic and developmental techniques

for influencing self-regulation. Therapeutic techniques are meant to induce calm states

but not necessarily to train the nervous system for future adaptation to stress (i.e., calm

states, not traits). Developmental techniques are more training-based in this regard, as

they train the body and mind to identify stressors and adaptively self-regulate.

Pranayama (literally ‘extension of the breath’) and qigong (literally ‘skills gained

through working with breathing’) are probably the oldest and most mature forms of

developing one’s self-regulative ability (Iyengar, 1985). Westerners often conflate these

psychophysiological practices with spiritual development but the two are independent.

Both approaches include mechanistic and prescriptive descriptions of techniques to

evaluate and develop one’s self-regulative ability. While focusing on the breath, both

systems describe the mechanistic interactions of the breath with the rest of the

cardiovascular system and patterns of the mind including thoughts, emotions, and focus.

The appearance of a related science of self-regulation in the West was autogenic

training (Luthe & Schultz, 1932) in Germany. This is a still-popular set of relaxation

practices with guided imagery that aims to influence and train the autonomic nervous

system. This was complemented in America by the therapeutic technique of progressive

Page 36: Augmented Self-Regulation (Moraveji, 2012)

18

muscle relaxation (Jacobson, 1938). Adoption or invention of other therapeutic

techniques has since blossomed in the West, delivered via haptic (e.g., touch, caress),

environmental (e.g., eco-therapy), auditory (e.g., soothing tones, mantras), and olfactory

(e.g., lavender) modalities.

Physiological approaches to developing self-regulative ability are complemented

by contemplative practices such as mindfulness (Kabat-Zinn et al., 1992; Langer, 1990)

and transcendental meditation (Russel, 2002). As of 2007, complementary and alternative

medicines (including contemplative practices) were in use by over 35% of the US adult

population (Barnes, 2002). Recent randomized controlled trials have also lent insight into

the efficacy of mindfulness training in children (reviewed by Obradovic, Portilla, and

Boyce, in press). These practices are used both therapeutically and developmentally

through the conscious intention of entering calm states despite subtle anxieties and

implied stressors. Contrary to popular misconceptions, contemplative practices are meant

to produce an aroused and focused state (Cahn & Polich, 2006), rather than the inert,

relaxed, and detached state often associated with meditation.

In the learning sciences, ‘self-regulated learning’ is used to describe a conscious,

metacognitive process by which learners can monitor their own learning, motivation, and

behavior (Corno & Mandinach, 1983; Corno & Rohrkemper, 1985). Three components of

self-regulated learning seem paramount (Pintrich and De Groot, 1990): metacognitive

strategies for planning, monitoring, and modifying one’s cognition (e.g., Brown,

Bransford, Campione, & Ferrara, 1983; Corno, 1986; Zimmerman & Pons, 1986, 1988),

self-regulating effort to overcome distraction on academic tasks (Corno & Rohrkemper,

1985), and cognitive strategies (e.g., rehearsal and elaboration) to learn, remember, and

Page 37: Augmented Self-Regulation (Moraveji, 2012)

19

understand the material (Corno & Mandinach, 1983). A review by Boekaerts (1997)

discusses how self-regulated learning can be domain-specific, where learners can apply it

in one subject but not in another. Stress has been shown to reduce plasticity of the limbic

system and one’s ability to learn (Sapolsky, 2003). Providing a neurological explanation

for how stress could inhibit metacognitive self-regulation.

Developmentally, behavioral self-control in children grows over time. “Brain

regions essential to self-control are immature at birth and develop slowly throughout

childhood” (Tarullo, Orbradovic, Gunnar, 2009). In children, developmental techniques

include training executive functions (EFs) that regulate behavior, attention, and emotion

(reviewed by Obradovic, Portilla, and Boyce, in press) – skills that are taken to be

precursors to academic and social achievement. Techniques with children have included

meditation, heart-rate variability feedback, behavioral interventions, and attention

training.

2.2 AUGMENTING SELF-REGULATION WITH BREATH MODIFICATION

The link between psychological state and respiration is non-intuitive but has been

well-studied for thousands of years. As B.K.S. Iyengar (1985) translates from Patanjali’s

Yoga Sutras (500-200 BC):

"Pranayama is the regulation of the incoming and outgoing flow of breath with retention. Pranayama has three movements; prolonged and fine inhalation, exhalation and retention; all regulated with precision according to duration and place. The fourth type of pranayama transcends the external and internal pranayamas, and appears effortless and non-deliberate. Pranayama enables the mind to become fit for concentration, and removes the veil covering the light of knowledge and heralds the dawn of wisdom."

Page 38: Augmented Self-Regulation (Moraveji, 2012)

20

More recently, in the West, Charles Darwin noted “the breathing to be hurried” in

men in stressful situations (Darwin, 1872). Sigmund Freud (1962) also recognized that

anxious patients often experienced difficulty breathing. This section motivates the use of

breath awareness and regulation as an effective technique for not only assessing one’s

state but influencing it. Table 2.1, at the end of this section, provides a useful summary

of the key concepts along with lay descriptions to ensure the reader can follow the myriad

concepts easily.

It is common knowledge that one’s heart rate increases in stressful conditions.

Less common is the understanding that heart rate variability (HRV) occurs continuously

and illustrates the body’s ability to adapt to even subtle demands of body and mind. An

example is the simple act of inhaling and exhaling: the heart rate increases during each

inhalation and decreases with each exhalation (Sherwood, 2006).

The degree of change in heart rate during respiration, or respiratory sinus

arrhythmia (RSA), is commonly used as a robust measure of parasympathetic nervous

system (PNS) activation. Higher RSA amplitudes (i.e., the difference in heart rate

between inhalation and exhalation) are associated with healthier individuals and longer

life expectancy. RSA can be averaged over time but changes can also be measured in

response to temporary demands. In general, high RSA is commonly used to indicate

autonomic adaptability to both positive and negative demands (Obradovic & Boyce,

2009).

RSA and other psychophysiological indicators such as skin conductance level

(SCL) and salivary cortisol responses do not by themselves alone explain one’s inability

to self-regulate. For example, environmental or familial influences play a role in dictating

Page 39: Augmented Self-Regulation (Moraveji, 2012)

21

a child’s adaptability and symptoms independent of individual stress reactivity

(Obradovic, Bush, Stamperdahl, Adler, Boyce, 2010). However, these

psychophysiological indicators can complement or provide cues for self-regulation

development. Respiration, in particular, is well-suited because it is relatively easy to both

consciously observe and modulate (unlike the others). It is the focus of this review.

The influence of stressors on RSA cannot be studied without accounting for

changes in respiration that result from engaging in those activities. Bernardi et al (2000)

found that “simple mental and verbal activities markedly affect HRV through changes in

respiratory frequency” (Bernardi et al., 2000; p. 1462). Simply put: engaging in a task

puts a certain amount of demand on the body, which is reflected by changes in the

interconnected cardiovascular system (including respiration).

Stress causes mild or acute hyperventilation (over-breathing) as dictated by the

stressors’ intensity or duration and the person’s innate or learned stress-reactivity (Suess

et al., 1980; Boyce et al., 1995; Van Diest et al., 2001). Dr. Herbert Fensterheim, clinical

professor of psychology in psychiatry at Cornell University Medical College in New

York City, notes that “depending on the person, any emotional stimulation can set off

over-breathing” (Flippin, 1992, p. 24).

The link between respiratory pattern and stress is often, but not always (Fried,

1990), studied in individuals with panic disorder (Wilhelm, Gevirtz, Roth, 2001; Meuret,

Wilhelm, Roth, 2001; Roth, 2005; Conrad et al., 2007), ostensibly because the

association is plainly evident, effects are more easily measured, and motivation is great

because it is a formalized chronic condition. However, chronic and acute stress are

experienced not only by those with a clinical condition but by almost anybody.

Page 40: Augmented Self-Regulation (Moraveji, 2012)

22

Therefore, we draw upon studies done with those panic disorder and find there are

similarities with theories posed by yogic and qigong scientists. There are two competing

theories (‘hyperventilation’ theory and ‘suffocation false alarm’ theory) to explain how

stress and respiration rate are related, summarized by Roth (1999):

The hyperventilation theory of panic postulates that falls in arterial blood CO2 precede and cause attacks, at least many at the beginning of the illness and when other variables such as the buffering effects of blood electrolytes are taken into account. Klein’s suffocation false alarm theory postulates that panic is provoked by sudden feelings of dyspnea triggered aberrantly in the same brain circuits that monitor blood CO2 levels. This theory sees hyperventilation as a compensatory reaction to this alarm rather than a cause of panic. Klein contrasts panic anxiety to ‘‘anticipatory anxiety’’, which is linked to external stimuli through Pavlovian conditioning and which is characterized more by cardiovascular than respiratory symptoms.

It has also been shown that those suffering from panic disorder “do not show a

unique tendency toward hyperventilation, but rather that their hyperventilatory symptoms

and perhaps intermittent over-breathing episodes are a function of the high levels of

anxiety they experience” (Holt, 1989). It follows that subjective stress or self-regulative

inability is the issue at hand rather than limiting the issue only to people labeled as having

panic disorder per se. People who are not classified as having ‘panic disorder’ have also

been repeatedly shown to react with faster breathing when exposed to a stressful situation

(Suess, 1980; Ley, 1999; Conrad et al., 2007; Roth, 2005).

In healthy individuals in non-stressful environments, one’s autonomic nervous

system determines the exact breathing pattern necessary to maintain homeostasis in the

body. As environmental and mental demands are made, this homeostasis is disturbed and

breathing patterns, being “exquisitely sensitive to [external or internal] stress” (Ley,

Page 41: Augmented Self-Regulation (Moraveji, 2012)

23

1999), adjust accordingly. Indeed, “distinct effects of stress are apparent even on a small

time scale (seconds-minutes)” (Bernardi, Porta, Spicuzza, Sleight, 2005). A number of

studies have demonstrated the relationship that affective state has with respiratory pattern

(Porges, 1994; van Diest et al., 2001). Just as the stressed mind is under conscious control

to change focus from stress-inducing thoughts, so is respiration under conscious control

to change patterns to calming ones. This is the bi-directional, causal relationship between

respiratory patterns and affective state.

In summary, the stress response prepares the body for quick action – including

shallow and fast respiration resulting from heightened heart rate – but prolonged

activation has significant costs on the brain and body. Respiratory pattern is an indicator

of one’s psychophysiological state that can be assessed without sophisticated feedback

and is one that can be modulated quickly and easily by practically anybody.

Though the link between stress and respiration is physiologically evident, it is

surprisingly not a common focus of research in psychophysiology. As Grossman (2007)

notes, “the significance of respiratory influences upon cardiovascular functioning has

been much neglected in the psychophysiological literature” while more attention is paid

to HRV (Mulder, 1992; Brown et al., 1993; Mulder, de Waad, Brookhuis, 2005). Perhaps

this is because respiration is under conscious control, making it more difficult to control

for in laboratory studies. Another reason could be because respiration is seen as

‘superficial’ compared to heart rate and rhythm. Yet another reason could be because

HRV and RSA are relatively easy to measure compared to subtle respiration

characteristics such as tidal volume, the amount of air breathed in a single inhalation or

exhalation, which are more difficult to measure unobtrusively.

Page 42: Augmented Self-Regulation (Moraveji, 2012)

24

Beyond the relatively little research that exists studying how stress and respiration

are linked, we observe even less scholarly work on how this bi-directional relationship

can be exploited to reduce stress-reactivity. In summary, for researchers interested in

augmenting (not only measuring) self-regulation, respiration is the central process to

study because it reflects mental demands, can be consciously regulated, and is relatively

easy to measure.

2.2.1 REGULATING RESPIRATION TO MITIGATE THE STRESS RESPONSE

Voluntary breath regulation is a common, empirically validated technique for

reducing stress and anxiety (Clark & Hirschman, 1990; McCaul, Solomon, Holmes,

1979; Sisto et al., 1995; Ley, 1999), relieving symptoms of asthma (Cooper & Oborne,

2003), reducing blood pressure (Grossman et al., 2001; Schein et al., 2001), and focusing

the mind (Ley, 1999). Changes in breath may result from frustration, pain, stress, test

anxiety (Ley & Yelich, 1998), and post-traumatic stress disorder (PTSD) (Zucker et al.,

2009). Schliefer and Ley (1994) investigated the effect of computer data entry compared

to baseline relaxed periods. They found that breath rate increased 26% during data entry

and that subjects exhibited decreased heart rate variability (HRV) and increased self-

ratings of tension. The differences in breath rate between resting and data entry were

comparable to those between rest and feeling threats of electric shock.

Humans can “regulate their respiration rates in a relatively short time period”

(Ley, 1999), making breath regulation a viable treatment for sporadic and subtle stressors

such as those that may be encountered in everyday settings such as information multi-

tasking. However, developing adequate continuous awareness and motivation to monitor

Page 43: Augmented Self-Regulation (Moraveji, 2012)

25

one’s breathing behavior is problematic and requires considerable attention and is often

conflated with spiritual pursuits, making it a suitable candidate for technology mediation.

Breath rate (breaths per min, or bpm) is calculated by the number of times the

chest rises (inhalations) in a minute. Breath rate, like heart rate, changes frequently even

in a resting state as it is affected by arousal, talking, posture, personal health, and other

factors (Ley, 1999). There is no standard resting human breath rate; studies have found

mean rates to range between 12-20bpm for Western adults (Ley, 1999; Sherwood, 2006).

There are different perspectives on what an optimal target breath rate should be. Song

and Lehrer found that an optimal range of 4-6bpm (i.e., very slow), usually in a state of

restful alertness, is correlated with the greatest HRV but that HRV generally increases as

breath rate decreases (Song, Lehrer, 2003). Stark et al. showed that a target rate too

different from one’s resting breath rate requires greater attention and effort to change it

(Stark, Schniele, Walter, Vaitl, 2000).

Referring to the hyperventilation theory, lowering the breath rate can bring carbon

dioxide back to normal levels, allaying neurological alarms (Ley, 1999). A 1970 study of

101 acute psychiatric patients in a hospital found that clinical improvement was

significantly associated with a mean decrease of resting breath rate of 3.4bpm (Skarbek,

1970). However, recent research has shown that if one breathes too slowly, carbon

dioxide levels may decrease so much that feelings of pressure or stress result (Roth,

2005). As a result, interventions must focus not on encouraging users to breathe at

arbitrarily low rates but at their own resting rate or perhaps at a rate in between.

Monitoring the breath, without instructed regulation, is itself an effective means of

regulating the breath. Conrad et al found that in a clinical study, simply “paying attention

Page 44: Augmented Self-Regulation (Moraveji, 2012)

26

to breathing significantly reduced respiratory rate and decreased tidal volume instability

compared to the other instructions” (Conrad et al., 2007).

The vast majority of studies we inspected were based on short breathing

instructions where the effects were measured over controlled durations in laboratory

environments. Even in these short durations of breath regulation, parasympathetic

activation is triggered. However, “this effect could be enhanced if respiration could be

maintained at a regular (rather than episodic) slower rate, and (by consequence) increased

tidal volume” (Bernardi et al., 2005). Training or triggering users to regulate their

respiration in light of even mild stress crosses both behavioral and physiological

responses in McEwen and Gianaros’s conceptual diagram of emergent allostatic load in

Figure 2.2. It is a conscious behavior but within the physiological system only, not

requiring tangible action. This means, we hypothesize, it can be done in parallel to

existing tasks and life events.

Meuret, Wilhelm, and Roth (2001) had participants with anxiety disorder use a

handheld sensor that senses the volume of the breath 5 times over a four-week period

(administered by a psychologist). The ultimate goal in their bio-feedback-augmented

respiration influence was slow, shallow, and regular diaphragmatic breathing. They

showed patients had significantly fewer panic symptoms and reduced physiological

indicators for stress and depression solely by inducing regularity and reduced rate of

respiration. Further, they showed that the effects were retained after the study. The team

speculated that the “biofeedback breathing training probably makes patients feel more in

control of their bodily reactions and makes them react less fearfully to them” (p. 600). It

was tidal volume, rather than rate alone, that the team used to explain the effect. This

Page 45: Augmented Self-Regulation (Moraveji, 2012)

27

further supports the claim that changes in rate must be complemented by changes in tidal

volume. A key limitation of the study was that the sensor was not ambulatory and could

not sense the patient continuously. The same team later identified a training method that

involves 4 characteristics of the breath: increased ventilation (respiratory rate x tidal

volume), breath-to-breath regularity in rate and depth, rate alone, and chest breathing

(Meuret, Wilhelm, Roth, 2004). It was feedback of breath rate and tidal volume alone that

helped participants “facilitate voluntary control of respiration and reduce symptoms” of

anxiety.

Though it may seem to follow from these studies that optimal respiration is that

which is constantly slow and deep, this is not true. Arousal, broadly speaking, is a natural

part of the experience of life and is crucial to physical and cognitive performance. The

aim is not to avoid all fluctuations in respiratory state but to support the ability to

effectively and efficiently recognize one’s sustained and unnecessary arousal and return

to a neutral state. This is self-regulation in its essence: the self-awareness to monitor

one’s state and the capability to influence it.

Part of the value of this dissertation is that it draws upon multiple fields:

psychophysiology, biofeedback, interaction design, electrical engineering, and behavior

change. Due to this diversity, Table 2.1 illustrates in a plain manner the key concepts

and mechanisms motivating breath regulation as a means of regulating

psychophysiological state.

Page 46: Augmented Self-Regulation (Moraveji, 2012)

28

# Concept Description

1 Psychophysiology Study of the link between psychological processes such as fear and physiological response such as dilated pupils. The link also exists in the opposite direction, between physiological processes such as increased breath rate and psychological processes such as anxiety.

2 Sympathetic response

The part of the human nervous system that prepares the body for action. Dilates pupils, contracts muscles, etc.

3 Parasympathetic response

The part of the human nervous system that initiates healing processes.

4 Allostatic load When the sympathetic nervous system is over-activated, the presence of stress hormones in the body ‘builds up’.

5 Self-regulation As it is used here: the innate, learned, and controlled process of maintaining balance between the sympathetic and parasympathetic systems.

6 Stress-reactivity One’s innate, learned, and controlled reactivity to environmental or imagined stressors.

7 Heart rate (HR) When the sympathetic system is activated, heart rate increases to pump blood more quickly to muscles so they can take the person more quickly away from the threat.

8 Heart rate variability (HRV)

The heart rate is actually constantly fluctuating. The amount of fluctuation is an indicator of how responsive the body is to its demands. High HRV is an indicator of general health.

9 Respiratory Sinus Arrhythmia (RSA)

The difference in heart rate between respiration inhalation and exhalation. The larger the amplitude, the more healthy the individual.

10 Respiratory pattern

Scientists have found a link between the way one breaths and their cognitive and emotional state. Breathing pattern includes rate, regularity, volume of air, etc.

11 Respiratory regulation

Consciously regulating one’s respiration pattern is a tangible means of influencing one’s mental state & reducing anxiety.

12 Respiration rate (RR)

Also called ‘breath rate’. High breath rate is indicative of anxiety and sympathetic activation. Low breath rate is correlated with a calm state of mind and high HRV.

13 Tidal volume The amount of air inhaled or exhaled. When reducing breath rate you want to increase tidal volume (slow, deep breathing).

14 Chest or diaphragmatic breathing

Another characteristic of the breathing pattern is the location: is the breathing ‘shallow’, where only the top of the lungs moves in and out? Is it from the diaphragm or belly? Or is it full-chest breathing?

15 Breath awareness The notion of simply devoting some amount of attention to the state of the breath, without consciously influencing it. This has been shown to effectively regulate and calm the breath.

Table 2.1: Lay descriptions of the key concepts motivating breath regulation as a means of regulating psychophysiological state. Not meant to be comprehensive.

Page 47: Augmented Self-Regulation (Moraveji, 2012)

29

2.3 TECHNOLOGY-MEDIATED RESPIRATORY SELF-REGULATION

Based on the review of the psychophysiological literature above, we identified

respiration rate as an appropriate candidate for modulating the stress response because

respiration:

- …is a direct mechanism for modulating the autonomic nervous system (as

opposed to heart rate, which we can control only indirectly).

- …is relatively simple to interpret (i.e., fast, slow, or average).

- …can be sensed relatively easily (e.g., using a thoracic strain gauge) in an

ambulatory manner.

- …can be regulated while one is engaged in other tasks.

- …does not require high precision sensor accuracy for reasonably useful

information. More granular information can be more useful but rate alone can

develop a sense of self-awareness and, as long as users increase tidal volume

accordingly, is a robust method of mitigating stress.

We do this with the caveat that rate is only one of four possible respiration characteristics

that could be used in aiding self-regulation. Tidal volume is relevant to both theories of

the relationship between stress and respiration.

Many people are motivated to, encouraged to, and even clinically prescribed to

practice breath regulation during their daily lives but they need support to not only assess

their state and guide breath regulation but to remember to do it at the most relevant time

(Meuret, Wilhelm, Roth, 2001; Wilhelm, Roth, Sackner, 2003). This section explores

technological methods of providing that support. The question of how best to motivate

users to do this is outside the scope of this dissertation.

Page 48: Augmented Self-Regulation (Moraveji, 2012)

30

The modern laboratory study of using technology to shape or develop one’s self-

regulative ability started, arguably, in the early 60’s by the research Armenian-Canadian

scientist John Basmajian published in Science (1963) and by the work of Neal Miller

(1972). Research in this vein has been active since it was popularized at Harvard

(Shapiro, Tursky, Schwartz, 1970). It waned but, with the advent of ubiquitous

technologies and the demands of information work, has grown in recent years.

Traditional biofeedback, which attempts to train patients to change their state on

command, is an example of a developmental technique to increase calm because it aims

to develop one’s own self-regulatory processes.

Clinical biofeedback practitioners have long encouraged patients to “use

breathing to focus attention, reduce arousal during the day, and inhibit the somatic

responses induced by stressful stimuli and pain” (Peper, 2003). “Appropriate

modification of the respiratory pattern can in fact induce changes that appear to have

useful clinical applications in different diseases” (Bernardi, Porta, Spicuzza, Sleight,

2005). Practitioners use a number of methods to influence breathing patterns, falling

generally into two categories: (1) self-assessment or (2) guiding breath regulation.

The first category offers users insight into their state for self-assessment. The

iPhone-based MyBreath (2011) infers respiratory pattern from the microphone’s input as

the user breathes into their headset. Azumio’s Stress Check (2011) uses the phone’s

camera as a pulse oximeter to detect heart rate through their finger and give the user

feedback about their state. The LifeShirt offers researchers a glimpse into the

psychophysiological state of the user as well (Wilhelm, Roth, Sackner, 2003). Similar

wearable research systems offer similar feedback to users about their state when they

Page 49: Augmented Self-Regulation (Moraveji, 2012)

31

synchronize the data with their computers and reflect about the patterns (Fletcher et al.,

2010).

Techniques and tools to guide breath regulation include visual animation (Sisto et

al., 1995), modulating ambient lighting in a room between binary states to guide

inhalation and exhalation (Lehrer, Vaschillo, Vaschillo, 2000; Stark, Schnienle, Walter,

Vatil, 2000), or a virtual metronome (Roth, 2005). Resperate (Gavish, 2010) is a

Walkman-like device that uses auditory tones to guide relaxed breathing to reduce blood

pressure. StressEraser (2005) and emWave (McCraty et al., 1999) measure and guide

users towards larger RSA amplitudes – the latter in coherence with other physiological

rhythms. The iPhone-based Pranayama (2011) uses pie charts and human figures with

animated chests to guide the practice of known stress-reduction techniques using paced

respiration.

In addition to mobile use cases, desktop computers hold potential use, primarily

because of the long durations that many users spend with them. The publicly available

‘Calm Down’ desktop application (2011) dims the screen to pace respiration in a calming

inhalation/exhalation pattern at predetermined intervals or upon user demand. Another

common technique is “break-reminder” software (Morris et al., 2008). Morris et al.

(2008) acknowledged the desire to maintain productivity while taking physical breaks.

Desktop-based biofeedback games have also been explored to motivate HRV regulation

and neurophysiological coherence (McCraty et al., 1999).

Pacing stimulus design can lead to ambiguity about how long each inhalation and

exhalation will be. Methods with binary states (e.g., when will the light turn off?) or

intensity (e.g., what is maximal brightness?) do not clearly indicate the end point of the

Page 50: Augmented Self-Regulation (Moraveji, 2012)

32

current inhalation or exhalation. These problems are not frequently acknowledged in the

literature to our knowledge, assumedly because the physiological implications are of

primary interest.

A noted trend in the literature is the practice of using multiple physiological

sensors to gain a comprehensive picture of the user’s state. If the user is shown multiple

feedback channels, perhaps they could more adeptly self-regulate their state. This

increases cognitive load so an important question becomes, ‘what physiological data is

best to feedback to users to most effectively help them self-regulate?’ A study by Reyes

del Paso, Godoy, and Vila (1992) comparing the effects of RSA amplitude, RSA

amplitude plus respiration, respiration biofeedback alone, and respiration instructions

showed that it was the respiration conditions that produced the most efficient changes in

RSA amplitude (as opposed to directly showing the user their RSA amplitude). Even the

condition that used RSA amplitude and respiration biofeedback resulted in slower

changes because of the extraneous cognitive load required for users to “search for [RSA]

control strategies” (p. 272). Further, participants reported they even though there was no

explicit instructions to do so, they used respiration to modulate RSA amplitude, a clear

indicator that RSA amplitude feedback is an indirect feedback element that requires

conscious attention, rendering it less useful for user feedback. The study showed

conclusively that “the parasympathetic cardiac outflow seems to be controlled easily in a

voluntary way in normal subjects by means of simple strategies for changing the

respiratory pattern toward slower respiratory rates and greater respiratory amplitudes,

with or without the help of biofeedback” (p. 273).

Page 51: Augmented Self-Regulation (Moraveji, 2012)

33

HCI research has also explored human breath as an input modality. Marshall et al.

(2011) reviewed relevant literature on using breath for two purposes: (1) an assistive

device to aid otherwise challenged users and (2) as expressive input into systems

attempting to infer affective state (e.g., games (Nacke et al., 2011)). Neither category

attempts explicitly to produce desired respiratory patterns in the user. Affective

computing (Picard, 2003), a related class of computing systems, attempts to infer, detect,

or influence affective state (rather than psychophysiological state). Some of these are

more recently being applied to health and self-awareness, e.g., the digital mirror (Poh,

McDuff, Picard, 2010). What differentiates our work is (a) the continuous nature of the

monitoring and feedback and (b) a focus on respiration and not affective state or valence.

Respiration rate is arguably the ideal data source for user feedback. Existing

systems suffer from requiring high amounts of self-discipline to use because they are

essentially means of explicitly training the user. We have not seen in the published

literature methods of continuously influencing breathing patterns to augment self-

regulation in a continuous manner.

2.3.1 SOCIAL INFLUENCE ON PHYSIOLOGICAL BEHAVIOR

Social awareness is “an understanding of the activities of others, which provides a

context for your own activity” (Dourish & Belotti, 1992). It can emerge via social activity

indicators (Ackerman & Starr, 1995) without requiring direct social interaction.

Peripheral displays have been used to make such indicators perceivable in a glance. For

example, EventManager (McCarthy & Anagnost, 2000) uses a peripheral display to

notify users of location and person-specific events that are relevant to collaborative

organizational work. Sideshow (Cadiz, Venolia, Janke, Gupta, 2001) displays personal

Page 52: Augmented Self-Regulation (Moraveji, 2012)

34

information management data in an always-visible peripheral desktop display, allowing

users to drill-down to details at any time.

In addition to studies of improved productivity, recent HCI research focuses on

using social systems to motivate behavioral change. For example, Maitland and Chalmers

(2011) conducted a study of how peer influence can be designed into systems promoting

weight management. They review systems (e.g., ActiveShare (Fialho et al., 2009))

employing socially motivating mechanisms such as setting commitments, sharing data

with others, and normative influences (Cialdini, 2008).

Though we draw on social systems for behavior change, our primary focus is on

mitigating stressors. In this vein, researchers have studied affective social games in which

users “out-calm” one another as measured by skin conductance and/or heart rate. Bersak

et al.’s (2001) early example found two primary problems with such systems: (1) players

were frustrated that they could not control these physiological measures directly and (2)

falling behind in competitive games can add to the user’s frustration.

In social psychology, synchrony can foster cooperation, rapport, in-group identity,

and altruism by strengthening social attachment among group members (Valdesolo et al.,

2010). Valdesolo et al. (2010) note that “[s]ynchronous others are not only perceived to

be more similar to oneself but also evoke more compassion and altruistic behavior than

asynchronous others experiencing the same plight.” Synchronous physical behavior is

rarely employed in virtual tools, assumedly because individual use is the norm. The

“parallel games” of Mueller et al. (2010) showed how shared physical experience bonds

and motivates distributed participants by connecting users as they jog or run.

Page 53: Augmented Self-Regulation (Moraveji, 2012)

35

We have not seen in the literature exploration into using the psychology of

synchronous physical behavior applied to cardiovascular or physiological behavior such

as breathing in order to intentionally influence respiration patterns. The literature points

to three reasons that synchrony may be useful for CSCW designers: synchrony may be

used to (1) address the free-rider problem by increasing empathy and relatedness

(Wiltermuth & Heath, 2009), (2) motivate and facilitate learning via imitation, and (3)

build connectedness.

2.4 SUMMARY AND IMPLICATIONS FOR CURRENT RESEARCH

The literature demonstrates a significant research opportunity to augment

continuous respiratory self-regulation with continuous respiration rate monitoring,

feedback, and incentive systems. Augmenting self-regulation, rather than motivating

users to do therapeutic exercises, represents a significant departure from prior approaches

and leads to interesting research questions that are of interest to researchers in HCI. The

creation of such novel tools could complement intensive training, regular practice, or

stand alone as a means of integrating calming influences into our technological

ecosystem.

In the present dissertation, our focus is solely on supporting and incentivizing

users to breathe at their personal resting rate in parallel to existing work tasks while at the

computer and beyond. We will investigate design techniques and evaluate their efficacy

while also investigating their effect on cognitive performance. It is our intention that

research in this domain can inform a class of future ‘calming technologies’ to improve

individual and collective health and productivity.

Page 54: Augmented Self-Regulation (Moraveji, 2012)

36

Page 55: Augmented Self-Regulation (Moraveji, 2012)

37

DESIGN AND VALIDATION OF A NON-CLINICAL AMBULATORY RESPIRATION SENSOR

This chapter provides a description of the sensors used in subsequent chapters.

There are many methods of sensing respiratory rate; Al-Khalidi et al., (2011) provide a

review of nine methods used in clinical research. Because the goal of this dissertation is

to support research on non-clinical interaction design and behavior modification, we do

not require clinical-grade accuracy, which allows one to assess precise tidal volume

among other characteristics. Rather, we are interested in sensing of respiration rate in a

continuous manner. As such, a high degree of wearability is desired. The data must also

be sent in real-time to a computing device.

Existing commercial sensors did not support our requirements. As a result, we set

out to design and build a sensor that fit our needs. We identified two dimensions,

accuracy and wearability, that describe a continuum of respiration sensors for the purpose

of continuous feedback of respiration rate. We then inspected the various methods of

sensing respiration and looked to see which would be amenable to our needs. A list of

known methods is shown in Table 0.1.

Page 56: Augmented Self-Regulation (Moraveji, 2012)

38

Method Description Acoustic modeling Uses a microphone near throat/nose. Thermistor Airflow measurement by measuring temperature using a

thermistor near the mouth (e.g., by putting sensor on a pair of glasses and requiring user to breathe only using nose).

Accelerometer Detection of movement of the chest or air passages in the throat using a tri-axial accelerometer.

Transcutaneous CO2 monitoring

Measuring CO2 diffusion using electrodes on skin.

Pulse oximetry Blood-oxygen saturation measurement inferred from changed in heartrate using pulse oximeter (on earlobe or finger), high-accuracy heart rate monitors, or plethysmogram.

Doppler Radar-based measurement on chest expansion at a distance. Optical Sensing of chest movement of stationary subjects using infrared

light projection or static cameras. Thermal Sensing of semi-stationary temperature changes in around the

mouth and nose at a distance. Strain gauge A stretch sensor measures force that the chest emits when

expanding upon inhalation. Elastomeric plethysmography

Detects changes in chest and/or abdomen expansion.

Respiratory inductance plethysmography

Uses thin bands of sinusoid coils woven to measure .changes in the magnetic field generated when they are stretched. One band on the abdomen, one on the chest. Maximum number of uses is 40-60 before they wear out.

Impedance plethysmography

Two electrodes are put on the skin and measures changes in current between them because of respiration expanding the chest.

Table 0.1: Known methods of sensing respiration.

A thoracic strain gauge to detect expansion and contraction of the chest and/or

abdomen (elastomeric plethysmography) was deemed the best choice given our needs:

wearability, non-invasiveness, and capability of sensing respiration rate accurately

enough for practical use. It is important to note that because our needs are not purely on

accurate description of tidal volume, we were able to focus on the most wearable means

of deriving respiration rate. Measuring changes in thoracic circumference is a robust and

straightforward measure of respiration for healthy individuals (i.e., those without apnea

or obstructed airways) as long as they are not engaged in high-movement activities such

as running (Al-Khalidi et al., 2011). In clinical research, Napal, Biegeleisen, and Ning

Page 57: Augmented Self-Regulation (Moraveji, 2012)

39

(2002) were able to use a thoracic strain gauge to detect sleep apnea and parametric

modeling, illustrating its accuracy.

2.4.1 SENSOR DESIGN

This section describes the basic design of the ‘breathbelt’ sensor. There are three

primary components: the belt, the sensor, and the communication mechanism. The strain

gauge sensor is a thin, 2-inch long cylindrical conductive rubber-like material that

increases resistance linearly as it is stretched. It is very sensitive to even subtle

movements – necessary for respiration.

The initial belt design (see Figure 0.1) consists of a single strap with an

adjustable buckle. Such a sensor could be built into clothing or undergarments and does

not require direct contact with the skin. Contrary to the name, the sensor is worn at an

arbitrary location on the abdomen, not around the waist (where belts are normally worn).

Page 58: Augmented Self-Regulation (Moraveji, 2012)

40

Figure 0.1: The initial breathbelt sensor. (Top) The adjustable sensor band with the Arduino Uno board. (Bottom) Close-up of the stretch sensor held in

place by two clips.

Our initial sensor was connected to an Arduino Uno, which in turn is connected to

a computer using a standard wired USB connection. The sampling frequency is 20Hz

(50ms) plus a small OS-level variable error from reading serial data (~2ms), making it

near-real-time.

The communication mechanism varies from wired to wireless methods. Initial

version used USB and grew to wireless (XBee and Bluetooth 4.0). Power consumption

did become an issue we had to solve when conducting continuous sensing.

Page 59: Augmented Self-Regulation (Moraveji, 2012)

41

The sensor delivers a continuous stream of integers representing the amount of

stretch in the sensor, with some noise. This raw signal, filtered signal, and detected peaks

(i.e., inhalation endpoints) are shown over a period of 30 seconds in Figure 0.2. Due to

the presence of high-frequency noise, we smoothed the signal using a Hanning window.

The window width is selected based on a reasonable maximum breath rate of 40bpm. The

peaks are detected by comparing the value of each data point in the filtered signal to data

points taken immediately before and after that point. The breath rate is then calculated

based on the number of peaks in a 30 second interval proceeding that time.

Figure 0.2: The respiration sensor’s raw signal (red) is filtered (blue) and then peaks (black) are detected using well-studied signal processing techniques. The Y-axis refers to raw sensor values (not normalized).

2.4.2 INDUSTRIAL DESIGN EVOLUTION

This section describes the evolution of the breathbelt’s industrial design from user

feedback, communication needs, and other necessities.

2.4.2.1 Iteration 1: Wired USB

The initial version of the strain gauge sensor used a rather bulky Arduino Uno

glued to the belt with a wired connection direct to the computer via USB (Figure 3.1).

The protective elastic material atop the stretch sensor was meant to protect the sensor

from breaking from too much tension, which happens occasionally.

Page 60: Augmented Self-Regulation (Moraveji, 2012)

42

2.4.2.2 Iteration 2: Wireless USB

The second version of the sensor used an XBee wireless communication device

housed on a Lilypad Arduino (see Figure 0.3) to allow for more free range of motion

while the user is at the computer. The XBee communicates with an XBee receiver

plugged into the USB port of a laptop computer.

Figure 0.3: Wireless USB version of the original breathbelt, using paired XBee wireless communication widgets and a Lilypad Arduino (the circular PCB).

The black plastic case on the top-most image holds 2 AAA batteries.

2.4.2.3 Iteration 3: Wireless Bluetooth

The third version of the sensor attempted to collect and transmit respiration data

more or less continuously (to a mobile device), rather than only when the user was close

Page 61: Augmented Self-Regulation (Moraveji, 2012)

43

to a laptop computer. This was a significant difference with previous versions of the

breathbelt. This resulted in two design goals:

1. A less intrusive belt that could be worn for multiple days on end.

2. Continuous data transmission.

To address the first design goal, bulk on the belt was minimized and we ensured

the battery did not require frequent change. The idea was that that sensor would be worn

under one’s primary shirt but over another shirt (i.e., not directly on the skin). Wires and

other components would be kept to a minimum so as to reduce strange bumps on the shirt

or the possibility of catching clothes on the breathbelt. Another innovation here was to

componentize the breathbelt so that the strain gauges, which have been known to break

during harsh use, can be replaced easily. To this end the existing sensors were modified

to be replaceable using snap-on/off buttons (see Figure 0.4).

Figure 0.4: The strain gauge (black) has two hooks (top), which was modified to include buttons to snap in and out of the sensor (bottom). This was necessary

because the strain gauge occasionally breaks and must be replaced.

To address the second design goal, the belt transmits data directly to a mobile

phone. To address power issues, this sensor uses the Bluetooth 4.0 Low Power protocol,

which consumes significantly less power than its predecessor, ideal for sensors. The

iPhone 4S is the first commercial mobile phone to support Bluetooth 4.0. A Texas

Page 62: Augmented Self-Regulation (Moraveji, 2012)

44

Instruments (TI) development kit board with on-board Bluetooth 4.0 chip was used to

make prototyping possible. The board uses a battery and is optimized for low power

consumption. This prototype also had two red interaction buttons on the transceiver itself,

allowing for on-belt feedback (see Figure 0.5).

Figure 0.5: The most recent breathbelt design includes 3 components: (a) adjustable band, (b) strain gauge, and (c) microprocessor Bluetooth 4

transceiver, allowing it to communicate continuously with a mobile phone.

2.5 SENSOR VALIDATION

The aforementioned thoracic strain gauge sensor was compared to a non-clinical,

commercial respiratory sensor manufactured by PASCO (see Figure 0.6) and used in

college and high school science experiments around the world. The PASCO sensor works

by having the participant wear an 8-inch-wide belt around their abdomen which is then

pumped full of air. Changes in thoracic circumference are then reflected in changes in air

pressure in the belt. A gas pressure sensor senses these changes and sends them to a

custom amplifier that is then connected via USB to the computer. Proprietary software

allows us to record and export data from the gas pressure sensor to compare it to changes

in values from our breathbelt sensor.

Page 63: Augmented Self-Regulation (Moraveji, 2012)

45

Figure 0.6: The PASCO respiration sensor used to validate our strain gauge sensor. It uses a gas pressure sensor to measure how air pressure in the belt

changes as the wearer breathes.

A researcher in our team wore both the PASCO device and the USB-wired

version of our thoracic strain gauge simultaneously for a period of 20 minutes. Both

devices were worn around the abdomen, the strain gauge immediately above (and not

touching) the PASCO device. The researcher then worked on a variety of unrelated tasks

on the computer, sometimes consciously changing his breath rate and sometimes not.

We were interested in the similarity between the frequency (not amplitude) of the

two signals. An excerpt of the two raw signals side-by-side is shown in Figure 0.7. We

attribute the small differences between the sensors to OS-level timing interruptions of

other processes which can delay writing to the log file in our sensor. For example, at 235

seconds in Figure 3.5 one can see the breathbelt (in red) skips a cycle.

Page 64: Augmented Self-Regulation (Moraveji, 2012)

46

Figure 0.7: X-axis is Time in seconds and the Y-axis is normalized sensor values. A 5-minute excerpt of our thoracic strain gauge breathbelt (red)

compared with the PASCO sensor (blue) shows their similarity, with some noise. Each peak is the apex of an inhalation.

We ran a smoothing algorithm on both signals to identify the breath rates

averaged over a 120sec. The width of the Hanning window used to smooth both signals

was 4 seconds. The value was selected after trial and error to most optimally delineate

peak detection. We normalized both signals and plotted them on the same graph, as

shown in Figure 0.8 over the full 20-minute period.

Page 65: Augmented Self-Regulation (Moraveji, 2012)

47

Figure 0.8: X-axis is Time in seconds. A comparison between the sensor gauge-based breathbelt and the commercial PASCO sensor that uses highly sensitive air

pressure fluctuations to measure breath rate.

First investigating the signals visually, we see there is a high amount of variance

in the signals as the user worked across different tasks including having a conversation.

The signals are very similar, sometimes diverging but returning to a similar breath rate

and neither signal is notably above or below the other consistently.

The mean difference of -0.028bpm was found between the two sensors. I.e., the

breath rate calculated by the strain gauge was, on average, 0.028bpm higher than that

calculated using the PASCO sensor. The cross-correlation, a measure of correspondence

between two time-series signals, between the two signals was 0.985, making the sensor

more or less identical to the commercial PASCO sensor.

Page 66: Augmented Self-Regulation (Moraveji, 2012)

48

2.6 CONCLUSION

This chapter motivated, described, and validated the design of an ambulatory

respiration sensor optimized for wearability, accuracy, and continuous data transmission.

Four iterations of the breathbelt’s industrial design were described along with a

comparison with a non-clinical commercial sensor, showing the accuracy between the

two sensors is comparable. Different iterations of the breathbelt are used in subsequent

studies presented in this dissertation as the studies progressed and wearability became a

greater emphasis.

Page 67: Augmented Self-Regulation (Moraveji, 2012)

49

CHAPTER 3 PERIPHERAL PACED RESPIRATION: INFLUENCING RESPIRATORY PATTERNS DURING INFORMATION WORK

This chapter explores methods for mitigating stress by motivating calming

physiological habits during computer work, with the advantages that such systems (1)

need not interrupt users’ work, (2) provide continuous, rather than sporadic, self-

regulation goals, and (3) complement relaxation therapies to work across computer tasks.

I investigate methods of integrating respiration-pacing techniques into the desktop

computing environment to enable peripheral paced respiration (PPR). Such a system

would allow users to engage in other tasks while regulating the breath, increasing the

accessibility and frequency of breath regulation. To that end, this chapter makes two

contributions. The first is a peripheral visual feedback technique to influence respiration

of a desktop user. The second is the results of a study evaluating the effects of PPR

compared to a control condition lacking visual feedback.

Page 68: Augmented Self-Regulation (Moraveji, 2012)

50

3.1 A PERIPHERAL PACED RESPIRATION INTERFACE

The prevailing interaction paradigm for pacing respiration is modal: interfaces

require the focused attention of the user. This is due to the common assumption that to

create a calm state, an interface must require the user to stop their current activity, which,

it is assumed, contains stressors. Given the demonstrated ease with which users can

entrain their breath to visual stimuli, this chapter proposes using short but frequent pacing

sessions that are peripheral, allowing the user to engage in other tasks while concurrently

pacing their respiration. Peripheral pacing requires continuous monitoring and a medium

for intermittent biofeedback. To our knowledge, the work here is the first instance of

peripheral respiration pacing being integrated directly into the desktop.

Our design goal was to evaluate the feasibility of peripherally regulating the

respiration of a desktop computer user. Such a system requires two functions: (1) sensing

user respiration and (2) feedback to pace the respiration across computing tasks.

3.1.1 WIZARD-OF-OZ PROTOTYPE

The research team first created a Wizard-of-Oz prototyping tool to assess

qualitative reactions to different forms of feedback. The researcher would sit behind the

user as they worked at their desk where a researcher would manually observe the user’s

breathe rate due to chest rises. They would approximate the rate and input it using a slider

onto a web form (see Figure 3.1). The pacing prototype installed on the user’s computer

read this data from the webpage and updated its current user breath rate accordingly.

Page 69: Augmented Self-Regulation (Moraveji, 2012)

51

Figure 3.1: The user interface of the web-based Wizard-of-Oz interface that study administrators would use to control the visualization of the desktop-based feedback of a user in real-time. The panel included elements that were

not yet in use (e.g. breath regularity).

An important early consideration was whether feedback should be integrated into

a specific application (e.g., programming code editor, web browser, or productivity

software) or system wide. Because information work involves multiple applications, the

team opted for the latter.

The team used visual pacing because information workers often listen to music

and work in office areas where others are working. The first prototype used a pulsing

circle atop other windows in the top-right screen corner as pacing stimulus.

Using this prototyping method with several users in our lab, two primary issues

were found. First, though seemingly useful, real-time feedback regarding one's breath

rate is highly distracting, as users frequently check the accuracy of the detected rate.

Page 70: Augmented Self-Regulation (Moraveji, 2012)

52

Second, the rate of the pulsing circle was not noticeable when users were deeply engaged

in their work, even when the pulse rate was exaggerated.

3.1.2 USER INTERFACE

The final design does not require any mouse- or keyboard-based interaction and

does not require researcher interaction. Our team implemented three pacing techniques

and a calibration mode. ‘Screen Dim Feedback’ sets the pacing stimuli to dim the entire

screen from near-black to maximum brightness at the target rate. ‘Menu Dim Feedback’

does the same but only to the Mac OS menu bar. ‘Bounce Feedback’ uses an animated

horizontal bar to pace respiration (see below). The calibration mode option toggles the

display of a gray bar whose y-position is controlled in near real-time by the resistance of

the stretch sensor. In calibration mode, the user can determine if the gray bar is indeed

moving up and down as they breathe or if the band requires tightening or repositioning.

Figure 3.2 shows how the study administrator selects the current pacing mode.

Figure 3.2: The user interface of the client was used by the study administrator to select the current peripheral pacing method.

Page 71: Augmented Self-Regulation (Moraveji, 2012)

53

3.1.3 PACING RESPIRATION PERIPHERALLY

Our team implemented and tested a pulsing light technique, two dimming

techniques (screen and menu bar) and one object animation technique. We chose the

animation technique because it is recommended to biofeedback practitioners (Lehrer,

Vaschillo, Vaschillo, 2000) and performed best in our early tests: users could identify

clear end points to inhalation and exhalation in their periphery.

The object animation technique (Figure 3.3) works by moving a screen-wide,

semi-translucent grey bar up and down across the screen, representing inhalation (up) and

exhalation (down). The ratio of up to down is 1:1 but could be adjusted or user-

configurable in future versions. Slow-in and slow-out animation (Lasseter, 1987) is used

to provide smooth movements and aid tracking. The bar moves across the lower third of

the screen to reduce both distraction and distance travelled (using the full screen height

would require a fast and distracting animation).

Page 72: Augmented Self-Regulation (Moraveji, 2012)

54

Figure 3.3: The peripheral paced respiration feedback used an animated, semi-translucent grey bar stretching across the screen. Vertical arrows on the left

indicate the full range of motion.

Each user’s target breath rate is set relative to an individual resting baseline,

rather than a universal target that might require too much effort (Stark, Schnienle, Walter,

Vatil, 2000). The target rate is set to 20% below their baseline to exaggerate the slow

rhythm of resting breath. This particular value was determined through internal testing to

be slow enough to not be distracting but not so slow so as to make it unrealistic for users

to do while engaged in another task.

The software continuously samples sensor data and determines when to display

visual feedback. If the user’s current breath rate is 20% above their resting rate, pacing is

triggered. This value was chosen through iteration with our research team members so as

to reduce frequent triggering as the user’s breath rate fluctuates frequently. The software

will also automatically trigger pacing at least once every six minutes to evaluate its

influence on breath rate. Similar to prior studies, a 2-minute duration was used (Bloch,

Page 73: Augmented Self-Regulation (Moraveji, 2012)

55

Lemeignan, Aguilear, 1991; Sisto et al., 1995). A study was conducted to evaluate the

efficacy and feasibility of using a PPR system to pace users’ respiration to a resting rate

while authentically engaged in information work.

3.2 STUDY

A study was designed to determine if PPR influences user respiration across

computing tasks in a naturalistic manner. Participants were recruited who had existing

work to do (e.g. research, programming, writing). Thirteen university students (9 male, 4

female, mean age=25.5) were recruited from computer science and related disciplines to

participate. They were told they could conduct their existing work during the experiment.

Participants were desired to be genuinely engaged and to work naturally (i.e., switch

windows and tasks as they normally do). Participants were not compensated. According

to a post-study questionnaire, no participants had existing respiratory conditions.

The study design was a counterbalanced, within-subjects experiment in which

participants were exposed to two conditions: (1) no feedback and (2) PPR feedback. As a

control, participants wore the respiration sensor in both conditions.

3.2.1 PROCEDURE

Participants first wore the sensor band and the administrator tightened it. They

were told that the sensor measures their respiration. A short calibration period ensured the

band was positioned accurately. Participants sat in a chair in front of their laptops,

working alone and not speaking. Their posture was not controlled, again to ensure a

naturalistic testing environment. Hence, they were allowed to lean backwards and

forwards in the chair, which can have an effect on respiration rate (Sisto et al., 1995).

Page 74: Augmented Self-Regulation (Moraveji, 2012)

56

Participants first completed a consent form. As in prior studies (Ley, Yelich,

1998), participants were then asked to close their eyes for 3-minutes and relax.

Unbeknownst to participants, baseline data was collected during this relaxation period.

Three participants did not take part in the relaxation period; their baseline was recorded

as they completed a pre-survey.

After recording the baseline, the experiment began with the participant being told

they could start working on their own tasks. They were also told that when the gray bar

appeared, it represents their target breath rate.

To guarantee that pacing would occur at least three times during the PPR

condition (at least once every six minutes), the duration of each condition was 20

minutes. When activated, PPR occurred for 2 minutes.

3.3 RESULTS

Consistent with prior studies, the mean breath rate across conditions was

16.67bpm (SD=4.28) and the breath rate of the relaxation period baseline was 9.33bpm

(SD=5.31). See Figure 3.4 for a visual illustration of the results.

Page 75: Augmented Self-Regulation (Moraveji, 2012)

57

Figure 3.4: (Top) Mean breath rate for the No Feedback and PPR conditions with standard error bars. (Bottom) Mean breath rate during the PPR condition when

PPR was on and off.

Figure 3.4 (top) shows the mean for each condition with 95% confidence

intervals. Paired t-tests were used to compare the means of each condition. A significant

difference was found between no-feedback (M=17.58, SD=4.18) and PPR (M=15.7,

SD=4.49) conditions; t(12)= 3.83, p<0.005. The mean difference between conditions was

1.8bpm.

Figure 4 (bottom) shows the mean breath rate during the PPR condition when

PPR was active or inactive. A paired t-test was conducted to compare the means between

when PPR was activated (M=14.96, SD=4.44) and when it was not (M=17.09, SD=5.25);

t(12)=3.5647, p<0.005. The mean difference was 2.13bpm. The mean proportion of time

that the feedback was activated was 0.60 (SD=0.14).

A paired t-test was used to compare the breath rate in the PPR condition when

feedback was unavailable (M=17.09, SD=5.25) and the breath rate during the no-

Page 76: Augmented Self-Regulation (Moraveji, 2012)

58

feedback condition (M=17.58, SD=4.18); t(12)=0.989, p>0.05. When feedback was not

present, breath rates returned to their working rates.

To illustrate how the breath rate is impacted by PPR, Figure 3.5 depicts the

breath rate of one participant in each of the conditions. The no-feedback condition (top)

shows a relatively consistent, high breath rate. While PPR was active (bottom) the breath

rate decreased.

Figure 3.5: Breath rate for one participant in both no feedback (top) and PPR

(bottom) conditions. Bold (orange) areas indicate where PPR occurred

Using Likert scales from 1 to 5 where higher scores represent greater levels of

‘annoyingness’, participants rated PPR 2.0 (SD=0.87). Concerning how much it

adversely affected productivity, PPR was rated 2.2 (SD=0.6). Lastly, participants gave a

score of 3.7 (SD=1.0) as to how likely they would be to use the software all day long

while working, were the sensor non-invasive.

Page 77: Augmented Self-Regulation (Moraveji, 2012)

59

3.4 DISCUSSION

Our hypothesis that PPR influences user respiration while they are engaged in

naturalistic tasks is supported; the peripheral feedback reduced breath rate significantly.

The 1.9bpm difference is almost identical to the 2bpm interval changes shown by Song

and Lehrer (2003) to correspond with significantly higher HRV amplitude.

Breath rates were observed to return to working levels between pacing instances.

Hence, there was no evidence of persistent rate change. This is a viable area for future

study; cues, social feedback, and game mechanics, among other methods, could help

motivate users to maintain low respiratory rates as they work.

Based on self-report data, PPR feedback was not too distracting and participants

expressed motivation to use PPR for sustained periods in the future. The results motivate

longitudinal research that attempts to motivate, trigger, or incentivize users to pace their

respiration even when pacing is not active. This allows for long-term respiration pacing

that could complement existing methods of respiratory habit-change. The pacing

algorithm could factor in the physiological and work history of the user. Such ‘context-

sensitive biofeedback’ could be used to influence the physiological factors underlying

cognition, and affect.

Quantitative measures of the level of distraction caused by PPR (such as working

while pacing) were not collected in this study. Further, the PPR method used goes beyond

breath rate and implicitly proposes time and duration for inhalation and exhalation, which

is beyond our goals and may have required greater effort than is necessary. This presents

an opportunity for future work to identify the optimal balance between pacing efficacy

and distraction.

Page 78: Augmented Self-Regulation (Moraveji, 2012)

60

The PPR method is but one possibility for visual PPR techniques. Similar

research could be done with auditory techniques that require headphones (or not), haptic

techniques in the mouse, chair, keyboard, or other peripheral, or other visual techniques.

The design space of integrated biofeedback and PPR is large and this chapter outlined the

most important design parameters. It is crucial to note that the goal is not feedback per se,

but rather feedback that is appropriately used to help users change their

psychophysiological state without requiring them to mentally search for control strategies

which would be distracting. That distraction would reduce compliance, being self-

defeating. The goal of systems that augment self-regulation is different from the goal of

enabling self-reflection. In the former, the purpose is to augment state change in real-

time. In the latter, the purpose is more intellectual, to give users feedback and data to

reflect on their state to motivate future state change.

As a tertiary contribution, this study is the first known that quantifies the effect of

naturalistic information work on respiration rate. In our case, resting breath rate was

almost half the working rate. This result highlights the issue of mild but chronic stress

that occurs during information work, and again recommends longitudinal studies.

3.5 CONCLUSION

This chapter presented the design of a peripheral paced respiration technique and

evaluated its efficacy in a naturalistic task environment. It was found that peripheral

pacing significantly reduces breath rate, but these changes are not sustained for the

duration of the tasks. The results recommend further research on using motivational cues

to amplify or sustain the effects. Further, the results motivate more research on

incorporating biofeedback directly into the desktop operating system to complement long

Page 79: Augmented Self-Regulation (Moraveji, 2012)

61

durations of information work in an attempt to reduce stress and increase productivity,

general health and wellness.

Page 80: Augmented Self-Regulation (Moraveji, 2012)

62

CHAPTER 4 BREATHCAST: A STUDY OF SOCIAL INFLUENCE ON BREATH MODIFICATION

The peripheral paced respiration (PPR) technique shown in the previous chapter

overlays a semi-transparent bar across the user’s screen, floating above other windows,

and animates it up and down across the lower third of the screen at a rate contingent upon

the target breath rate and rhythm. The user’s resting breath rate is used as their target rate.

The bar’s visibility is triggered by the user’s current breath rate (i.e., cycles of breathing

per minute). Our aim was to extend this method with social feedback; our study examines

the effect that synchronicity of feedback has on breath rate.

There are many possible methods of adding social activity indicators to PPR. The

Breathcast system is grounded in two design goals:

1. There is a shared goal for Breathcast users to breathe calmly while working, and

2. The more one user breathes calmly, the more they influence others to stay calm

(and vice versa).

In the Breathcast version of the PPR bar, participants are told these principles outright but

in future work these will be communicated via the interface. Avatar icons of other

Page 81: Augmented Self-Regulation (Moraveji, 2012)

63

Breathcast users are displayed atop the PPR bar (see Figure 4.1). The selection of which

icons to display is dictated by the breath performance of those users: the more one

breathes at their resting rate, the more their picture appears on the bars of others. ‘Breath

performance’, here, is defined as the proportion of time one is breathing at one’s

individual resting rate in a recent time range.

Figure 4.1: Breathcast works by intermittently animating a semi-transparent bar across the bottom third of the user’s screen. The inset shows how profile icons

of other Breathcast users are discreetly displayed on the bar. The vertical arrow on the lower right illustrates the range of bar movement. In asynchronous

mode, the bar is blue to aid differentiation.

Rather than offering social presence indicators to simply enhance behavioral

awareness of others, the system attempts to motivate desired breathing patterns by

making the visibility of those indicators contingent upon the performance of the social

agents in question. This introduces a subtle but potentially effective motivational cue that

one’s behavior is being measured and will receive due positive recognition.

Page 82: Augmented Self-Regulation (Moraveji, 2012)

64

Breathcast supports two types of temporality: (1) asynchronous, where users’ past

performance dictates display on your bar and your performance dictates how often your

profile is displayed on future users’ screens, and (2) synchronous, where bi-directional

influence occurs with other concurrent Breathcast users. A third, “coupled” form of

synchrony, in which the actual inhalation and exhalation intervals are conveyed, was not

included because users found such pacing distracting in early tests.

In sum, users are only aware that others are intending to maintain their own

respective resting breath rate. The social feedback received from other users concerns

how well they are able to do this (their “breath performance”), in order to motivate the

current user.

4.1 STUDY

A controlled, within-subjects laboratory study was designed to compare the

effects of synchronous or asynchronous peer influence on PPR efficacy. The study design

is comparable to that shown in the previous chapter.

4.1.1 PARTICIPANTS

Thirteen university students (8 male, 5 female, mean age 26.9) with no respiratory

conditions or prior PPR experience participated, receiving a $5 gift card at the end of the

study. They were recruited by being told they could participate in a study while doing

their own computer work. To sense respiration, participants were fitted with the wired

USB thoracic strain gauge sensor used in prior studies.

Page 83: Augmented Self-Regulation (Moraveji, 2012)

65

4.1.2 PROCEDURE

Participants were asked not to leave their seat for the duration of the study. They

first spent five minutes completing a consent form, adjusting and calibrating the sensor,

and establishing a resting breath rate by closing their eyes for three minutes (as in prior

studies (Ley, 1999; Moraveji et al., 2011). Each user was asked for a profile picture and

username that would represent them on the PPR bars of other participants.

Participants were then told they could start working and asked not to play video

games but to do their other work as usual. The tasks were not controlled in order to

ensure naturalistic work. They first worked for 10 minutes without any PPR feedback in

order to establish a working baseline rate. The subsequent two 20-minute conditions were

counter-balanced and participants were told that when a bar appeared, it would “guide

them back to their resting breath rate.” In actuality, the rate of the bar was slower than

their (often quite fast) resting rate so as to not distract. They were told that the shared

goal was for all students on campus to breathe calmly. In the asynchronous condition,

participants were told that the visible icons were from prior participants, and that their

own performance would dictate their appearance on the bars of future participants. In the

synchronous condition, they were told that icons correspond to other real-time

participants in the study.

At the outset of the synchronous condition, an administrator spoke on their phone,

pretending to coordinate the start of the “real-time” feature. Participants were told that

three other students were doing the study at the same time and therefore the start time

was important.

To control for differentiated feedback, the social data was identically simulated

across participants in each session. The icons of the simulated social agents were all

Page 84: Augmented Self-Regulation (Moraveji, 2012)

66

“unknown similar peers” represented using icons of random Twitter users across genders.

Different bar colors (green and blue) were used to differentiate the conditions but

otherwise feedback was identical. It was hypothesized that synchronous feedback would

motivate users to be more vigilant with their respiratory regulation than with

asynchronous.

The bar was activated when the user’s breath rate was significantly higher than

their resting breath rate, at least once and at most three times every four minutes so as to

avoid distraction. When activated, PPR lasted a random amount of time between 1-2

minutes. This is because the goal was to make the study feel ‘real’ and that it was

automatically detecting the appropriate amount of time to pace the user’s respiration

(over a minimum of 1 minute).

4.2 RESULTS AND DISCUSSION

The mean breath rates for each participant in each condition are shown in Figure

4.2, with a dotplot illustrating the distribution in each condition in Figure 4.3. Both

figures omit a single outlier to enhance readability.

An omnibus repeated measures ANOVA test found mean breath rates differed

significantly across conditions, F(3)=2.88, p<0.05. Planned paired t-tests were then used

to compare the mean breath rates of the different conditions. Replicating prior work

(Moraveji et al., 2011), the mean resting baseline breath rate (M=15.75, SD=4.5) was

lower than that of the working baseline (M=20.17, SD=2.48), t(12)=-5.135, p<0.01.

Page 85: Augmented Self-Regulation (Moraveji, 2012)

67

Figure 4.2: Mean breath rates for participants in each condition.

Mean breath rate during the asynchronous (M=17.78, SD=4.16) condition was

lower than during the working baseline (M=20.17, SD=2.48), t(12)=3.09, p<0.01. The

same was true for synchronous (M=16.91, SD=4.43) feedback, t(12)= 3.868, p<0.01.

More surprising was that both synchronous (t(12)=-0.0873, p=0.3991) and asynchronous

(t(12)=-1.668, p=0.1212) conditions did not differ significantly from the resting baseline.

Page 86: Augmented Self-Regulation (Moraveji, 2012)

68

Figure 4.3: Mean breath rate for each condition. BL=baseline, WBL=working baseline, A=asynchronous, S=synchronous.

The average breath rate in the synchronous feedback condition was 0.87bpm

lower than the asynchronous condition, supporting our primary hypothesis. However, this

small difference is only weakly significant, t(12)=2.156, p=0.052. There are two

hypotheses that might explain the observed difference: a magnify effect, in which breath

rate changes were more pronounced for comparable durations, and a persist effect, which

supposes users maintained a lower breath rate over a longer duration.

Visually inspecting each participant’s data, there is no noticeable trend. The

difference in breath rates when the PPR bar was shown and hidden in each condition was

also inspected. In the asynchronous condition, when PPR guidance was activated it

produced a mean difference of 0.74bpm while in the synchronous condition it was

1.49bpm. This is evidence of a potential magnify effect, albeit small. Figure 4.4 below

shows a time series of a single representative user that illustrates this dynamic.

Page 87: Augmented Self-Regulation (Moraveji, 2012)

69

Figure 4.4: Breath rate of a user with a resting rate of 19.5bpm. PPR occurrences are orange. The working baseline condition (top) saw the breath

rate climb upwards. With synchronous feedback (middle), it decreases noticeably during PPR. Asynchronous feedback (bottom) saw rates drop little and continue

to climb overall.

Survey results show that PPR was occasionally distracting (Likert scale of 6,

M=3.3, SD=1.1) and, when asked who they would expect to use the system with,

participants claimed it would be known friends or perhaps clinical caregivers, yoga

instructors, meditators, or other authoritative people.

Page 88: Augmented Self-Regulation (Moraveji, 2012)

70

This study represents the first known instance of social feedback being used to

influence psychophysiological behavior. Clearly, this study has only demonstrated the

feasibility of utilizing social feedback for influencing breathing behavior – it has not duly

investigated all the ways in which social feedback could be used and for different

purposes. Other social influences may also be explored including indicators of another

user’s breathing patterns over time. Another dimension could be to manipulate the

reputation of the social agent in order to magnify the effect more (e.g., if one’s doctor,

yoga instructor, or favorite celebrity was the social agent in question).

4.3 CONCLUSION

This chapter offered two primary contributions. The first is the interaction design

method for socially motivating respiratory regulation in parallel to information work. The

second is an evaluation of temporality in the system, showing that expectations of

synchronous feedback magnified the effect of social feedback. The implications for

researchers are that continuous physiological behaviors such as respiration can be

socially influenced and that synchronous behavioral feedback from unknown peers, even

at a distance, can enhance the effect. These results recommend further investigation into

design techniques to motivate consistent respiratory change without pacing.

Page 89: Augmented Self-Regulation (Moraveji, 2012)

71

CHAPTER 5 BREATHAWARE: CONTINUOUS INFLUENCE OF SELF-REGULATION DURING INFORMATION WORK

The previous two chapters examined how visual stimuli could prompt users to

pace their respiration rate during information work. However, the system relied on

repeated intermittent visual cues to remind the user to bring their breath rate down. The

drawbacks in that model are (1) that the system is not adaptive to task valence, (2) the

user is not relying upon their innate self-regulation, and (3) the user only has negative

motivation (in the form of keeping the PPR bar hidden) to maintain low breathing rates.

This chapter investigates different methods of augmenting self-regulation without

explicitly pacing respiration. It contains three contributions. The first is the set of 10

design principles, motivated by the literature on stress and stressors, which describe

necessary attributes of a system that effectively augments a user’s respiratory self-

regulatory processes. The second is the design, implementation, and exploratory

evaluation of a prototype system that realizes these design principles. The third

contribution is the outline of a research agenda and accompanying research questions for

Page 90: Augmented Self-Regulation (Moraveji, 2012)

72

further study of interactive systems that influence and augment our fundamental innate

self-regulatory processes.

5.1 DESIGN PRINCIPLES

Based on a review of the literature, related tools, and our own experience

designing interfaces and getting user feedback from previous chapters, a set of 10

generalizable design principles (DPs) is presented, each responsive to a challenge of

influencing a user’s psychophysiological state using interactive technology. We refer to

these systems as autonomic interaction systems (AISs) and, when used in these

principles, generally refers specifically to interaction with respiration. The challenge is

shown in italics and the rationale follows. These principles are not meant to educate or

motivate the user to engage in breath awareness or regulation. They assume the user is

motivated to a significant degree and that the use of the technology is only meant to

complement their own self-regulatory processes.

A caveat about these principles is that they are principles, not empirically

evaluated design heuristics. That is, they are drawn from research on self-regulation,

breath regulation for panic disorder, the yogic and qigong practices, human-computer

interaction, and contemplative science. As such, they contain a certain degree of

speculation and, though we attempted to make them as objective as possible, may contain

bias towards the intention to influence users in the direction of calm states or neutral

arousal. A summary of all 10 principles is listed in Table 5.1 below.

Page 91: Augmented Self-Regulation (Moraveji, 2012)

73

# Design Principle Challenge

1 Accommodate different levels of attention

One’s breath is ‘always-on’ but shouldn’t be all-consuming.

2 Sustain motivation Because we are always breathing, regular attention and motivation to regulate it can be easily dismissed as unimportant.

3 Demonstrate desired patterns of breathing

Few people have a good sense of their desirable breathing pattern at work and at rest.

4 Personalize the feedback Desirable breathing patterns vary widely between users.

5 Reinforce the relationship between the breath and the body

The relationship between breathing patterns and the rest of the body (such as muscles) is not often apparent

6 Avoid exasperating stress with negative feedback

Negative feedback can cause feelings of inadequacy, competition, or resentment.

7 Develop awareness at different timescales

Breath awareness is different at different timescales.

8 Encourage internal self-assessments

Technology could create dependency, hindering true self-regulation

9 Consider secondary components of respiration

Sighing, wheezing, and related events are important contributors to improved self-awareness.

10 Protect the privacy of the breath

One’s breathing patterns are intimate and private and should be treated as such.

Table 5.1: 10 design principles for interactive systems aiming to influence respiratory self-regulation.

5.1.1 ACCOMMODATE DIFFERENT LEVELS OF ATTENTION

One’s breath is ‘always-on’ but shouldn’t be all-consuming. The goal of

respiratory self-regulation is not to have users constantly watching their breath and ignore

their primary tasks. Instead, systems should support the user in returning from a

heightened state of arousal in a reasonable time frame or to refrain from repeated,

frequent sympathetic activation (Obradovi!, Boyce, in press). This means that users may

not pay direct attention to their breath for spans of time where they are focused on their

task yet in a calm manner. As such, AISs must support various levels of attention:

sometimes full attention, sometimes partially aware and looking for feedback in intervals,

Page 92: Augmented Self-Regulation (Moraveji, 2012)

74

and other times attempting to maintain a certain state while focused on another task (e.g.,

writing an essay, creating a spreadsheet, or manipulating an image).

‘Level of attention’ encompasses granularity of data; precise values may not be

necessary to communicate their importance to users. Gross, relative values may have the

highest impact and require less cognitive load to process. Providing relative values may

require the system have some sort of baseline, whose definition should be communicated

to user (e.g., compared to their individual resting rate, yesterday’s rate, similar peers,

prior best, and so on). Level indicators can also be relative, and descriptions of the levels

should make this apparent. Social comparison is an oft-used method of influence but, in

the context of self-regulation, should be used in a manner that supports self-regulation

and understanding of one’s own state and behavior. It should not compromise one’s

privacy (see 5.1.10).

This principle could be evaluated by identifying what kinds of information users

are able to glean from what kinds of feedback. Using different feedback conditions, an

evaluator could interview participants to ascertain the accuracy of that user’s self-

awareness as augmented by the feedback. Motivation and task would need to be

controlled in such a study.

5.1.2 SUSTAIN MOTIVATION

Because we are always breathing, regular attention and motivation to regulate it

can be easily dismissed as unimportant. It is easy to ignore or not notice the effects and

implications of changing one’s breathing patterns or awareness of breath. Unlike

exercise, a user may not notice outwardly visible effects that others can comment on (or

that a mirror can make salient). Instead, maintaining a calm state or disposition is

Page 93: Augmented Self-Regulation (Moraveji, 2012)

75

primarily an internally visible change, the effects of which may include greater subjective

self-awareness, less agitation, less cognitive rumination, greater positive affect, and

greater awareness of issues that were invisible before such as minute physical

discomforts. Sustaining one’s motivation to develop and practice breath awareness and

self-regulation can be done in myriad ways and, in this domain in particular, users have

very different levels of motivation. Some users with consistently high levels of stress may

consider that the norm and not even realize the benefits of calm states. They may equate

stress with productivity and calm with sleep, a non-productive state that they address

nightly. In truth, many people dedicate their lives to practicing breath awareness,

teaching others about the breath, and using the breath for personal growth. AISs may

assume high levels of motivation or attempt to inform users about the various reasons to

develop their self-regulatory ability. AISs may appeal to the user’s desire for greater

health, productivity, more positive affect, reduced negative affect, and so on. Such

systems must also take care to not overwhelm users with information and triggers to pay

attention to their breathing; it requires a delicate balance of bringing attention back to a

user’s existing motivation to enhance their own self-regulation.

5.1.3 DEMONSTRATE DESIRED PATTERNS OF BREATH

Few people have a good sense of their desirable breathing pattern while at work

compared to at rest. Most individuals in modern society either do not have a good sense

of what a calm breathing pattern is or how to maintain it in parallel with their

participation in other tasks. Many have never taken part in breathing exercises or

dedicated time to consciously experiencing different breathing patterns. It is important

that any system be able to guide users to experience calm patterns, such as is often done

Page 94: Augmented Self-Regulation (Moraveji, 2012)

76

in current commercial solutions reviewed in section 2.3. Even users who have

experienced calm breathing patterns may not know how to incorporate them into a

workday. Due to the nature of psychophysiological state itself influencing attention and

psychology, it is difficult to maintain cognitive awareness of a reference point because

cognition itself is affected by aberrant breathing patterns. One’s respiratory state is not

normally quantified in the same way that heart rate is, making respiratory state less

tangible.

Further, the breath has multiple characteristics including depth and regularity,

further complicating the process of remembering one’s neutral patterns of calm breathing.

Simply measuring one’s resting rate is an imprecise practice because it does not define

explicitly what ‘rest’ means. ‘Rest’ normally refers to physical rest but breathing patterns

are influenced by both physical and mental activity. For example, watching a video, even

a calm one, influences the state of the mind, which influences the state of the breath.

Even when closing one’s eyes to meditate, one finds that their breathing patterns and

heart rate rhythms fluctuate consistently with their emotional state.

5.1.4 PERSONALIZE FEEDBACK

Desirable breathing patterns vary widely between individuals. Resting breathing

patterns vary significantly between individuals (Ley, 1999; van Diest et al., 2001) and

depending on their current state of mind (Porges, Doussard-Roosevelt, Malit, 1994;

Biotin, Brigade, Witnesses, 1994), physiological health (Grossman, 1983), and physical

state (e.g., posture). Interactive systems will need to take these differences into account,

rather than guiding users towards a non-existent ‘universally optimal’ breathing pattern.

Further, AISs should aim to personalize the strictness with which they trigger changes in

Page 95: Augmented Self-Regulation (Moraveji, 2012)

77

behavior. That is, users have very different desires for stringency based on their personal

goals, environmental contexts, and the ebb and flow of work intensity. Feedback while at

home may be very different than feedback at work. Likewise, feedback in the early part

of the workday may be different from the latter.

5.1.5 REINFORCE THE RELATIONSHIP BETWEEN BREATH AND BODY

The relationship between breathing patterns, the body, and the mind is not often

apparent. Users have very different motivations for wanting to develop self-regulation.

However, any AIS should ideally develop the user’s innate self-regulatory ability to

avoid over-dependence and to develop that user’s experiential understanding of the

nature of the mind-body relationship. Developing a user’s understanding and experience

of this relationship is among the most profound intentions that a tool can have for a

human being. An ideal AIS system would reinforce the relationship between breathing

patterns and affect, ability to focus, and other aspects of the human experience (Biotin,

Brigade, Witnesses, 1994; van Diest et al., 2001). For example, systems could uncover

when stressful events lead to halted or hyperactive breathing. In the other direction, a

system could show how a change in breathing patterns brought the user from fragmented

to focused attention. Awareness of any component of this interconnected system could

strengthen self-regulation and self-awareness.

5.1.6 AVOID EXASPERATING STRESS WITH NEGATIVE FEEDBACK

Negative feedback can exasperate the problem. Environmental stressors (i.e., real

physical threats to one’s safety) account for only a small proportion of stress typical

experienced by a modern human being in the West. However, many individuals see

external stimuli as stress-inducing because it triggers their own anxieties or needs to

Page 96: Augmented Self-Regulation (Moraveji, 2012)

78

comply with external pressures. As a result, interactive technologies are often seen as

stress-inducing.

Feelings of social evaluative threat from a social agent, even a piece of

technology, can do the same. Any interactive system that gives feedback on performance

has the potential of delivering what can be construed as negative reinforcement. That is,

the feedback could make explicit bad breathing performance or, worse, falsely diagnose

what was in reality a calming breathing pattern (false negatives). In the domain in

question, this is particularly dangerous because negative feedback could cause additional

stress, exasperating rather than mitigating the problem. Even accurate positive feedback

(e.g., “Great job, you just had 23 minutes of calm.”) can interrupt a calm state by

triggering feelings of social evaluation and so must be designed in light of this.

5.1.7 DEVELOP AWARENESS AT DIFFERENT TIMESCALES

Breath awareness is different at different timescales. Breathing patterns can be

observed at different levels of analysis: e.g., second-to-second, recent past, portions of

the day, day, few days, week, month, segment of year, year, life stage, and lifetime. There

are different reasons to focus on different levels of analysis at different times: they

influence motivation differently, reinforce the relationship between mind and body

differently, and help diagnose overarching patterns differently. Because changes in the

breathing patterns are constantly occurring, it may be possible to become ‘over-vigilant’

and hyper-focused on moment-to-moment changes. Testing the effects of different levels

of analysis is non-trivial because users must be wearing the sensor for such durations of

time in order to make the data useful and ensure they take the feedback seriously. For this

Page 97: Augmented Self-Regulation (Moraveji, 2012)

79

purpose, simulated data can be used to evaluate impact on users if it is explained that they

should suspend disbelief.

5.1.8 ENCOURAGE INTERNAL SELF-ASSESSMENTS

Technology could create dependency, hindering true self-regulation. Any AIS

system faces the risk of creating an unhealthy dependency between the user and the

system. This is a well-known problem in “scaffolding” human performances in education

(Pea, 2004). Though humans can be conceptualized as co-evolving with technology, as

described by Douglas Engelbart (Bardini, 2000), self-regulation is ultimately an

introspective process independent of external tools (which serve only to augment and

strengthen one’s innate self-regulatory process). It is therefore important that the system

take measures to encourage an introspective or reflective internal self-assessment.

From relevant literature on stress-reduction techniques and programs, we

identified a practice of an internal awareness ‘check-in’ where one assesses their own

state to determine their own heart rate, breath rate, cognitive state, affective state, muscle

tension, or other indicators of sympathetic or parasympathetic activity. Different

techniques or exercises use different phrases to describe this process: centering, stress

check, cool-down, spot-check, etc. The general goal is to assess one’s state on a regular

but non-disruptive basis; the simple process of assessment is usually enough to influence

one’s state. This practice should complement tool-based feedback, which brings the

user’s attention out from their bodies and onto the external feedback display. One

difficulty in designing for internal self-assessment lies in evaluation because, by

definition, they often do not involve system interaction.

Page 98: Augmented Self-Regulation (Moraveji, 2012)

80

5.1.9 CONSIDER SECONDARY COMPONENTS OF RESPIRATION

Sighing, wheezing, and related events are important contributors to improved

self-awareness. A number of relevant phenomena or events influence one’s breathing

patterns such as breath-holding during work tasks (Stone, 2008), sighing (found to

increase during high cognitive load tasks (Vlemincx et al., 2011)), sleep apnea during

sleep, and speaking. These events often go unnoticed and their impact on overall

breathing patterns unacknowledged. Note these are not studied in this dissertation though

they are measurable and potentially significant for learning purposes. They can act as

punctuated events to give practical texture to more persistent feedback such as breathing

rate and depth. They can even act as primary feedback mechanisms, making their

detection easy to subjectively validate.

5.1.10 PROTECT THE PRIVACY OF BREATH

One’s breathing patterns are intimate and private and should be treated as such.

Like many other aspects of life, if privacy is broached in what was once a private,

intimate aspect of one’s life, problems can arise. Breathing patterns of others can be

interpreted in undesirable ways, requiring a means of protecting one’s privacy. The

patterns could be used by insurance companies, employers, interviewers, doctors,

therapists, romantic partners, and other social agents as a reflection of the user’s mind,

health, emotional patterns, and cognitive load. As in other aspects of data privacy, users

may be willing to relinquish data privacy to garner well-documented benefits, but such

trade-offs require study outside the scope of this dissertation.

Page 99: Augmented Self-Regulation (Moraveji, 2012)

81

5.2 PROTOTYPE DESIGN

The prototype described here instantiates the aforementioned design principles

using an iterative, user-centered design research process. The prototype, Breathaware, is

meant to help users develop awareness of their breathing patterns while doing work on a

laptop or desktop. This differs significantly from PPR, whose aim was to visually pace

users to a specific breath rate. An assumption of the present prototype is that the user is

already aware of the benefits of maintaining a calm state. The prototype system has three

components the user interacts with: sensor, client, and social network.

To date, methods of influencing physiological change relied on traditional modal

biofeedback, requiring the user’s full attention. A stationary, as opposed to mobile,

context was chosen for multiple reasons. First, it is a more controlled environment.

Second, it was expected that the technology would be somewhat easier to evaluate. Third,

it allows experimentation with both interruptive and peripheral means of influence.

Fourth, sensor readings of breath rate during user verbalizations are notoriously difficult,

so a context was chosen where users do not regularly speak.

5.2.1 CLIENT

The software client, written in Objective-C for OSX, is a system tray application.

It reads data from a USB receiver and creates a log of all respiration sensor readings.

Because users are expected to go out of range of the computer (e.g. to attend meetings

and meals), the client will automatically detect signal loss, pause, then restart sensor

recordings. The user interface and features of the client are described in the next section.

All log data is stored both on the local drive (retaining the user’s respiratory

history while the sensor is on) and is uploaded to a centralized web repository, associated

Page 100: Augmented Self-Regulation (Moraveji, 2012)

82

with a unique identifier for each user (email address). A separate process is launched

every 3 minutes to upload both the physiological data and client events via HTTP POST.

The client also stores an adjustable baseline breath rate to ensure that it is tailored

not only to one’s resting rate, but also to their different states and goals (addressing DP4).

The default baseline is 15bpm but it can be set at any time by toggling the record button

on and then off.

5.2.2 SOCIAL NETWORK

The web repository (Figure 5.3), Breathcast, is accessed via a client menu item or

by web browser. An online profile stores their log data and those of social connections. In

future versions, the user will be able to control their privacy settings at a more granular

level.

5.3 INTERACTION DESIGN

This section describes the strategies used to meet the design principles identified

earlier. Because the research here is focused on stationary users at work, it is crucial that

such strategies are usable in a longitudinal setting in parallel to existing work, unlike

dedicated tools or exercises used in isolation. The goal of the prototype was to help users

understand their breathing patterns and support them to breathe in their desired manner.

This section outlines the features of Breathaware and the design principle(s) they address.

Principle 9 is not addressed, as algorithms for secondary characteristics were not included

in the prototype.

Page 101: Augmented Self-Regulation (Moraveji, 2012)

83

Figure 5.1: The BreathTray and its drop-down menu in two states: sensor disconnected (left) and connected (right). The user’s current breath rate is

displayed on the user’s system tray.

5.3.1 BREATH RATE (IMMEDIATE) – DP3, DP7

A core feature is displaying the user’s current breath rate in near real-time, available

at a glance. This is done by using the drop-down menu itself as the feedback panel

(Figure 5.1). The update interval can be adjusted manually but is set to every second.

Further study is required to ascertain an optimal interval update rate to reduce distraction

while maintaining desired impact on breath rate.

As the active application changes on OSX, system tray icons can be ‘pushed off’ and

made invisible. We wanted to reduce this occurrence as much as possible so that users are

able to see their current status while in any application. The system tray displays three

fields at all times:

• Breath rate: The user’s current breath rate, in units of bpm (breaths per minute).

• Difference from user’s baseline breath rate: The user’s current breath rate in

relation to the current baseline (breath rate while at rest). Here, “131.2%” means

the user is breathing 31.2% faster than their personal resting rate (recorded earlier

or set manually).

Page 102: Augmented Self-Regulation (Moraveji, 2012)

84

• Calm points: an indicator of past performance (described below)

After several iterations, these three fields above were considered most interesting to

display and display technique was the smallest way to display them without taking too

much room on the system tray. This display is what we call the BreathTray. When the

sensor is disconnected or is not in use, the BreathTray reverts to “<Breathe>” to indicate

to the user that their breath rate is not currently being detected.

5.3.2 BREATH RATE (DAILY) – DP5, DP7

Another user need identified is to be aware of trends that impact one’s daily life

(e.g., stressful situations or activities). This can help users develop an awareness of

particularly stressful or calm moments in their day to develop their self-awareness for the

future. To this end, the user can select “Today’s Highs and Lows” from the drop-down

menu (shown in Figure 5.2) to trigger the display of a window that displays the three

highest and lowest breath rates (consistent within 1bpm for at least 1min). The client goes

a step further to associate a context of work (represented by a screenshot of the entire

screen) and the time of day for each high and low. This strategy aims to strengthen the

mental association between one’s respiratory state and work activities. This can lead to

insights such as “I find that I begin to stress out when I check email too often – or vice

versa”.

Page 103: Augmented Self-Regulation (Moraveji, 2012)

85

Figure 5.2: ‘Today’s Highs and Lows’ shows three desktop screenshots from each category with relevant metadata (breath rate, time of day). Here, the user’s breath rate was highest when working in email and on a presentation. It was

lowest when viewing their calendar and viewing a document.

5.3.3 BREATH RATE (IMMEDIATE BUT RELATIVE) – DP3, DP7

Rather than displaying to the user a decimal value, another means is using color to

communicate a general relative breath rate frequency. This enables the user to have a

peripheral view of his or her breath rate without looking at precise percentage values.

When this function is toggled on, the entire menu text is displayed in the color according

to the proportion of the baseline (blue: at or below baseline, red: above baseline). Other

levels (e.g., particularly calm or stressed) could easily be added by using other colors or

different intensities of the chosen colors.

Page 104: Augmented Self-Regulation (Moraveji, 2012)

86

5.3.4 BREATH RATE (LONGITUDINAL) – DP7

The last way to inspect one’s breathing patterns by breath rate is to use the

website to inspect one’s profile. This allows one to see a cumulative average (Figure

5.3, top), a day-by-day breakdown in list form, and a visualization of one’s history data

in a line graph. This also helps realize DP7.

Figure 5.3: The user’s Breath.fm profile for an imaginary user, ‘KKP’. The top shows their overall data including name, last activity update, total calm

points accumulated, mean baseline, and mean BPM. The area below shows their activity stream with event notifications updated in real-time: positive and

negative reinforcement messages and milestone images.

5.3.5 CALM POINTS – DP2

Displayed continuously in the system tray alongside the user’s current breath rate

is the current number of calm points they have accumulated (see Figure 5.1). It is a daily

running total, resetting each morning to zero. A user gets points by breathing calmly; the

calmer they breathe relative to their baseline, the bigger the increment. Points are never

taken away, they are always positive. One’s point total at any time is, therefore, a

function of how long they’ve been at the computer, wearing the sensor, and breathing

Page 105: Augmented Self-Regulation (Moraveji, 2012)

87

calmly. Breathing at one’s resting rate for 1min adds 2 points. The increment increases

linearly as the user’s breath rate decreases. Because points are accrued while doing any

activity at the computer, they can be accrued while working, watching videos, chatting

with friends, reading, or any computer activity.

5.3.6 DAILY MILESTONES – DP2

Based on the number of calm points achieved, a window is displayed atop all

other open windows at the 40, 80, 120, 160, 200, 240, 300, 400, 600, 800, and 1000-point

milestone markers. At each milestone, one of 50 different nature-inspired, calming

images is selected at random and displayed in a large window, above all other windows

(Figure 5.4). The user must close the window manually to hide the image. The sequence

of images is never repeated.

Figure 5.4: Calm point milestones. The desktop of a user who achieved the 80-point milestone. The inspiring images are always randomized as an attempt to

create anticipation for the different milestones.

Page 106: Augmented Self-Regulation (Moraveji, 2012)

88

5.3.7 ENCOURAGING MESSAGES – DP2

Giving the user positive feedback when they have been breathing calmly is useful

because (a) the user may not realize it, (b) if they do realize it, they may not know for

how long, and (c) the knowledge that durations (not only instances) of calm breathing are

acknowledged can be motivating. Figure X depicts one of the two types of positive

notification windows included in this prototype. The user can manually dismiss the

notifications as well.

Modeless ‘toast’ notification windows (see Figure 5.5) were implemented that

fade away after five seconds (the default duration in Growl, a popular notification

management system). The notifications appear in the bottom-right of the screen. Future

versions would use standard notification practices, customizable by the user.

• Notable duration: “Very cool – 7:16 of calm.” This contains two elements: a

congratulatory message and the duration that they user stayed below their

baseline. An interesting motivational mechanism included is that the user sees this

positively reinforcement message at an opportune moment: when they are

breathing above their baseline.

• Exceptional duration: “Moment of Zen. 12:37 of calm breathing.” When the user

breathes calmly for at least 10min, a different category of positive reinforcement

is denoted which contains a different congratulatory message to motivate users to

reach higher (and breathe calmer).

Page 107: Augmented Self-Regulation (Moraveji, 2012)

89

Figure 5.5: The prototype system showing two types of notifications in the lower-right corner of the screen: (a) Left, positive feedback gives the user a congratulatory message and a duration for which they were breathing relatively calmly. (b) Right, a cautionary message tells users how long they have been

breathing relatively fast.

5.3.8 CHECK-IN – DP8

A “Check-in” button on the system tray dropdown (see Figure 5.1) that simply

reports back the user’s breath rate averaged over a 30-second window. This allows the

user to communicate their intention to conduct an internal self-assessment, even while

they continue working. Our original design had the breath rate over the 30-seconds

becoming the user’s new baseline, but this proved too onerous in internal testing. In sum,

pressing the “Check-in” button could connote one of several different intentions such as:

• “I am coming back from a stressful state.”

• “I have been feeling calm recently.”

• “I have been feeling stressed recently but would like to be calm.”

There are no extrinsic motivators for pressing the check-in button. It is meant to draw

only on a user’s latent intention to maintain a calm state as they work.

Page 108: Augmented Self-Regulation (Moraveji, 2012)

90

5.3.9 CAUTIONARY FEEDBACK – DP6

The notion of giving users negative reinforcement around stress can seem

counterproductive. However, reminders about sustained stress can be motivating for

some users if they are not penalizing. As such, two types of notifications exist in the

prototype:

• Notable duration (Figure X): “Caution – watch your breath. Your breath rate is

17.2 (115%)!” A cautionary message is displayed alongside the duration for

which the user has been breathing over their baseline for at least 3min. A soothing

audible chime is played when the window appears.

• Exceptional duration: “Moment of Stress: 7:18 of especially rapid breathing

detected.” The user’s long duration of fast breathing is called a ‘Moment of

Stress’.

5.3.10 ACTIVITY STREAM – DP2, DP10

The events that one experiences with the client are all stored on their profile and,

at present, are public to other users of the (private and closed) system on the social

network (see Figure 5.3). Cautionary and encouraging feedback, and milestones

achieved, are all listed with their corresponding icon and background color to aid

delineation. The knowledge that one’s activity stream is part of one’s profile could

motivate users to remain engaged in the system. Future iterations would support

manipulation of privacy settings for different types of events. Screenshots of the user’s

desktop are not stored in their web profile.

One intention of the website is to enable people to develop reputations for being

‘always green’ and possibly establish a following by other users, as in Twitter or other

Page 109: Augmented Self-Regulation (Moraveji, 2012)

91

social media. In this way, one’s state of mind and body can act as a physiological status

indicator alongside traditional online status indicators (i.e., available, busy, away).

Clearly, there is a need for fine-grained data privacy controls at this level as well (DP10).

5.3.11 BUDDY LIST – DP10

As multiple users could be using the system simultaneously, one research goal is

to experiment with methods of motivating calm breathing socially. The first strategy is to

show other users who are using the system and their respective breath rates precisely and

relative to their baseline (if they opted in to revealing that data to you – DP10). Figure

5.6 shows the drop-down menu with other users online. Their username (email address)

is displayed, alongside the number of points they currently have and their most recently

reported breath rate.

In its current incarnation, all users of the system are in a fully connected graph.

This creates a scenario that actually violates DP10 because users have no privacy over

their respiration data (and this information is told to them upon using the prototype). This

is only a temporary solution as we experiment with the impact of social influence and

before privacy settings and the ability to construct one’s own social graph or inherit it

from other social networks are implemented.

Page 110: Augmented Self-Regulation (Moraveji, 2012)

92

Figure 5.6: When other users have recently logged in and had data sent to the web repository, the drop-down menu also doubles as the location of the buddy list. Usernames, current point values, and last recorded breath rates are

displayed.

5.3.12 RE-RECORD AND MODIFY BASELINE – DP3, DP4

At any time, the user can re-record (by toggling the ‘Record Baseline’ button) or

manually modify the current baseline (by pressing ‘Set New Baseline’) that points and

notifications are in terms of. This allows users to adapt the system to their desired

feedback level or arousal level. This is one way to support goal-setting in that a user can

decide to record a low baseline if they want to ensure they stay especially calm while

working. This could happen in response to a stressful episode, headache, pain, or training

session.

5.4 TEST DEPLOYMENT

A small longitudinal test deployment of our prototype system was implemented to

get feedback that would guide further design iteration and to shed light on how relevant

the design principles are. Two male users self-identified as ‘early adopters’, aged 26 and

Page 111: Augmented Self-Regulation (Moraveji, 2012)

93

28 (User A and B, respectively), used the system for a total of 34 hours over 3 and 4.5

workdays, respectively. Each worked primarily at their laptop computer during the day.

At the culmination of the deployment, participants were individually interviewed. The

buddy list feature was deactivated for the purposes of the deployment and users were not

allowed to view the data of other users so the website could be used only to view one’s

own past data (DP10).

5.4.1 RESULTS AND FEEDBACK

It quickly became apparent that the two users represented two very different

personas. What was most interesting about this deployment was the way in which these

different users interacted with the system over time. User A had an average breath rate of

18.3 (SD=4.91) while User B’s was 24.0 (SD=5.1). The users gained a very different

number of calm points: 402 for User A and 41 for User B. User B received 74 cautionary

messages while User A received 0. User A received 22 notifications with positive

reinforcement while user B received 5. User B hit 2 milestones while User A reached 14.

The baseline was adjusted/ recorded four times by User B (ranging from 18.78 to 28.75)

and only once by User A (Baseline=18.81). The longest non-‘Zen’ duration of calm

breathing sensed was 8.8 (by User A) and there was 1 ‘Moment of Zen’ total (User A).

In some ways, the users acted similarly. They checked in very little (4 times total)

and visited the website only once each. Each user also used the check-in feature 3 times.

User A viewed their daily highs and lows a total of 5 times while User B viewed it twice.

After the users returned their sensors, both users reported a sense of breath

awareness that they did not have previously. User A: “I didn't think about it [my breath]

much before. The only time I used to think about it was when I'd get nervous before a

Page 112: Augmented Self-Regulation (Moraveji, 2012)

94

presentation or something like that. Now I use it to keep stress down so it doesn’t build

up.” This is an indicator of initial success of the prototype’s primary intent. Both

participants also noted that their awareness would be stronger and last longer if they used

the prototype for a longer duration.

From user interviews it was discovered that User A was already motivated to keep

their breath rate low throughout the day while User B was more interested in the

relationship between breath rate and task. This difference speaks to the fact that the

application does not clearly impose a preferred model of use (for better or for worse).

User A developed a goal to maintain a breath rate at or below their baseline. He

reported that when he received a notification informing him of a notable duration of

calm, it motivated him to maintain that low breath rate. This positive reinforcement had

the intended effect given that it appears when the breath rate climbs above the baseline

and motivated the user to bring (and keep) it down.

User B found their resting breath rate to be consistently lower than the breath rate

they felt comfortable working at. As such, he accrued very few points each day until he

“cheated” by raising his baseline manually. This user felt he had a significantly different

resting and working rate and this was not a problem to him.

User A found the milestone intervals to be “unpredictable” but “it didn’t bother

[him]”. The unpredictability speaks to the general feeling among the users that the

algorithm for increasing calm points was not clear. This was by design: we wanted to

avoid attempts at gaming the system and distractions from the user’s task. Both users

voiced appreciation for the inspirational imagery shown at milestones (one even took

several screenshots of those images as an expression of pride).

Page 113: Augmented Self-Regulation (Moraveji, 2012)

95

Both users noted the system was not disruptive despite the regular desktop

notifications. Regarding negative feedback, both users were not discouraged by it (even

User B). Rather, they felt it was “more of a reminder”. We also noted the suggestion to

see one’s breath rate at different times of day and in different contexts (e.g., morning,

Fridays, at home). This deeper level of analytics could come by opening up a user’s data

to third party components via an API, being sure to heed privacy control settings.

5.5 DISCUSSION AND IMPLICATIONS FOR RESEARCH

Breathaware is the first system we are aware of whose goal is to develop

respiratory self-awareness in parallel to other tasks. The test deployment confirmed it met

its goals and motivates further inquiry into this area of research.

We can use insights from this deployment to iterate further. First, the system’s

simple threshold values and lack of adaptation to user performance means users who do

not perform in an expected fashion may very well ignore feedback over time. Attempts at

motivating users who are not breathing at their resting rate cannot rely on cautionary

messages alone. Perhaps the baseline must be automatically adjusted and respiratory

consistency encouraged. To this end, future improvements can draw upon prior research

in adaptive feedback and pedagogical agents that adapt to user performance and adjust

difficulty accordingly.

Second, setting breathing goals is currently not straightforward and it may be best

to add a goal-setting feature.

Third, the user cannot be expected to remember to ‘check-in’; perhaps introducing

a random notification (if the user’s breath rate has high variance) would be a welcome

feature.

Page 114: Augmented Self-Regulation (Moraveji, 2012)

96

Fourth, the system provided a greater variety of feedback to the user who was

more consistently calm. This positive bias was an artifact of our desire to avoid

extraneous negative reinforcement. However, it is clear the cautionary messages did not

have the intended effect and perhaps violated DP6, causing the user to ignore them.

Finally, the system should not assume that users always want to keep their breath

rate low; identifying patterns and relationships between breath rate and activity may be

more interesting. This would strengthen the user’s own self-regulative awareness, rather

than only enforcing a particular type of breathing.

The primary contribution of this study is the design of a system that augments

respiratory self-regulation by developing breath-awareness, based on design principles

inferred from a review of relevant literature and tools. Further iteration and evaluation of

individual components of the system will shed further light on those design principles

while motivating further research into systems that augment human self-regulation.

These results triggered a number of research questions listed below to motivate

other researchers and designers to answer pressing questions into techniques for most

effectively augmenting the human self-regulatory system.

• How (if at all) should adaptive feedback be integrated?

• What is the role of goal-setting in such systems without distracting users from

their task?

• What is the role of negative feedback –in both rare and regular cases?

• How can the system help users make educated choices about what baseline value

to choose?

Page 115: Augmented Self-Regulation (Moraveji, 2012)

97

• What privacy settings are optimal for protecting user privacy while still

promoting social motivation?

• How should the system adapt to user context and user preference? E.g., should

systems differentiate between different work activities?

• What positive motivation exists beyond the novelty of ‘points’?

• For mobile contexts, how can systems provide feedback without constant

interruption (which may be while driving or handling heavy machinery)?

• What are the optimal sensors to be used for sensing the regulatory processes

(including respiration)?

• What other processes are most interesting and productive to experiment with (e.g.

muscle tension)?

• What are the most important patterns to display to users that would best influence

their self-regulatory practices? E.g., differentiation by time period, application

use, time of week/year?

• How best to determine for any given system user, what is the optimal update rate

of peripheral biofeedback for them such that it provides maximum benefit and

influence while minimizing distraction?

5.6 CONCLUSION

There is no question that society is turning its attention to methods and techniques

for sustainable living in an environmental sense and an individual health and productivity

sense. As a result, interactive tools are emerging that hope to influence user state in such

a way as to produce the psychophysiological states appropriate for productive work. The

Page 116: Augmented Self-Regulation (Moraveji, 2012)

98

focus of this chapter was to investigate this space by creating and testing a prototype that

aims to reduce stress by increasing the prevalence of calm states through breath

regulation. Also guiding this examination are 10 design principles for designing such

applications, which yielded several research questions to provoke further research. The

number of such research programs will likely increase in the coming years as

corporations and governments are economically incentivized to ensure optimal

psychophysiological states in their members more frequently.

Page 117: Augmented Self-Regulation (Moraveji, 2012)

99

CHAPTER 6 BREATHTRAY: CONTINUOUS RESPIRATORY FEEDBACK AND ITS EFFECTS ON COGNITIVE PERFORMANCE

The previous chapter outlined the design space and research questions for methods

of designing self-regulatory feedback directly into the desktop. One of the fundamental

questions to this research agenda is the effect that such feedback has on the user’s breath

and their work; this chapter addresses that question.

Methods to augment user self-regulation during information work have so far relied

on notifications and explicit prompting of physiological change. Peripheral paced

respiration intermittently paces respiration according to the user’s individual resting rate.

Earlier, it was shown that this method had users rely on external prompts as reminders

rather than strengthening or amplifying their own internal self-regulatory habits.

This chapter presents an empirical evaluation of an interaction design technique to

motivate and incentivize self-regulation without explicit prompts and pacing. Further, this

study uses controlled tasks that induce cognitive load to evaluate the effect that such

feedback, and respiratory self-regulation itself, has on cognitive tasks.

Page 118: Augmented Self-Regulation (Moraveji, 2012)

100

Our interest is in augmenting the ability of information workers and students to

self-regulate despite cognitive demands. To that end, the research question asked was,

“Can continuous feedback be used for ongoing self-regulation and, if so, does it come at

the cost of task performance from the additional cognitive load?” This chapter makes two

contributions by (1) studying the extent to which peripheral respiratory feedback and

monitoring impacts respiratory patterns as the user is engaged in other tasks and when

compared to motivation alone and (2) examining the extent to which these changes are

accompanied by impacts on cognitive performance on the user’s tasks.

6.1 BREATHTRAY DESIGN

BreathTray continuously displays respiratory feedback in the system tray,

implemented here on Mac OSX. It contains four feedback elements: (1) one’s breath rate

in real-time updated every second, (2) breath rate displayed as a percentage of their

individual resting breath rate, (3) earned ‘calm points’, an indicator of the number and

duration of previous calm moments (see below), and (4) being below or above one’s

resting breath rate, which colors the text of the entire display blue or red, respectively.

Figure 6.1: The BreathTray shows 4 components: calm points, breath rate, percent relative to resting rate, and whether they are above (red) or below

(blue) their resting breath rate.

Page 119: Augmented Self-Regulation (Moraveji, 2012)

101

BreathTray can be complemented with notifications (see the ‘Breathaware’ chapter)

and prompts but can stand alone as a peripheral awareness cue of respiratory behavior.

Text color provides peripheral awareness without requiring reading.

6.2 STUDY DESIGN

A counter-balanced, controlled, within-subjects study was administered that first

evaluates the effect of the interface on breath rate (calculated as the number of inhalations

in one minute) versus when the interface is not available, controlling for motivation.

Effect of the additional cognitive load of the interface on two different tasks was also

measured.

6.2.1 PARTICIPANTS

There were 14 participants (7 female, average age 34.2, SD=9.23), each given a

$5 gift card to a local eatery as remuneration. Participants ranged from working

professionals to university students; all had a working understanding of Mac OSX and

had not taken part in studies in this domain.

Two different tasks were used to evaluate whether or not the task type affected

whether or not the feedback helps regulate breathing. Both tasks are designed to

introduce significant cognitive load but in different ways, as described below.

The two tasks used are straightforward; most participants had no clarifying questions

after a practice round (see below). The durations of each task, based on study iterations,

were adjusted to be as long as possible without significantly frustrating participants while

fitting the 30-minute study duration.

Page 120: Augmented Self-Regulation (Moraveji, 2012)

102

6.2.2 SERIAL SEVENS

The first task was the Serial Sevens test adapted to a web-based environment. The

Serial Sevens test (reviewed by Taylor, 1988) is a well-studied tool to assess cognitive

impairment because it introduces significant cognitive load (Sweller, 1988). Participants

are presented with a number from which they are instructed to serially subtract in sevens.

The participant must retain that last number in working memory, perform a mental

calculation, derive the new number, write that new number down, go to the next problem

and ensure that they hold that resulting number in working memory to start the process

again as quickly and accurately as possible.

Figure 6.2: The ‘Serial Sevens’ task adapted to a web-based interface. A starting number was shown (top) and numbers disappeared when participants typed

and pressed Enter (bottom).

In the present study, the test was 5 minutes. Text instructions were standardized,

asking that the participant aim for speed and accuracy. The user was shown an empty text

box below the (randomized three-digit number between 900 and 1000) ‘starting number’

(see Figure 6.2). They typed the number which was 7 below that starting number and

pressed Enter, upon which both the starting number and the number they typed would

disappear. After inputting that starting number minus 7, no other text was shown and the

user had to remember the previous number in order to subtract 7 from it and type in the

new number.

Page 121: Augmented Self-Regulation (Moraveji, 2012)

103

The task was scored as total number of correct subtractions. The number of incorrect

responses was also recorded, to ensure that there was no condition-specific ‘speed-

accuracy trade-off’. In the case of an error, subsequent responses were scored as positive

if they were correct in relation to the last number entered.

6.2.3 PROBLEM-SOLVING WITH AUDITORY DISTRACTORS

The second task simulated a form of multi-tasking with high cognitive load:

problem-solving with auditory distractors. For 6 minutes, participants had to choose the

true mathematical expression of two side-by-side expressions (order randomized) (see

Figure 6.3). Each expression had a combination of multiplication and addition in it, so

order of operations had to be adhered to. Concurrently, users listened to pre-recorded

conversations (order of segments randomized). The conversations were between a man

and woman, who are discussing various topics such as where to move. The conversations

varied in affective tone and valence and were recorded specifically to be distracting:

numerous questions are asked between each conversant and distracting comments were

made. Most participants remarked that they were indeed distracted.

Figure 6.3: The math problem in the ‘Problem-Solving with Audio Distractors’ task. Users were to choose the correct expression using the radio buttons and

press the “answer” button.

Page 122: Augmented Self-Regulation (Moraveji, 2012)

104

This task is fundamentally different from the Serial Sevens task in that it contains

a primary task (a math problem that does not require one to use significant working

memory) and a secondary ‘task’ (hearing an interesting conversation). The mathematical

calculations required were not as intensive as working memory was not used as heavily;

users could rely on reading the numbers visually and nothing needed to be remembered

from one problem to the next.

Figure 6.4: The study setup with the USB-connected wearable respiration sensor (left), computer, headphones, and external mouse. The web browser was maximized

to fill the screen for all conditions.

6.2.4 PROCEDURE

The participant sat down at a desk jkwith a computer in front of them (see Figure

6.4). The study administrator sat in an available chair in the room and read a book while

the participant completed the study. Participants first completed a consent form, and then

filled out a short web-based pre-survey with demographics questions and no information

Page 123: Augmented Self-Regulation (Moraveji, 2012)

105

about breathing or self-regulation. The web browser was maximized such that only the

browser and the menu bar were visible. The resolution was 1440x900, making the

dimensions of the BreathTray 1.0”x0.16”. Participants then wore a USB-connected

thoracic strain gauge respiration sensor whose function was earlier described (see Figure

6.4). They watched a 2-minute web-embedded video of the study administrator

describing the benefits of respiratory self-regulation during mild and acute stressors. The

video motivates viewers to maintain calm respiration during work, play, and life in

general. During this 2-min video, the BreathTray feedback was hidden but the user’s

resting breath rate was recorded in the background. Participants wore headphones during

the task. The video was not shown at full screen and participant posture was not

controlled throughout the study, though they remained seated.

Each user did a 90-second practice session of each task in the same order: (1)

Serial Sevens and (2) Problem-Solving with Auditory Distractors. There was no

BreathTray feedback during the practice round. Following these practice sessions, the

administrator asked if there were any questions about each task – rarely were there any.

Participants performed each series of tasks twice, once in each condition. The two

conditions were BreathTray and NoBreathTray. The order of the conditions was

randomized, removing any practice or training effects and minimizing novelty effects.

Novelty effects were also minimized by having participants do the repetitive tasks for a

significant period of time (22min total). Before the BreathTray condition, the participants

were asked to read text that explained what the BreathTray was showing. They were

allowed to ask any questions about BreathTray but were never told to breathe a certain

way. After doing both conditions, participants did a short post-survey.

Page 124: Augmented Self-Regulation (Moraveji, 2012)

106

Our hypothesis was that the feedback would help augment the user’s self-

regulation in multiple ways: (1) the red/blue color in their periphery while they were

focusing on a task, (2) the visual reminder of their own breath rate, (3) the subtle desire to

accrue calm points. However, our study design does not distinguish between which of

these would be the primary cause of any obtained difference in breath rate or cognitive

performance.

6.3 RESULTS

Table 0.1 shows the mean (and SD) breath rates for each of the conditions and

tasks as well as aggregated across both conditions. Figure 6.5 illustrates these results.

Task BreathTray No BreathTray Both

Both 19.7 (1.9) 20.7 (2.2) 19.9 (2.2)

Serial Sevens 19.5 (2.6) 21.3 (2.5) 20.4 (2.6)

Problem-Solving with Audio Distractors

19.4 (1.6) 19.9 (2.3) 19.9 (2.1)

Table 6.1: Mean (and standard deviation) breath rates across both BreathTray and NoBreathTray conditions in each task and across both tasks together.

Page 125: Augmented Self-Regulation (Moraveji, 2012)

107

Figure 6.5: Mean breath rates in each task, split by condition and also aggregated across all condition. The difference between BreathTray and No

BreathTray is significantly different in the Serial Sevens task.

6.3.1 BREATHTRAY IMPACT ON BREATH REGULATION

The first research question our analysis serves to answer is “Did the BreathTray

feedback influence breath rate significantly more than NoBreathTray condition (i.e.,

motivation alone)?”

There was a condition effect on breath rate. Paired t-tests with Welch corrections

were conducted because the variances between tasks and conditions were not equal.

Breath rates in the BreathTray condition (M=19.7, SD=1.9, Med=19.7, where ‘Med’ is

the mean median breath rate across participants) were significantly less than that of the

NoBreathTray condition (M=20.7, SD=2.2, Med=20.9) when comparing both medians

(t=3.17, df=13, p=0.007) and means (t=2.12, df=13, p=0.028). The stronger result in the

median comparison is especially encouraging given how median somewhat captures the

variance of individual breath rate. The median difference in breath rate in the BreathTray

Page 126: Augmented Self-Regulation (Moraveji, 2012)

108

condition was 1.2bpm. A difference of 1bpm corresponds with significant differences in

HRV amplitude (Song, Lehrer, 2003).

The number of calm points received by participants in the BreathTray condition

(M=2.7, SD=3.8), a measure of episodic breath regulation, was larger than that of the

NoBreathTray condition (M=1.6, SD=2.9) but the effect was not statistically significant

(t=1.61, df=13, p=0.13). The majority of users received 0 calm points, indicating that,

more than anything else, the current requirement for calm points is currently too

stringent.

There was a condition effect on the breath rate of the Serial Seven task but not

Problem-Solving. There was a significant difference between conditions in breath rate

between the Serial Seven task (t=3.17, df=13, p=0.007) but not in the Problem-Solving

task (t=0.07, df=13, p=0.944). Thus, it was the Serial Sevens task that accounted for the

majority of the effect between conditions. Participants breathed markedly differently

between the BreathTray (M=19.5, SD=2.6, Med=19.7) and NoBreathTray (M=21.3,

SD=2.5, Med=21.2) conditions during Serial Sevens.

6.3.2 MAGNIFY OR PERSIST?

An important question in the study of peripheral paced respiration (Chapter 4)

was whether participant respiration was decreasing overall or only when the PPR bar was

displayed. It was shown that it was the latter. While one cannot ask exactly the same

question here because BreathTray does not explicitly prompt users to change their

breathing pattern, one can ask, “How were respiration patterns influenced when

BreathTray was available?” It is impossible to know exactly when and how participants

utilized the feedback given that it was peripheral. We visually surveyed the individual

Page 127: Augmented Self-Regulation (Moraveji, 2012)

109

differences in the Serial Sevens task when BreathTray feedback was available or not. One

trend was noted. We quantitatively inspected the standard deviation of the two conditions

of Serial Sevens with the expectation that if the standard deviation of the BreathTray task

was significantly lower, then the feedback caused breath to be lower more consistently.

The two were not significantly different, t=0.059, df=13, p=0.954. This means that there

is no clear sign indicating exactly how the BreathTray feedback influenced breathing: a

persistently lower breath rate or occasional glances up that brought the user’s attention to

their breath to lower for a short period of time.

6.3.3 IMPACT ON COGNITIVE PERFORMANCE

The second research question was “Did having the continuous feedback negatively

impact the performance on the tasks?” This is a non-obvious question because both tasks

required a good deal of cognitive load. Participants frequently remarked about the

difficulty of each of the tasks and that they were unable to spend the desired amount of

time looking at the BreathTray or focusing on their breath. The difference in each task

between conditions was non-significant in number correct, error rate, and, crucially,

response time. There was no condition effect on task performance. I.e., users in the Serial

Seven condition were able to regulate their breath rate without adverse impact on task

performance or response time.

6.3.4 QUALITATIVE FEEDBACK

Post-survey Likert scales of 7 (1=not at all, 7=very much so) revealed that

participants, on average, reported they were not distracted by the feedback (M=2.4,

SD=1.8) and that they felt it influenced their breathing (M=4.1, SD=2.2) but not their

Page 128: Augmented Self-Regulation (Moraveji, 2012)

110

performance (M=3.3, SD=1.6). They also expressed their likeliness to use BreathTray or

similar feedback on their own computers (M=4.6, SD=1.9).

6.4 DISCUSSION

The results of the study show that technology-augmented self-regulation is

effective for a certain class of cognitively demanding tasks represented by Serial Sevens.

That task contains a single intensive task with no distractors. The cognitive load imposed

by the task is significant: users had to focus and were very silent during the task. It

requires users to store elements in working memory, unlike the second task. Unlike the

multi-tasking task, the user was able to attend to the BreathTray feedback while

regulating their respiration. The results suggest that self-regulatory feedback will be

effective during cognitively intensive tasks such as writing and programming. On the

contrary, the results show that one cannot multi-task and expect to be able to use

peripheral biofeedback simultaneously.

The BreathTray design exemplified an approach to designing feedback to

accommodate both elements of self-regulation: monitoring and influence. The changes in

color augmented the user’s ability to monitor changes in their state. The fact that they

knew it was about their breathing gave them actionable feedback to modify that state,

aiming to reinforce their own self-regulative ability.

Levy et al (2011) showed that mindfulness-based practices reduce stress and

improve memory during multi-tasking but that it does not change performance of those

tasks. The present study motivates study of integrating mindfulness-related information

into multi-tasking environments but that it may be the times in between multi-tasking that

Page 129: Augmented Self-Regulation (Moraveji, 2012)

111

the feedback can effectively be used by participants. Perhaps, then, this could reduce

multi-tasking behavior.

The quantitative difference in breath rate between the two conditions during the

Serial Sevens task was 1.8bpm (see Table 6.1), greater than results found in previous

research that used intermittent pacing (i.e., peripheral paced respiration in Chapter 4). It

is noteworthy that BreathTray was able to effect breathing comparable with intermittent

pacing. Future systems would rely on both explicit prompting and peripheral feedback to

augment self-regulation in a form of autonomic interaction design. This concept refers to

the notion that our interactive systems interact not only with our conscious minds and

commands but also with our autonomic nervous systems. Specifically, they can not only

detect and react to changes in our psychophysiological state but can be designed to

influence that state as well.

6.4.1 STUDY LIMITATIONS

Some participants mentioned they felt some ‘math anxiety’ initially; it is assumed

this was reduced in all but the most extreme cases during the practice tasks or early on in

the study. Similarly, the exact seating location of the study administrator may have

caused more or less anxiety in different participants in case they felt they were being

spied on. Again, it is proposed that any such effects waned over the course of the practice

sessions.

Participants may have inferred that the study administrator was interested in calmer

breathing during the BreathTray condition and may have been especially motivated by

this, but (a) they were not told of the BreathTray until it was used (and the order of the

Page 130: Augmented Self-Regulation (Moraveji, 2012)

112

conditions was randomized) and (b) any extraneous motivation was likely reduced due to

the cognitive load of having to do the task itself.

This study does not explicitly identify which of BreathTray’s feedback elements

was responsible for the influence. Our hypothesis is that there is no ‘silver bullet’

feedback that is effective but rather that different people use different influence

mechanisms and that it depends also on task difficulty, eyesight, multi-tasking ability,

and self-regulation performance. Note that specific contributions of different design

elements can be studied in future work.

6.5 CONCLUSION

The BreathTray feedback technique was introduced. A study demonstrated that (1)

peripheral feedback alone can modulate breath rate better than motivation alone without

explicitly prompting or pacing users’ respiration and (2) peripheral physiological

feedback does not draw sufficient attention away from intensive cognitive tasks so as to

negatively impact them. When engaged in a task that simulated multi-tasking, users were

unable to utilize the feedback to regulate their respiration. During an intensive cognitive

task, however, they were. Future operating system enhancements may profitably include

physiological feedback. This study sheds light into how we can expect such feedback to

impact both user physiology and cognitive performance.

Page 131: Augmented Self-Regulation (Moraveji, 2012)

113

CHAPTER 7 BREATHWEAR: AMBULATORY INFLUENCE ON RESPIRATORY PATTERNS

Breathwear is a system for continuous (1) monitoring and (2) influence of one’s

psychophysiological state as reflected by respiratory pattern. In this chapter, design

challenges and opportunities are presented for sociotechnical systems that influence state

for augmented self-regulation (ASR). The iterative, user-centered design of the system is

then described, followed by the qualitative and quantitative results of a longitudinal

exploratory user study to assess user impact and usage. The results suggest further

inquiry into game mechanics and goal-setting as means of conditioning changes in

respiratory pattern and also indicate the robustness of the non-invasive respiration sensor

described. The chapter concludes with a discussion of the derived insights, which are

useful to researchers and designers of systems meant to induce real-time physiological

change using wearable devices.

There has been a surge in the availability of wearable sensors for both research

and consumer use. There are two primary aims in this trend: (a) consumer desire for self-

knowledge to improve health or performance behavior and (b) researcher desire for

Page 132: Augmented Self-Regulation (Moraveji, 2012)

114

ecologically valid and longitudinal data that can reveal insights into human health and

performance phenomena.

These two desires overlap when systems emerge that both help researchers collect

useful data and provide feedback or data that is immediately useful to users themselves.

An example of this is wearable sleep monitors such as the Zeo for sleep researchers but

also for people with sleep disorders (which commonly use accelerometer or EEG

sensors).

In addition to diet, fitness, and sleep, psychophysiological state is a domain of

health that is impacted by lifestyle and behavior. It underlies many, if not all, aspects of

physical and mental health and performance yet few technological approaches have been

made to influence one’s state continuously. Such approaches might attempt to reduce

stress, increase calm, and train users to recognize and adapt to changes in state even when

not using a technological device.

Breath regulation currently requires a great deal of practice, training, and a high

level of compliance that stymies adherence. Our long-term goal is to employ wearable

sensors for continuous monitoring and mobile phones for continuous feedback to

augment respiratory self-regulation and, by association, psychophysiological state. This

chapter first explores related literature, and then describes how user feedback led to two

significant design iterations. The user study presented here is the first known exploratory

study of the ambulatory influence of respiratory patterns using interactive technology.

The chapter concludes with an analysis of the results and discussion of the insights.

Clinical research on the link between psychophysiological state and respiratory

patterns has placed an emphasis on accurate monitoring, detection, and description of the

Page 133: Augmented Self-Regulation (Moraveji, 2012)

115

stress response but not necessarily on influence. For example, Wilhelm et al. (2006) had

users wear electrodes and a computer waist pack (Wilhelm, Roth, Sackner, 2003) and

collected ECG, skin conductance (on fingers), two thoracic strain gauges, finger

temperature gauges, accelerometer, ambient temperature, and barometric pressure in

three situations: quiet sitting, physical exercise, and a short commercial airplane flight.

Their goal was to distinguish between psychophysiological influence on sensors and

physical activity (exercise). They found that, of all the indicators, “certain parameters of

irregularity in breathing were [responsive] to anxiety” and that it is only respiratory

patterns that separate the resting, physical activity, and emotional conditions,

“emphasizing the high informational content of respiratory pattern analysis”.

Pfaltz et al (2009) conducted a robust ambulatory respiration monitoring study –

in this case two 24-hour time periods separated by 1 week, comparing subjects with

‘panic disorder’ (PD) with healthy controls. They replicated effects found in the

laboratory that distinguished changes in respiration during states of anxiety in both

groups. However, the results did not show the two groups were significantly different.

These results can be interpreted to mean that (a) wearable sensors can detect states of

anxiety and that (b) anxiety spans all populations of people and that separating treatment

for only those with PD is not meaningful. The sensor system employed to conduct these

studies is wearable but is not something one can be expected to wear for many

consecutive days.

Our intent, differentiated from related work, is not to improve or validate the

sensing of ambulatory respiratory patterns but to influence it. Clinical-grade accuracy is

not required for our goal, nor are all related respiration parameters. The respiratory

Page 134: Augmented Self-Regulation (Moraveji, 2012)

116

characteristic with the greatest utility for this purpose (breath rate) was identified,

motivating focus on studying the interaction design of a system for feedback and to

influence that characteristic by a gross amount. The next section outlines design

opportunities and challenges for designing an ambulatory version of such a system.

7.1 DESIGN OPPORTUNITIES AND CHALLENGES

Using a mobile device is a qualitatively different user experience than designing

for a laptop (Pea & Maldonado, 2006). The device sits on your person, feels like an

extension of the body/mind, has a dramatically different screen size, has opportunities for

other forms of physiological sensing, and perhaps most important, is with the user more-

or-less continuously. Mobile phones already extend one’s cognition (Pea, 1985) and are

the logical medium for augmenting self-regulation. Before designing, we identified

potential design opportunities and challenges for augmented self-regulation (ASR) using

wearable sensors and feedback on a personal device (e.g., mobile phone). ASR systems

are a class of autonomic interaction design (AIS) systems that use AIS approaches

explicitly to augment self-regulation.

7.1.1 CONTINUOUS STATE INFLUENCE

Imagine the difference between a pedometer that gives you feedback only at the

end of the day compared to one that tells you how many steps you’ve logged thus far in

the day. Now apply that analogy to psychophysiological state and stress. With desktop-

based influence, demonstrated with peripheral paced respiration (Chapter 3), the user

associates monitoring and influence with a stationary laptop that must be turned on.

Using continuous feedback, the user can then begin to notice differences between

Page 135: Augmented Self-Regulation (Moraveji, 2012)

117

information tasks as well as difference between physical tasks or activities such as eating,

reading, listening to a lecture, driving, watching a movie, and so on.

7.1.2 CONTEXT-AWARENESS

Using mobile devices, feedback from the system can take into account different

characteristics of the environment or user’s behavior (Fogg, 2002). For example, if they

are walking, behavioral triggers can be delayed. If the user is co-located or even socially

connected with other users, feedback can take this into account. Time of day can

influence cortisol levels and other stress-related indicators in the body (Sherwood, 2006)

and can also be considered.

7.1.3 SLEEP

One’s state has an impact on sleep quality and patterns (Carskadon & Dement,

1981; Shahar, et al., 2001). The body’s physiological indicators also change according to

stage of sleep, including respiration (Shewood, 2006). Though not an ostensible goal of

our work, there is a great deal of research on the role of respiration in sleep disorders.

Through a mobile device, users can monitor and perhaps even influence sleep patterns.

Partners could watch over one another during sleep or parents could care for children by

monitoring their respiration (co-located or remote).

7.1.4 SOCIAL

Opportunities exist in co-located and distributed scenarios for socially influenced

synchronization or influence on motivation (see Chapter 4). Synchronous physical

behavior could be complemented or replaced with synchronized physiological behavior,

Page 136: Augmented Self-Regulation (Moraveji, 2012)

118

synchronizing states between users. This could be useful for meetings, meditation classes,

yoga courses, or other intimate social gatherings.

7.1.5 ANNOYANCE

A primary challenge of our domain is that it attempts to influence a process that is

continuously occurring. This is dissimilar from exercise, diet, and other related domains

of intervention. There is a high probability that system designs in this area will annoy

users with frequent notifications or making the user feel that a very intimate aspect of

themselves is being monitored.

7.1.6 EVOLVING USER GOALS

Users have different goals and intentions for different times of day, weeks out of

the year, contexts, and tasks (Fogg, 2002). While studying for an exam, the user may

have one goal that the system could support but while taking an exam, they could have a

completely different goal. While meditating, for example, the user may have very

stringent goals for system guidance. However, they may wish monitoring to be

continuous. Either the system must automatically detect these different contexts and

intentions or must be very simple to customize according to not only individual desires

but also contextual desires.

7.1.7 OVER-DEPENDENCE

A potential issue in ASR systems is the level of granularity at which feedback can

be useful. One’s state is constantly being managed by the body and the mind; looking to

an external device for feedback on self-regulation can create an over-dependence on the

Page 137: Augmented Self-Regulation (Moraveji, 2012)

119

system. Systems could customize the amount and granularity of feedback given or the

situations that should produce feedback.

7.1.8 INACCURACY

Monitoring one’s state is different from monitoring steps. In the latter, if there is

some level of inaccuracy, the user can clearly blame the system. Because breathing

patterns are not as easily discernible as steps taken yet are more visible than indicators

like heart rate or skin conductance, feedback of breathing patterns can be suspect. Worse,

false negatives can create stress when it did not exist or had subsided.

7.2 INTERACTION DESIGN GOALS

This section describes the interactive client component of the Breathwear system

(on the mobile phone). At first glance, respiration feedback alone may seem like the

primary function of the client. However, our goal is to influence the user’s

psychophysiological state in a continuous manner, not only when the user wants to reflect

about their state or remembers to check the interface. Feedback per se is not the goal of

the system. As such, three primary design goals (monitor, influence, and customize) are

presented below. They follow the components of self-regulation identified earlier, namely

the self-awareness to monitor one’s own state and the capability to influence it.

7.2.1 MONITOR

At its most basic level, the system must ensure the user that their state is being

monitored while the user is engaging in other tasks. This must be as invisible as possible

so the user can be fully engaged in their work and life tasks. Reducing sensor bulkiness is

a big part of this but the software design must also ensure the user knows the system is

Page 138: Augmented Self-Regulation (Moraveji, 2012)

120

‘on’, is recording breath rate to an appropriate level of accuracy, and is storing the data

for later perusal (without compromising the user’s privacy).

7.2.2 INFLUENCE

What differentiates the system from related systems in the literature is that the

system has a goal beyond assessment: it attempts to influence the user’s behavior. One

method of doing this is by simply knowing the system is monitoring one’s behavior.

Another method is with behavioral triggers (Fogg, 2002). To support this, Breathwear

supports ‘push notifications’ on the iPhone user interface. These are both

positive/congratulatory messages and negative/cautionary messages to support operant

conditioning of breathing behavior (Ley, 1994).

7.2.3 CUSTOMIZE

The user should be able to re-record or adjust their baseline easily because people

naturally breathe at different rates. The baseline becomes not only a baseline of one’s

state at rest but also an intention of their desired state. This can change quickly, as was

found in early user evaluations.

7.3 DESIGN ITERATION 1

An initial prototype on an iPhone 4S was created to receive data continuously and

wirelessly from the breathbelt sensor. The initial prototype attempts to address the three

design goals identified above.

The Dashboard (Figure 7.1, left side) shows the user’s real-time breath rate, raw

sensor value (‘Raw Data’), number of calm points (see Chapter 5.3.5), and number of raw

sensor records counted so far. Sensor values are provided for debugging purposes.

Page 139: Augmented Self-Regulation (Moraveji, 2012)

121

Buttons to connect/disconnect and trigger a buzzer on the belt are also shown. The

progress bar on the bottom is the sensor’s battery level.

Calm points are analogous to steps with a pedometer. They only increase and are

reset to zero at the start of each day. A user gets points by breathing calmly relative to

their own resting rate; the calmer they breathe, the bigger the increment. They are

awarded a point for each 30sec they are at or below their resting rate. Breathing at or

below one’s own resting rate for 30sec adds 1 point. The increment increases linearly as

the user’s breath rate decreases.

Figure 7.1: The initial prototype. Dashboard (left) shows real-time feedback and Settings (right) allows the user to make changes to their breath rate baseline and to manually send data back to the research team via email.

When the user clicks ‘Notifications’ in the bottom, they are shown a simple list of

notifications that have appeared so far. The Settings (Figure 7.1, right) screen allows for

debugging and testing of parameters during this phase of research. The user can change

their breath rate manually and can manually send the data via email to research team

members for analysis.

Page 140: Augmented Self-Regulation (Moraveji, 2012)

122

To influence user respiration, the system employs real-time behavioral triggers in

the form of push notifications for operant conditioning (learning based on modifying the

environment to reinforce particular behaviors): negative feedback (Figure 7.2, left) and

positive encouragement (Figure 7.2, right). The notifications that appear to the user

mirror those used in Breathaware (see Figure 5.5). These are shown on the phone’s lock

screen as well as during phone use (Figure 7.2, center). The number of notifications that

can be ‘stacked’ on the lock screen is first determined by the system settings

Figure 7.2: Push notifications on an iOS5-based mobile phone based on initial prototype design. Each notification has a type (positive or

negative/cautionary) and duration that the system detected the user was in that state. For example, the user had been breathing above their resting rate for

15.2min at left, at or below their resting rate for 12.7min on right.

7.4 Primary insights (PIs) from participant feedback

The first iteration of the prototype was tested with three friends of the lab wearing

it for several hours a day for 2-5 consecutive days. The following primary insights (PIs)

were found from user feedback and interviews:

1. Users felt a part of them was being monitored that they were not used to: “How

I’m being”.

Page 141: Augmented Self-Regulation (Moraveji, 2012)

123

2. Users desired feedback about not only the present moment but also their recent

history. 10 minutes seemed to be a suitable amount of time.

3. Users desired a qualitative measure of performance, not only quantitative

indicators.

4. Users wanted to change the baseline according to their context (or that the system

would automatically do it). At times, they wanted support for being very calm but

at other times they wanted fewer notifications.

5. Notifications became ‘stale’ quickly. If they were not seen immediately, they were

not useful later.

6. Battery consumption on the phone was noticeably increased.

7. Users had no precedent for what were ‘low’ and ‘high’ breath rates so the

feedback was difficult to interpret.

8. When notifications ‘piled up’, they felt like spam and were stacked on the phone’s

lock screen.

Design of the client interface was iterated upon to address a significant number of these

PIs. These are presented in the next section.

7.5 DESIGN ITERATION 2

The primary goals of the second design iteration were first to support quick-

glance, qualitative comprehension of one’s state and, second, to support customization

that would improve subsequent studies. Based on initial user testing, the interface was

changed (see Figure 7.3). The changes are summarized in the sub-sections below and

include reference to the primary insights (PIs) from the previous section they are meant to

address. They still aim to address the three design goals described in Chapter 7.2.

Page 142: Augmented Self-Regulation (Moraveji, 2012)

124

7.5.1 RECENT ACTIVITY INDICATOR

To address PI2, a visual indicator of recent activity and performance was added

based on 10 minutes of prior data. This number was selected from user interviews with

prior users of the system and domain experts. The bar can be three different colors: red

(mean of last 10 minutes is more than 1.5bpm above resting rate), yellow (within a range

of 1.5bpm above or below resting rate), or green (below 1.5bpm of resting rate) (to

address PI3). The user can make a quick glance at the interface to infer how the recent

past has been for them. There were a number of possible designs, including adding a

slider to adjust the time window in real-time (to see, for example, mean performance over

the past hour or day). However, we elected to maintain a low level of complexity given

that the design goal was to create an interface that could be consumed with a glance.

Figure 7.3: The second iteration of the Breathwear client interface, which includes a recent activity indicator (left, top, in green) and additional configuration options in the Settings screen (not shown). The center image

shows when the user is hovering around their baseline (hence the yellow) and the sensor is connected (hence the ‘Time Connected’ indicator). During a high breath rate state, the band is red (not shown). The baseline here is set to

15bpm (the default).

Page 143: Augmented Self-Regulation (Moraveji, 2012)

125

7.5.2 RELATIVE BREATH RATE

To address PI7, the relative breath rate indicator was changed to show not just the

total percent comparison but to use the words “below” and “above” to indicate exactly

how much above or below the baseline the user is breathing right now (see Figure 7.3,

left and center).

7.5.3 BREATHBELT FEEDBACK

Additional feedback was added to the breathbelt to address user complaints. First,

the belt will beep when one presses one of the two buttons (in order to confirm it has

battery power and is reacting to user input). Second, the belt beeps multiple times when it

loses connection with the phone (either via distance, Bluetooth software problem, or

other). This can help avoid long periods of time without data collection.

7.5.4 ADDITIONAL SETTINGS

Three additional settings were added, not shown in the screenshots above. First,

the option was added to disable push notifications completely (to partially address PI5). It

is unknown if users would want to use the application this way and, if so, how it would

be used. For example, perhaps users would turn off notifications but glance at the

application more often but only during periods in which they desired to be calm (e.g.,

taking a test). Second, the ability to hide breath rate feedback on the dashboard was

added. The reason for this was research: the desire to use Breathwear to measure baseline

breath rates for users as a control and then use the system with feedback (and

notifications) to see if it influenced their breath rate. Third, another parameter added was

the ability to modify the breathbelt sample rate (to address PI6). This was added to

address prior concerns that the application drains the iPhone battery too quickly. The

Page 144: Augmented Self-Regulation (Moraveji, 2012)

126

team was not certain how frequent one could make the sample rate before the breath rate

accuracy was compromised so multiple levels exist that the user can select from.

The application was instrumented to log anonymous but granular information

including application usage, changing settings, respiration rate, baseline, sensor data, date

and time, notification events, and application usage summaries.

7.6 STUDY

An exploratory in situ, longitudinal user study was conducted with 6 users (2

female, 4 male, median age 31.5) who used Breathwear over the course of a 5-day

workweek. The users varied in profession and background but all were technologically

savvy and were interested in maintaining a calm state. Participants were recruited through

university and common interest email lists. Participants were warned the prototype may

contain bugs and that the goal was to get feedback and improve the design of the sensor

and feedback. One user’s data was not collected because their sensor repeatedly broke.

This was the only user who broke a sensor.

Participants met with the study administrator individually and were given

instructions for how to manually install the application on their phones and were shown

how to wear the adjustable sensor. They were asked to use the system as much as was

comfortable for several days and the study administrator would send them intermittent

SMS reminders and updates over the course of the study. The participants started on

different days and each had the same model of mobile phone (Apple iPhone 4S).

Participants were given a short URL to access system troubleshooting tips and

feedback form (see Appendix D). They were told they should visit this page and add any

comments, feedback, questions, or criticisms at any time. In this way, the research team

Page 145: Augmented Self-Regulation (Moraveji, 2012)

127

hoped to collect user feedback close to the time it occurred. Each user was occasionally

sent an SMS reminder to disconnect the sensor before going to sleep each night (to

conserve battery). Before doing so, they were also to send the log file to the study

administrator and clear yesterday’s data using the Settings screen (see the right side of

Figure 7.3). They were also asked to remove the sensor during exercise.

7.6.1 RESULTS

Participants used the system for an average of 37.26 hours (median=20.5,

SD=30.2) over a 5-day period for a total of 186.3 hours. Of that possible duration, 140.5

hours (75.4%) of the breath rate data was deemed usable and the remaining was deemed

unusable due to sensor battery problems, software malfunctions, or a mis-worn

breathbelt. This data loss was expected given the exploratory nature of the study and the

fact that participants were wearing the sensor in their daily lives. It can be improved by

improved robustness in belt and sensor design. Users opened the application an average

of 3.0 times per hour (median=2.9, SD=1.7). Generally speaking, users did not manually

adjust their baseline breathing rates, keeping them at the default 15bpm.

The mean breath rate across all participants over the entire duration was 13.64

(median=13.77, SD=1.97). However, individual breath rates were distributed differently

over the course of the study, as shown in Figure 7.4 below. Drawing upon the literature

on stress and the stress response, one can always expect to encounter relatively high

breath rates even over a relatively calm week. These can occur for myriad reasons that, in

Western culture, are mostly subjective or psychological (rather than resulting from a

physical threat to one’s safety).

Page 146: Augmented Self-Regulation (Moraveji, 2012)

128

Figure 7.4: Distribution of breath rates of each participant. X-axis is “Breath rate in Breaths per Min”. Y-axis is ‘Frequency in Seconds’.

Clockwise from top-left, users 1, 2, 4, 5, and 3.

Page 147: Augmented Self-Regulation (Moraveji, 2012)

129

The reader will note here that, in all four cases (including the omitted), breath rate

distribution is normally distributed. This is as one would expect, where one’s breath rate

has regularity to it although there are different patterns of breathing discernable. Because

ambulatory respiration sensing is so rare in the literature, figures such as these are rarely

(if ever) found and interesting to analyze. For example, User #3 has almost two ‘styles’

of respiration: high (10-15bpm) and low (4-9bpm) breath rate. This user later was found

to have used the system during meditation practices. Users #1 and #4 have a very normal

distribution of breath rates while User #5 had little slow respiration duration. An

intelligent mobile system could identify such longitudinal patterns and recommend

changes in lifestyle or behavior. User #2, a frequent meditator, has a distribution that

illustrates the amount of breath awareness he maintained, with a spike in frequency

appearing around 5bpm. In post-hoc user interviews, the user meditated on average about

1 hour per day. This consistent practice is clearly evidenced in the graph.

Users received an average of 45.2 (median=41, SD=43.1) push notifications (1.9

per hour). Of those, 32.4 were positive (‘Calm’ and ‘Zen’), 12.8 were negative (‘Stress’)

feedback. ‘Zen’ notifications were given during especially long calm episodes and were

relatively rare and difficult to achieve. Interestingly, one user received almost no

notifications over 50 hours of usage. Figure 7 shows the frequency and types of

notifications received across the five study participants. No clear trend is evident,

speaking to the system adapting to each user’s individual behavior.

Page 148: Augmented Self-Regulation (Moraveji, 2012)

130

Figure 7.5: Frequency and type of push notifications received by participant. ‘Calm’ and ‘Zen’ are two types of positive notifications. This graph shows that there was no discernable trend around type or frequency of push notifications but that one can characterize an individual’s respiratory patterns to some

degree using this visualization.

User #2, the frequent meditator, is shown to exhibit very little fast breathing

episodes while Users #3 and #4 both experienced more stress states than calm. User #5

received no push notifications during the 18 hours he used the system. This could be

because he hovered exactly at his resting rate for the duration of the study. In post-hoc

interviews, it was discovered he was a computer programmer and spent the vast majority

of his day working in a consistent environment and pace.

Users were awarded an average of 105.9 calm points per hour (median=78.6,

SD=68.3). This number, which seems high, is attributed to the fact that one user received

disproportionately more calm points and to the fact that users could receive greater

numbers of calm points the slower they breathed. User #2 (the frequent meditator) was an

outlier, receiving 226.9 calm points per hour (see Figure 7.6). The median number of

calm points granted hourly, 78.6, is more useful here, and is still higher than expected.

Page 149: Augmented Self-Regulation (Moraveji, 2012)

131

Given the baseline was usually 15bpm and that the average breath rate across participants

was 13.6bpm, one can see how users were granted calm points so frequently. Calm points

were designed to be relatively easy to get (i.e., as easy as steps are using a pedometer).

The user interface did nothing with the calm points and users were not explicitly told how

many points they received yesterday or how many their peers received. Later, it was

found that participants noticed them and wished to use them more explicitly.

Figure 7.6: Calm points per hour granted to each study participant. This graph shows how User #2, who had a great deal of meditation experience, was rewarded

a great deal more calm points per hour than the other participants.

Individual breath rates varied over the course of the study as illustrated by one

user who took the liberty to send herself a day’s worth of data, label it, and produce a

graph to gain insight into how her breath rate changed across tasks (see Figure 7.7). This

was an exciting and unanticipated instance of a user being personally interested in her

Page 150: Augmented Self-Regulation (Moraveji, 2012)

132

data. She later described this insight clearly to us in user interview although it had

happened days earlier. It had clearly made an impression on her.

Figure 7.7: A line graph produced by one participant with data she labeled herself, “meditating”, “surfing the web”, and “reading”.

Looking at the user-generated figure, the reader will note three distinct breathing

patterns (although we cannot objectively confirm their validity): meditation, information

work, and reading. The breath rate during meditation hovered around 4bpm while reading

was slightly more erratic however still low at 6bpm. Surfing the web, albeit for only

10min, resulted in an increasingly fast breath rate, replicating findings from our earlier

related studies. It is not clear how fast the breath rate would have been had the user

continued using the computer for an hour.

The experience of seeing a visualization of their own data made the user more

aware of their daily experiences. She later told us “It was surprising to see how my breath

rate would go up and down over the course of an hour. [The system] gave me an

understanding of my breath rate during different activities. So, like, in meditation my

breath rate could get very low versus walking. But there’s even a difference between

reading and watching a show on television. I’m 10-13 when I’m chilling out but 15 when

I’m working on the computer.”

Page 151: Augmented Self-Regulation (Moraveji, 2012)

133

The occurrence and effect of push notifications is difficult but interesting to

measure. The notification itself is meant to give the user a sense of their recent

performance but also to influence their subsequent respiratory patterns. depicts a

representative session of use (a session being the time the sensor was connected to being

disconnected) with the different types of push notifications represented (User #3).

First, one notes that the placements of the notifications are in logical locations.

For example, when the user’s breath rate is elevated above 15bpm, a stress notification

appears in a group (approximately 250-270min) until the user notices and takes action

(from 280-350min). Though the data is not labeled by activity, the reader will see that the

feedback does seem to maintain a breath rate at or under the baseline (15bpm).

Page 152: Augmented Self-Regulation (Moraveji, 2012)

134

Figure 7.8: A line graph showing the relationship between a user’s breath rate punctuated by the different types of push notifications (green=calm, blue=zen, red=stress). The line indicating the user’s breath rate is by default gray and then colored according to the duration of the state detected by the subsequent

push notification.

Page 153: Augmented Self-Regulation (Moraveji, 2012)

135

7.6.2 USER FEEDBACK

Given the relatively small sample size and exploratory nature of this study, user

feedback is highly valued and can lead to important design considerations. The study

participants were enthusiastic about the possibilities of such a system but had a number of

concerns that must be addressed in future iterations.

The notion that the system is monitoring something that was previously

unavailable was not lost to the participants. This speaks to a simple measurement effect

that, in and of itself, is a useful feature of the system. One participant noted that “having

my phone track this data is making me more self-aware about it.”

The ‘baseline’ concept became used more like a ‘target’ because users could

manually change it and would change it arbitrarily in order to instruct the system to give

different types and frequency of feedback.

Breathwear monitors something that is changing continuously: one’s

psychophysiological state. As such, it is something that isn’t only useful upon later

reflection (such as how pedometers and other fitness devices are used) but in real-time.

Users voiced the desire for an even easier and quicker access to their state via the

iPhone’s lock screen (which cannot be modified programmatically without ‘jail-breaking’

the phone). As one user put it, “I got into the habit of looking at the screen and I wish it

were easier to see that.”

Calm points as an experiment were useful to a degree but were clearly not used to

their potential. One user said they “could be interesting but I’m not sure about milestones

or what I’m shooting for. Maybe I can be told the # of points that others reached? It

Page 154: Augmented Self-Regulation (Moraveji, 2012)

136

didn't give me something to aim for. Like ‘today 56min versus yesterday 32min’. I

wanted something to aim for day to day.”

After one user’s battery died, they mentioned they "miss wearing it and getting

information about breath rate." After using the system for 5 days, that user found there

were “definitely times when I'm walking and will think about my breath, not something

I'm used to doing.”

One concern our team had was whether sensor accuracy would hold up outside

the lab. Indeed, one user asked, “if it could detect if I was moving it could have a separate

moving state.” She found that the sensor readings looked jumpy while she walked.

User #1 took intricate notes about how her behaviors were affecting her breathing

patterns. She then sent SMS messages directly to the study administrator as they

happened. Examples included screenshots of the application screen and text

accompanying it describing that it was taken “Fri night after a dinner and then two

colas”. Later: “Never realized how much caffeine affects me!” and “rate still high two

hours after caffeine. Even run-walking w my dog, Savannah, I only reached yellow--still

a lesser effect than caffeine!” Clearly, one cannot draw a causal connection between

caffeine and breath rate but providing users with the data to self-experiment can be a

robust means of identifying causal relationships for the purpose of changing one’s

behavior.

User #1 also mentioned that the system told her she was in a ‘green’ state when in

fact she was so stressed that she was engaging in breath-holding. This is clearly a new

feature of the system that future research must address.

Page 155: Augmented Self-Regulation (Moraveji, 2012)

137

7.7 DISCUSSION

This study explored the impact and usage of a system for continuous monitoring and

influence of psychophysiological state based on sensing, feedback, and incentivizing

changes in respiratory patterns. To interpret these results, we refer back to our three

design goals: monitor, influence, and customization.

To support the goal of effective monitoring, the system performed well for a research

prototype. 25% of the data was lost due to disconnections, sensor drops, and other

unexplained incidents. This can clearly be improved. Another place for improvement is

the potential for using additional sensors to detect physical activity such as walking. The

placement of the sensor, over or under one’s shirt, resting on the hips, and the tightness of

the belt itself, seemed to work well. The system was able to differentiate between

different activities and illustrated differences between people who regularly practice

meditation.

To effectively influence the user, the chapter discussed how notifications were used

and iterated towards user interface design techniques that allow quick glances at both

quantitative and qualitative indicators of respiratory performance. Calm points were

indicative of a user’s performance but are not clearly designed as motivators. There is a

great deal of work to be done around arranging goals, milestones, and other methods of

influence.

The different ways in which calm points and push notifications were delivered to the

user shows one way the system tailors feedback to users. We did not see users explore

changing their baseline breath rate. This would be a way of influencing the frequency and

type of notifications (and calm points) but we did not discuss it with users in great depth.

Page 156: Augmented Self-Regulation (Moraveji, 2012)

138

This presents a large area for future possible research: automatically adjusting the

baseline breath rate so that system responses influence and reward the user appropriately.

Participants were intrinsically interested in their own data, asking to wear the sensor

longer, for visualization of their data, and for more features.

The push notification logic in this prototype was rather simplistic and though only

occasionally useful, one can imagine the potential utility of taking the user’s context into

consideration in the notification logic. Such context-sensitive influence could be

improved over time such that the system does not attempt to influence the user during

exercise, in a meeting, or similar.

This chapter began by describing the intention of some system designers to create

systems that are useful to both researchers and users themselves. The early results of this

system deployment lend insight into the design of sociotechnical systems that help users

in a real-time manner while providing researchers with data previously unavailable

without highly invasive sensors. The team plans to iterate and use this system to amass a

large, labeled dataset of respiration data of users in their real lives. This is of use to the

research community to both educate and influence the public to take steps to regulate

their state to ensure a healthy and a productive population.

Another next step is to explore design methods of enabling users to label their own

data in real-time, as this has proved to deliver insights to users about the effect that their

behaviors are having on their state. The system could detect changes in breath rate and

prompt the user to label their current activity. These results could then be used on graphs

to provide contextual data of their state and perhaps to identify patterns that were buried

before.

Page 157: Augmented Self-Regulation (Moraveji, 2012)

139

Lastly, the system must differentiate between slow breathing and breath-holding.

This is non-trivial because while User #1 engaged in breath-holding, it could look similar

to being in a meditative state with a slow breath rate. Differentiating between them may

require looking at tidal volume, respiration regularity, and inhalation depth, all things the

research team is now attempting to infer from the sensor data.

7.8 CONCLUSION

Two design iterations of an ambulatory respiration monitoring and influence system

were presented. The results of an exploratory, longitudinal user study showed the system

effectively assessed different individual respiratory patterns and responded to changes in

state. User feedback shed light into how the system could better motivate and make

tangible the process of breath regulation using goals, visualizations of individual episodes

and data, and improved methods of quickly assessing one’s state in the present moment

and recent past.

Page 158: Augmented Self-Regulation (Moraveji, 2012)

140

CHAPTER 8 CONCLUSIONS AND FUTURE WORK

This dissertation motivated then investigated technology-mediated

psychophysiological self-regulation via stationary and ambulatory respiration monitoring

and influence as a means of augmenting human self-regulation.

8.1 SUMMARY OF FINDINGS

We refer to our initial research questions to summarize the findings.

• Q1: Is it feasible to augment respiratory patterns of information workers as they

are engaged in meaningful information work?

Chapter 3 demonstrated that visual pacing feedback can be designed to complement, not

interrupt, the user’s task across operating system windows and applications (PPR). The

pacing method reduced mean breath rate by almost 2bpm but failed to develop user

consistency in maintaining that low breath rate. This motivated further inquiry into social

feedback to motivate that consistency.

Page 159: Augmented Self-Regulation (Moraveji, 2012)

141

• Q2: How does synchronous social feedback compare with asynchronous feedback

in peripheral paced respiration?

Chapter 4 investigated two types of social feedback: asynchronous and synchronous.

While it did not develop the respiratory consistency expected, it did uncover an

interesting effect of synchronous feedback magnifying the effect of the social feedback.

We also replicated findings from the prior study showing that information work in

general results in significantly breathing rates.

• Q3: How can sociotechnical systems be designed so as to motivate respiratory

change without prompting it explicitly?

Chapter 5 identified 10 design principles for designing desktop-based systems that

influence respiratory patterns and habits without requiring user interruption. The systems

utilize motivational cues and are clear in their intent to develop (not replace) the user’s

innate self-regulation skill. The Breathaware system was designed and implemented to

demonstrate the designs.

• Q4: How do peripheral feedback and motivation cues influence respiration and

does that feedback negatively influence cognitive performance?

Chapter 6 also included a controlled evaluation of a primary component the Breathaware

system: the BreathTray. It aided respiratory self-regulation without compromising

cognitive performance in two types of tasks: problem-solving while multi-tasking and a

more intensive single cognitive task. Further, users were able to effectively utilize the

peripheral feedback, which lacked any prompting or pacing during the single task

condition only (i.e., not while multi-tasking).

Page 160: Augmented Self-Regulation (Moraveji, 2012)

142

• Q5: How must these techniques be adapted to be effective in a mobile setting?

This final study presents the design opportunities and challenges for mobile, continuous

monitoring and ambulatory influence of respiration. Two design iterations, user feedback,

and results from a longitudinal user study were presented showing that the system

adapted feedback according to individual respiration patterns. Recent activity indicators

were used to give users both quantitative and qualitative understanding of their state in

the present moment and the recent past. Users were intrinsically interested in their own

data and the sensing method was found to be relatively robust for daily use.

8.2 LIMITATIONS

This section contains overall limitations of the approach taken here and, as a

result, of conclusions that can be drawn.

8.2.1 BREATH RATE ALONE IS NOT ILLUSTRATIVE OF AUTONOMIC ACTIVITY

We identified breath rate as the optimal characteristic to begin with but some

studies focus also on tidal volume as a key to assessing user state. The reason is to

understand whether users are breathing slower but more shallow, which is associated

with anxiety (Roth, 2005). At present, the breathbelt does not detect tidal volume but we

are working towards that goal. The Breathwear client can also work with other respiration

sensor systems that do detect tidal volume. Our focus has not been on sensor design but

on behavioral influence.

Further, breath rate has been studied less than heart rate and heart rate variability,

which has been shown to be a highly robust and rich measure of arousal and

psychophysiological state. Future studies should collect both measures to ensure the

Page 161: Augmented Self-Regulation (Moraveji, 2012)

143

findings are validated. Further, a bi-directional link between breath rate and heart rate has

not been firmly established and is likely mediated by other factors in the autonomic

nervous system. In short, replicating the effects with heart rate as an additional measure

will provide needed validation of the study claims.

8.2.2 THE EVALUATOR EFFECT

In the laboratory studies, the presence of the evaluator can influence user

behavior, perhaps motivating them to ‘do what the administrator wants’ so that they get

desired results on their study. We attempted to mitigate these as much as possible. Even

in ambulatory studies, there can exist an evaluator effect that may wane when the study is

over.

Few of the study participants had had experience with biofeedback or

physiological computing in general. As a result, users could have also been subject to

novelty effects that may have stimulated arousal and sympathetic nervous systems,

resulting in even higher breath rates (but perhaps more attention paid to respiration

feedback).

8.2.3 COMPETITION CONFOUND

In controlled studies there is often a competitive undertone given that participants

know that the study will involve other participants as well. In at least one study

(BreathTray), we did ask users to do the best they could – and this may have been

interpreted to mean breath regulation as well. As a result, users could have felt that they

were ‘competing to be calm’, influencing results at least in the initial part of the studies.

Page 162: Augmented Self-Regulation (Moraveji, 2012)

144

8.2.4 SINGLE DATA SOURCE

I have motivated the use of breath regulation as a means of influencing

psychophysiological self-regulation using prior controlled and ambulatory studies but we

have not validated it ourselves. We run the risk of relying upon prior studies without

replicating those results here. For example, adding multiple sources of physiological

input to our client could discriminate between high arousal/low breath rate and true

resting breath rates.

8.3 SUPPLEMENTARY CONTRIBUTIONS AND IMPLICATIONS

This section offers supplementary contributions encountered while addressing the

research questions above. Following these is a discussion of the implications of these

supplementary contributions for the field and humanity writ large.

8.3.1 AUTONOMIC INTERACTION DESIGN

The studies presented in this dissertation are essentially a means of initiating and

motivating a field of autonomic interaction design (AID) through interactive technology

that aims explicitly to influence psychophysiological state during everyday usage (rather

than limiting such impact to biofeedback ‘training’ sessions and the like). AID is a field

relevant to the study of autonomic interaction systems (AISs) outlined earlier. Each study

attempts to operationalize the notion of augmenting human self-regulation by creating or

furthering real, working systems and evaluating real, observed changes on human

physiology.

A working definition of autonomic interaction design (AID) is designing to

purposefully influence or interact with the human autonomic nervous system. Principles

Page 163: Augmented Self-Regulation (Moraveji, 2012)

145

of AID are also useful in affective computing (Picard, 2003), which aims to interact and

adapt to human emotions. AID is concerned less with human emotions and more with

autonomic self-regulation itself, which influences and regulates emotions experienced by

the mind. The utility in making this distinction between affective computing and AID is

that the latter can serve to help augment the human ability to experience, amplify, and

regulate emotions. This helps develops agency in how one chooses to experience

emotions and is a key to wellbeing. Augmented self-regulation systems (ASRs) are AISs

that explicitly aim to augment the user’s innate ability to self-regulate (rather than simply

to interact with the autonomic system in general).

The AID approach maintains that technology must be consciously and

compassionately designed to augment our innate self-serving abilities, not only to replace

or manipulate them. When we approach technology design with this perspective,

technology becomes again a meaningful tool to support human beings to live meaningful

lives, rather than creating new technologies for its own sake or to solve inadequacies

found in prior technology.

8.3.2 ASR AND INCENTIVIZING SELF-REGULATION

The irony of the idea of ASR systems, of course, is that users can develop over-

dependence on external tools to improve their own self-regulatory ability (see Section

7.1.7). Though this is a common concern in HCI systems, it is especially interesting in

this case because the systems is aiming to develop the user’s self-regulatory ability,

which itself would monitor and protect one from over-dependence.

What differentiates ASR systems from others is the notion that such systems must

not incentivize their own use but, rather, effective self-regulation. Though this seems a

Page 164: Augmented Self-Regulation (Moraveji, 2012)

146

logical conclusion, we consider it novel because inherent to so many interactive systems

is to design continued engagement into the system.

The implicitly valued HCI principle that effective tools are ones that are used

frequently, or even well liked by users, is a crucial issue that fails to honor the humanity

of the user. By making themselves indispensable, technology tools can become

‘permanent crutches’ that produce in the user a feeling of need or dependence, similar to

one addicted to a pain-killing substance that was once a tool but has now become the

source of the problem and must itself be escaped. Such technologies unwittingly serve

themselves (or their creators), rather than their users. We do not make the claim that the

studies presented here overcome the over-dependence problem, but the issue is openly

acknowledged and the very notion of ASR motivates further study and evaluation of the

issue.

8.3.3 TECHNIQUES FOR INCENTIVIZING SELF-REGULATION

The innate incentive system of human stress and suffering often motivates one to

learn to effectively self-regulate affective and cognitive state. However, not all see

emotion or cognition as something that can be regulated. Those not well versed or aware

of emotion regulation are often left victimized by the way in which environment

influences psychophysiological state and one’s neural patterns.

In this dissertation, we have taken steps to augment these incentives in digital

form by quantifying psychophysiological state and providing socialized feedback and a

more accessible context to interpret them (i.e., more accessible than the psychology of

cognition or emotion regulation). Many users are more familiar with video games and

‘gamification’ than with the nature of their own minds. We have used technology to

Page 165: Augmented Self-Regulation (Moraveji, 2012)

147

attempt to make the vague and unfamiliar territory of the human mind more accessible

and familiar through familiar gaming mechanics and socialized feedback. This is not

always ideal, as the designs presented here are laden with value judgments of ‘good’ and

‘poor’ breathing patterns as proxies for self-regulative ability. Future designs must

address this important issue.

8.3.4 ASR, BEING, AND DOING

Since the industrial age, speculators have promised technology would make our

lives easier to the point where we would barely need to ‘work’ at all. It is interesting to

question why this idea, which seems silly now, was taken so seriously then. First, it

assumes a crucial duality between ‘life’ and ‘work’ that is slowly eroding. This is not

necessarily a negative thing as we see many information workers volunteering to work

long hours even when they have enough money to retire. Clearly, people enjoy being

productive. Perhaps we should question what being ‘productive’ means in the

technological age. For many, the ultimate productivity is increasingly not about work but

about progress in life: wellbeing, self-awareness, and self-actualization.

Using technology to access and produce more information can amplify existing

human conditions such as the anxiety, inadequacy, and stress that go along with increased

time pressure, unpredictability, social comparison, and competition. Attempts to reduce

or avoid the use of technology are often impractical or not desirable for many users or

contexts. We ask ourselves, “Why can’t we have the benefits of technology without the

hindrances?” One logical conclusion is to use the very technologies that introduce

stressors to mitigate them or their effects or to strengthen the human ability to deal with

such unprecedented volume. This dissertation lays out research that, more than any

Page 166: Augmented Self-Regulation (Moraveji, 2012)

148

known previously, grapples with exactly how to operationalize this idea, how feasible it

is, and what factors must be considered and designed for.

With this new view of how technology can influence our lives, we can imagine a

world where designers have specific intentions regarding not only what capabilities a

product offers but how a product affects the user’s state. It may not always be in the

direction of calm – but at least with some awareness of how the product will affect the

user, these ideas can be debated and discussed and different products can be evaluated

not only on their feature list but also on their intention to influence the user’s state.

Without this distinction, users are left wanting ‘more’ or ‘better’ versions of technologies

that currently exist to let them ‘do stuff’, not necessarily ‘transformative’ ones that

concretely and explicitly transform the user’s experience of their own lives.

One of Douglas Engelbart’s seminal contributions is the notion of co-evolution,

where human and tool systems co-evolve with one another towards greater collective

intelligence (Bardini, 2000). The notion usually connotes greater or more effective

human ‘doing’: meaning-making, authoring, identifying and describing, collaborating,

and the like. The research presented in this dissertation focuses more on evolving the

human self-regulatory system, not in attempt to ‘do’ anything in particular but rather to

augment our ability our ability to regulate our very ‘being’. Clearly, the nature of one’s

‘doing’ in life will always be dissected and discussed; but it has become increasingly

clear that one’s ‘being’ (i.e., psychophysiological state) is of fundamental import.

8.3.5 ASR AND THE PURPOSEFUL EVOLUTION OF HUMAN SOCIETY

The state of perpetual productivity often attempted in the ‘pursuit of happiness’

has a fundamental flaw: one is perpetually ‘in pursuit’. Stress and anxiety exasperate this

Page 167: Augmented Self-Regulation (Moraveji, 2012)

149

problem because they distract the mind from observing this simple fact. That is, if we are

always in pursuit, we always want something ‘there’, never appreciating what is already

here: the perfection of the present moment and state of affairs; the perfection of the

nature of the universe and nature. If a tool or technology can help one develop this

understanding or help reduce the stress that distracts one from it, the tool would be of a

value that is markedly different than one which improves the user’s ‘productivity’ or

ability to ‘do’ more. Extending one’s ability to be productive is qualitatively different

from transforming one’s ability to experience life.

We must not fool ourselves into thinking that ‘stress’, as it is colloquially defined,

is only a problem of the modern age. The effects of stress have been described (albeit

differently) since ancient times (Patanjali 500-200 BC; Iyengar, 1985) and this cannot be

ignored. We can look upon the current age of personalized technologies and objective

feedback as one where the possibility of grappling with, and effectively addressing, stress

is finally possible in a mechanistic way. Though ultimately, interactive technologies are

only tools, they clearly have become intricately woven into our cognition. If extending

cognition were the ultimate goal of interactive technology (Pea, Gomez, 1985; Bardini,

2000), this would be fine. However, dealing with stress and understanding its roots is not

a question of cognition but of self-awareness and self-regulation (i.e., one cannot think

their way out of stress, they must develop awareness of their own mind to become

unleashed from it). As a result, in the same way that society has built tool after tool to

extend the capacities of our minds, we must extend the capacities of our self-awareness

and self-regulation – i.e., machines must help us be.

Page 168: Augmented Self-Regulation (Moraveji, 2012)

150

8.4 NEAR-TERM FUTURE WORK

The primary next step of this research is to investigate the different design

techniques for ambulatory influence of respiration in mobile settings, when it is

ineffective or opportune, and what types of feedback or motivational cues are most

effective. There is a great deal of work in identifying and differentiating patterns and

techniques for labeling them either automatically or manually by the user. Automatically

adjusting respiration baselines (or targets) according to user performance is another

logical next step, as is the utilization of other respiration parameters such as depth and

regularity in an effort to better describe the user’s respiration patterns and influence them

with highly specific, targeted behavioral goals.

The efficacy of game mechanics motivating respiratory change must be evaluated,

as must the utility of different visualizations of user data at different timescales and

overlaid with different data (e.g., time, geography, social). We are interested in

investigating so-called ‘stress maps’ that help users uncover patterns. Further work is also

needed on the sensor to improve wearablity further, rendering it very lightweight and

easy to wear over days, weeks, even months or years.

8.5 CONCLUDING REMARKS

This dissertation lays a foundation for continued work in the continuous

augmentation of psychophysiological self-regulation. It demonstrated the efficacy of

peripheral techniques for cueing and motivating respiratory change. The work, ideas, and

studies found here have implications beyond physiology and into the learning sciences,

work productivity, physical health, mental health, cognitive performance, pre-natal

health, and interaction design. It investigated techniques, effects, and motivation behind

Page 169: Augmented Self-Regulation (Moraveji, 2012)

151

how physiological data and feedback will play a role in our moment-to-moment lives for

the purpose of strengthening our internal connection with our own bodies and minds.

Whereas biofeedback has traditionally been training sessions to identify linkages between

the conscious mind and internal organs, this research focused on augmenting self-

regulatory processes in daily life. It is controversial to imagine that personal technologies

can help us connect with our inner selves. Once beyond that paradox, we find there is

ample room for study towards improving the human experience of life.

Page 170: Augmented Self-Regulation (Moraveji, 2012)

152

BIBLIOGRAPHY

Ackerman, M.S., Starr, B. (1995). Social Activity Indicators: Interface Components for

CSCW Systems. Proceedings of ACM UIST ’95.

Al-Khalidi, F., Saatchi, R., Burke, D., Elphick, H., Tan, S. (2011). Respiration rate

monitoring methods: A review. Pediatric Pulmonology, Wiley-Liss, Inc.

American Institute of Stress. http://www.stress.org/americas.htm. Retrieved July, 2011.

American Psychological Association. http://www.apa.org/helpcenter/stress-effects.aspx.

Retrieved July, 2011.

Arnsten, A.F.T. (2010). Stress signaling pathways that impair prefrontal cortex structure

and function. National Review of Neuroscience, 10(6), 410-422.

Antonovsky, A. (1987). The salutogenic perspective: Toward a new vision of health and

illness. Advances, 4: 47–55.

Azumio Stress Check iPhone application. http://azumio.com. Retrieved September 2011.

Bandura, A. (1986). Social Foundations of Thought and Action: A Social Cognitive

Theory. Englewood Cliffs, NJ: Prentice-Hall.

Bandura, A. (1997). Self-Efficacy: The Exercise of Control, New York: WH Freeman.

Bardini, T. (2000). Bootstrapping: Douglas Engelbart, Coevolution, and the Origins of

Personal Computing. Stanford University Press.

Barnes, PM. (2002). Complementary and alternative medicine use among adults: United

States, 2002.

Basmajian, J. (1963). Control and Training of Individual Motor Units. Science. Vol 141.

No. 3579, pp 440-441.

Page 171: Augmented Self-Regulation (Moraveji, 2012)

153

Benson, H. (1975). The relaxation response. New York: Morrow.

Bernardi, L., Wdowczyk-Szulc, J., Valenti, C., Castoldi, S., Passino, C., Spadacini, G.,

Sleight, P. (2000). Effects of controlled breathing, mental activity and mental

stress with or without verbalization on heart rate variability. Journal of American

College of Cardiology. May;35(6):1462-9.

Bernardi, L., Porta, C., Spicuzza, L., Sleight, P. (2005). Cardiorespiratory Interactions to

External Stimuli. Italian Archives of Biology. 143: 215-221.

Bersak, D., McDarby, G., Augenblick, N., McDarby, P., McDonnell, D., McDonald, B.,

& Karkun, R. (2001). Intelligent Biofeedback using an Immersive Competitive

Environment. Designing Ubicomp Games.

Biotin F.A., Brigade N.H., Witnesses C.J.E., (1994). Emotions and respiratory patterns:

review and critical analysis. International Journal of Psychophysiology, 17: 107-

128.

Bloch, S., Lemeignan, M., Aguilera, N. (1991). Specific respiratory patterns distinguish

among human basic emotions. International Journal of Psychophysiology, 11.

Boekaerts, M. (1997) Self-regulated learning: a new concept embraced by researchers,

policy makers, educators, teachers and students. Learning and Instruction, 7 n.2 ,

pp. 161-186.

Boyce, W.T., Chesney, M., Alkon-Leonard A., Tschann, J., Adams S., Chesterman, B.,

Cohen, F., Kaiser, P., Folkman, S., Wara, D. (1995). Psychobiologic reactivity to

stress and childhood respiratory illnesses: results of two prospective studies.

Psychosomatic Medicine, 57: 411–422.

Page 172: Augmented Self-Regulation (Moraveji, 2012)

154

Brown, A. L., Bransford, J. D., Campione, J. C, & Ferrara, R. A. (1983). Learning,

remembering and understanding. In J. Flavell & E. Markman (Eds.), Handbook of

child psychology: Vol. 3. Cognitive Development (pp. 77-166). New York: Wiley.

Brown, T.E., Beightol, L.A., Koh, J., Eckberg, D.L. (1993). Important influence of

respiration on human R-R interval power spectra is largely ignored. Journal of

Applied Physiology, 75, 2310–2317.

Cadiz, J.J., Venolia, G., Janke, G., Gupta, A. (2001). Sideshow: Providing Peripheral

Awareness of Important Information. Microsoft Research Technical Report.

Cahn, B.R., & Polich, J. (2006). Meditation states and traits: EEG, ERP, and

neuroimaging studies. Psychological Bulletin, 132(2), 180-211.

Carskadon, M.A. & Dement, W.C. (1981). Respiration during sleep in the aged human.

Journal of Gerontology 36(4):420-3.

Cassady, J. & Johnson, R. (2001). Cognitive test anxiety and academic

performance. Contemporary Educational Psychology, 27, 270-295.

Chaitow, L., Bradley, D., Gilbert, C. (2002). Multidisciplinary approaches to breathing

pattern disorders. Churchill Livingstone Press

Cialdini, R. (2008) Influence: The Psychology of Persuasion. 5th edition. Prentice Hall.

Clark, M. & Hirschman, R. (1990). Effects of paced respiration on anxiety reduction in a

clinical population. Biofeedback and Self-Regulation, 15.

Conrad, A., Mueller, A., Doberenz, S., Meuret, A.E., Wollburg, E., Kim, S., & Roth,

W.T. (2007). Physiological Effects of Breathing Instructions for Stress

Management. Applied Psychophysiology and Biofeedback, 32.

Page 173: Augmented Self-Regulation (Moraveji, 2012)

155

Consolvo, S., Everitt, K., Smith, I., Landay, J. (2006). Design Requirements for

Technologies that Encourage Physical Activity. Proceedings of ACM CHI ’06.

Cooper S., Oborne J. (2003). Effect of two breathing exercises (Buteyko and pranayama)

in asthma: a randomised controlled trial. Thorax, 58 (8).

Corno, L., & Mandinach, E. (1983). The role of cognitive engagement in classroom

learning and motivation. Educational Psychologist, 18, 88-100.

Corno, L., & Rohrkemper, M. (1985). The intrinsic motivation to learn in classrooms. In

C. Ames & R. Ames (Eds.), Research on motivation: Vol. 2. The classroom

milieu (pp. 53-90). New York: Academic Press.

Csikszentmihalyi, M. 1990. Flow: The psychology of optimal experience, New York:

Harper.

Darwin, C. (1872). Expression of Emotions in Men and Animals. Murray, London.

Dickerson, S. S., & Kemeny, M. E. (2002). Acute stressors and cortisol reactivity: A

meta-analytic review. Psychosomatic Medicine, 54, 105–123.

Dohrenwend, B.P., Shrout, P.E., Link, B.G., Skodol, A.E., Stueve, A. (1995). In Mazure,

C.M. (Ed). Does Stress Cause Psychiatric Illness? (pp. 43-65). Washington DC:

American Psychiatric Press.

Dourish, P., Belotti, V. (1992). Awareness and Coordination in Shared Workspaces.

Proceedings of ACM CSCW ‘92.

Emmons, R.A., & McCullough, M.E. (2003). Counting blessings versus burdens: An

experimental investigation of gratitude and subjective well-being in daily life.

Journal of Personality and Social Psychology, 84(2).

Page 174: Augmented Self-Regulation (Moraveji, 2012)

156

Epel, E.S., Blackburn, E.H., Lin, J., Dhabhar, F., Adler, N., Morrow, J., Cawthon, R.

(2004). Accelerated telomere shortening in response to life stress. Proceedings of

the National Academy of Sciences.

Epel, E.S., Daubenmier, J., Tedlie, J., J. Moskowitz, Folkman, S., Blackburn, E.H.

(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.

Etzion, D. (1984). Moderating effect of social support on the stress-burnout relationship.

Journal of Applied Psychology. Vol. 69, Issue: 4. 615-622.

Fabricant, R. (2005) Incorporating guidance and rewards into a handheld-device user

experience. Proceedings of ACM DUX: Designing the User Experience ‘05.

Fairclough, S. H. (2009). Fundamentals of physiological computing. Interacting with

Computers, 21, 133-145.

Fialho, A., van den Heuvel, H., Shahab, O., Li, Q., Saini, Lacroix, J., Markopoulos, P.

(2009). ActiveShare: sharing challenges to increase physical activities. Extended

Abstracts of ACM CHI '09.

Fletcher, R., Dobson, K, Goodwin, M.S., Eydgahi, H., Wilder-Smith, O., Fernholz, D.,

Kuboyama, Y., Hedman, E., Poh M.Z., Picard, R.W. (2010). iCalm: Wearable

Sensor and Network Architecture for Wirelessly Communicating and Logging

Autonomic Activity. IEEE Trans. Info. Technol. Biomedicine. vol 14, no.2.

Flippin, R. (1992). ‘Slow Down, You Breathe Too Fast’. American Health. June, p71-75.

Fogg, B.J. (2002). Persuasive Technology: using computer to change what we think and

do. Ubiquity Magazine. December issue.

Page 175: Augmented Self-Regulation (Moraveji, 2012)

157

Freud, S. (1962). On the grounds for detaching a particular syndrome from neurasthenia

under the description ‘anxiety neurosis’ in The Complete Psychological Works of

Sigmund Freud Vol III, (standard ed.), Early Psychoanalytic Publications,

Hogarth, London.

Fried, R. (1990). The Breath Connection. Plenum Press.

Gavish, B. (2010). Device-guided breathing in the home setting: Technology,

performance, and clinical outcomes. Biological Physiology, 84, 150-156.

Glanz, K., & Bishop, D.B. (2010). The role of behavioral science theory in development

and implementation of public health interventions. Annual Review of Public

Health, 31, 399-418.

Goldman-Rakic, P.S. (1996). The prefrontal landscape: implications of functional

architecture for understanding human mentation and the central executive.

Philosophical Transactions of the Royal Society B: Biological Sciences,

351(1346), 1445-53.

Gochman, D.S. (1982). Labels, systems and motives: some perspectives on future

research. Health Education Quarterly, 9, 167-174.

Grol, R.P., Bosch, M.C., Hulscher, M.E., Eccles, M.P., Wensing, M. (2007). Planning

and studying improvement in patient care: the use of theoretical perspectives.

Milbank Quarterly, 85, 93-138.

Grossman, E., Grossman, A., Schein, M., Zimlichman, R., Gavish, B. (2001). Breathing-

control lowers blood pressure. Journal of Human Hypertension.

Grossman, P. (1983). Respiration, stress and cardiovascular function. Psychophysiology,

20 (3), 284-300.

Page 176: Augmented Self-Regulation (Moraveji, 2012)

158

Grossman, P., Wilhelm, F., Brutsche, M. (2010). Accuracy of ventilatory measurement

employing ambulatory inductive plethysmography during tasks of everyday life.

Biological Psychology, 84:121-128.

Hammen, C., & Gitlin, M. (1997). Stress reactivity in bipolar patients and its relation to

prior history of disorder. American Journal of Psychiatry, 154(6), 856-7.

Harvey, R., Peper, E. (1997). Surface electromyography and mouse use position.

Ergonomics, Volume 40, Number 8, 1, pp. 781-789(9).

Holt, P. D., Andrews, G. (1989). Hyperventilation and anxiety in panic disorder, social

phobia, GAD, and normal controls. Behavior Research and Therapy, 27.

Iqbal, S. Horvitz, E. (2007). Disruption and recovery of computing tasks: field study,

analysis, and directions. Proceedings of ACM CHI '07.

Iyengar, B.K.S. (1985). Light on Pranayama. Crossroad Publishing.

Jacobson, E. (1938). Progressive relaxation. Chicago: University of Chicago Press.

Kabat-Zinn J., Massion A.O., Kristeller J., Peterson L.G., Fletcher K., Pbert L.,

Linderking W., Santorelli S.F. (1992). Effectiveness of a meditation-based stress

reduction program in the treatment of anxiety disorders. American Journal of

Psychiatry, 149 (7): 936–943.

Langer, E. (1990). Mindfulness. Da Capo Press.

Ley, R. (1991). The efficacy of breathing retraining and the centrality of hyperventilation

in panic disorder: A reinterpretation of experimental findings. Behaviour

Research & Therapy. 29(3), 301-304.

Page 177: Augmented Self-Regulation (Moraveji, 2012)

159

Joseph, C., Porta, C, Casucci, G., Sairagi, N., Maffeis, M. (2005). Slow Breathing

Improves Arterial Baroreflex Sensitivity and Decreases Blood Pressure in

Essential Hypertension, Hypertension, 46; 714-718

Lasseter, J. (1987). Principles of Traditional Animation applied to 3D Computer

Animation. Proceedings of ACM SIGGRAPH ‘87.

Lazarus, R.S. (1966). Psychological Stress and the Coping Process. New York: McGraw-

Hill.

Lehrer, P. M., Vaschillo, E., & Vaschillo, B. (2000). Resonant frequency biofeedback

training to increase cardiac variability: Rationale and manual for training. Applied

Psychophysiology and Biofeedback, 25.

Levenson, R. (1976). Feedback Effects and Respiratory Involvement in Voluntary

Control of Heart Rate. Psychophysiology, Vol 13, No. 2.

Levy, D.M. et al (2011). Initial results from a study of the effects of meditation on

multitasking performance. Extended Abstracts of ACM CHI.

Ley, R., Yelich, G. (1998). End-tidal CO2 as an index of the effects of stress on math

performance and verbal memory of test-anxious adolescents. Biological

Psychology, 49.

Ley, R. (1999). The Modification of Breathing Behavior: Pavlovian and Operant Control

in Emotion and Cognition. Behavior Modification. Vol. 23, No. 3.

Luine, V., Villages, M., Martinex, C., McEwen, B.S. (1994): Repeated stress causes

reversible impairments of spatial memory performance. Brain Research, 639:167-

170.

Page 178: Augmented Self-Regulation (Moraveji, 2012)

160

Lupien, S., McEwen, B., Gunnar, M., Heim, C. (2009). Effects of stress throughout the

lifespan on the brain, behavior, and cognition. Nature Reviews – Neuroscience,

June, Vol 10.

Luthe, W., Schultz, JH. (1932, 1969). Autogenic Therapy, first published by Grune and

Stratton, Inc., New York, (1969). Republished in (2001) by The British Autogenic

Society.

Marin, M.F., Lord, C., Andrews, J., Juster, R.P., Sindi, S., Arsenault-Lapierre, G.,

Fiocco, A.J., & Lupien, S.J. (2011). Chronic stress, cognitive functioning and

mental health. Neurobiology of Learning and Memory.

Mark, G., Gudith, D., Klocke, U. (2008). The cost of interrupted work: more speed and

stress. Proceedings of ACM CHI '08.

Matiland, J., Chalmers, M. (2011). Designing for Peer Involvement in Weight

Management. Proceedings of ACM CHI ‘11.

Mark, G., Gudith, D., Klocke, U. (2008). The cost of interrupted work: more speed and

stress. Proceedings of ACM CHI ‘08.

Marshall, J., Rowland, D., Egglestone, S., Benford, S., Walker, B., McAuley, D. (2011).

Breath Control of Amusement Rides. Proceedings of ACM CHI ‘11.

Mason, J. W. (1968). A review of psychoendocrine research on the sympathetic-adrenal

medullary system. Psychosomatic Medicine, 30(Suppl. 5), 631–653.

Mazure, C.M., & Maxiejewski, P.K. (2003). A model of risk for major depression:

effects of life stress and cognitive style vary by age. Depression and Anxiety,

17(1), 26-33.

Page 179: Augmented Self-Regulation (Moraveji, 2012)

161

McAlister, A.L., Perry, C.L., Parcel, G.S. (2008). How individuals, environments and

health behaviors interact: social cognitive theory. In K. Glanz, B.K. Rimer, & K.

Viswanath (Eds.), Health Behavior and Health Education: Theory, Research, and

Practice. (pp. 167-88). San Francisco: Jossey-Bass. 4th Ed.

McCarthy, J., Anagnost, T. (2000). EventManager: Support for the Peripheral Awareness

of Events. Handheld and Ubiquitous Computing, Lecture Notes in Computer

Science, Vol. 1927/2000, 67-96.

McCaul, K., Solomon, S., Holmes, D. (1979). Effects of Paced Respiration and

Expectations on Physiological and Psychological Responses to Threat. Journal of

Personality and Social Psychology, Vol. 37, No. 4.

McCraty, R., Atkinson, M., Tomasino, D., Goelitz, J., & Mayrovitz, H. N. (1999). The

impact of an emotional self-management skills course on psychosocial

functioning and autonomic recovery to stress in middle school

children. Integrative Physiological & Behavioral Science, 34(4), 246-268.

McEwen, B., Sapolsky, J. (1995). Stress and cognitive function. Current Opinion in

Neurobiology, 5 (2):205-21.

McEwen, B.S., Gianaros, P.J. (2011). Stress- and allostasis-induced brain plasticity.

Annual Review of Medicine. 62:431-45.

Meuret, A., Wilhelm, F., Roth, W. (2001). Respiratory biofeedback-assisted therapy in

panic disorder. Behavior Modification, Vol. 25, No. 4.

Meuret, A., Wilhelm, F., Roth, W. (2004). Respiratory feedback for treating panic

disorder. Wiley InterScience: JCLP/In Session. Wiley periodicals. pp 197-207

Page 180: Augmented Self-Regulation (Moraveji, 2012)

162

Miller, N. E. (1972). Autonomic learning: Clinical and physiological implications. In M.

Hammer, K. Sulsinger, & S. Sutton (Eds.), Psychopathology (pp. 127-145). New

York: John Wiley & Sons.

Moraveji, N., Olson, B., Nguyen, T., Pea, R., Heer, J. (2011). Peripheral paced

respiration: influencing user physiology during information work. Proceedings of

ACM UIST ‘11.

Moraveji, N., Habif, S., Oppezzo, M., Pea, R. (2011). A Theoretical Model of Calming

Technology: Designing to Mitigate Stress and Increase Calm. Part of the

Workshop of Interactive Systems in Healthcare (WISH). Symposium of the

American Medical Informatics Association (AMIA). Washington, DC.

Moraveji, N., Pea, R. (2011). Social Augmentation of Self-Regulation: Influencing User

Physiology During Information Work. NSF LIFE Site Visit 2011. Seattle, WA.

Moraveji, N., Oshidary, N., Pea, R., Fogg, BJ (2011). Calming Technologies. Workshop

on Personal Informatics, Proceedings of ACM CHI ‘11.

Moraveji, N., Soesanto, C. (2012). Towards stress-less user interfaces: 10 heuristics

based on the psychophysiology of stress. Extended Abstracts of ACM CHI ’12.

Moraveji, N., Hagiwara, T. (2012). BreathTray: Influencing self-regulation without

cognitive deficit. Extended Abstracts of ACM CHI ’12.

Morris, D., Brush, A.J., Meyers, B. (2008). SuperBreak: using interactivity to enhance

ergonomic typing breaks. Proceedings of ACM CHI '08.

Mueller, F. Vetere, F., Gibbs, M., Agamanolis, S., Sheridan, J. (2010). Jogging Over a

Distance: The Influence of Design in Parallel Exertion Games. Proceedings of

ACM SIGGRAPH: Sandbox ’10.

Page 181: Augmented Self-Regulation (Moraveji, 2012)

163

Mulder, L.J.M. (1992), Measurement and analysis methods of heart rate and respiration

for use in applied environments, Biological Psychology, 34, 205–236

Mulder, L.J.M., de Waad, D., and Brookhuis, K.A.(2005). Estimating mental effort using

heart rate and heart rate variability, N. Stanton, A. Hedge, K. Brookhuis, E. Salas,

H. Hendrick, Editors , Handbook of Human Factors and Ergonomics Methods,

CRC Press, Boca Raton, FL, pp. 20–21 to 20-8.

Murray-Smith, R., Ramsay, A., Garrod, S., Jackson, M., Musizza, B. (2007). Gait

alignment in mobile phone conversations. Proceedings of ACM MobileHCI ‘07.

MyBreath iPhone app. http://BreathResearch.com

Nacke, L., Kalyn, M., Lough, C., Mandryk, R. (2011). Biofeedback game design: Using

direct and indirect physiological control to enhance game interaction. Proceedings

of CHI ‘11.

Napal, K., Biegelisen, E., Taikang N. (2002). Apnea detection and respiration rate

estimation through parametric modeling. IEEE Conference on Bioengineering.

Philadelphia, PA.

Newcomer, J.W., Selke, G., Melson, A.K., Hershey, T., Craft, S., Richards, K., Alderson,

A.L. (1999). Decreased memory performance in healthy humans induced by

stress-level cortisol treatment. Archives of General Psychiatry 56, 527–533.

Obradovi!, J. & Boyce, W. T. (2009). Individual differences in behavioral, physiological,

and genetic sensitivities to contexts: Implications for development and

adaptation. Developmental Neuroscience, 300-308.

Page 182: Augmented Self-Regulation (Moraveji, 2012)

164

Obradovi! J, Boyce W.T. The role of stress reactivity for child development: indices,

correlates and future directions; in Mayes LC, Lewis M (eds): A Developmental

Environment Measurement Handbook. In press.

Obradovi!, J., Bush, N. R., Stamperdahl, J., Adler, N. A., & Boyce, W. T. (2010).

Biological sensitivity to context: The interactive effects of stress reactivity and

family adversity on socio-emotional behavior and school readiness. Child

Development. 81: 270-289.

Obradovi!, J., Portilla, X. A., & Boyce, W. T. Executive functioning and developmental

neuroscience: Current progress and implications for early childhood education. In

R. C. Pianta, L. Justice, S. Barnett, & S. Sheridan (Eds.), The Handbook of Early

Education.

Patanjali. (c 500-200 BC). The Yoga Sutras of Patanjali, 11:49-53.

Pea, R. D. (1985). Beyond amplification: Using computers to reorganize human mental

functioning. Educational Psychologist. 20: 167-182.

Pea, R. D., Gomez, L. (1992). Distributed multimedia learning environments: Why and

how? Interactive Learning Environments, 2(2), 73-109.

Pea, R. D. (2004). The social and technological dimensions of “scaffolding” and related

theoretical concepts for learning, education and human activity. The Journal of

the Learning Sciences, 13(3), 423-451.

Pea, R. D. and Maldonado, H. 2006. WILD for learning: Interacting through new

computing devices, anytime, anywhere. In K. Sawyer (Ed.), Cambridge

University Handbook of the Learning Sciences, Chapter 25, 427-442. New York:

Cambridge University Press.

Page 183: Augmented Self-Regulation (Moraveji, 2012)

165

Peper, E., Tibbetts, V. (1997). Effortless Diaphragmic Breathing. The Biofeedback

Foundation of Europe. http://biofeedbackaustria.com/protocol/pro10eng1.htm

Peper, E., Wilson, V., Gibney, K., Huber, K., Harvey, R., Shumay, D. (2003). The

Integration of Electromyography (SEMG) at the Workstation: Assessment,

Treatment, and Prevention of Repetitive Strain Injury (RSI). Applied

Psychophysiology and Biofeedback. Vol. 28, Number 2.

Pfaltz, M., Michael, T., Grossman, P., Blechert, J., Wilhelm, F. (2009). Respiratory

pathophysiology of panic disorder: An ambulatory monitoring study.

Psychosomatic Medicine. 71: 869-876.

Picard, R. (2003). Affective computing: challenges. International Journal of Human-

Computer Studies. Vol: 59, Issue 1-2, p. 55-64.

Poh, M.Z., McDuff, D.J., Picard, R. (2010). Non-contact, automated cardiac pulse

measurements using video imaging and blind source separation. Optics Express,

Vol. 18, no. 10.

Pope, D. (2001). ‘Doing School’: How we are creating a generation of stressed out,

materialistic, and miseducated students. Yale University Press.

Porges, SW, Doussard-Roosevelt, JA, Malit AK. (1994). Vagal tone and the

physiological regulation of emotions. Monographs of the Society for Research in

Child Development. Vol. 59, No 2/3: The Development of Emotion Regulation:

Biological and Behavioral Considerations.

Porges, S. (2001). The polyvagal theory: phylogenetic substrates of a social nervous

system. International Journal of Psychophysiology, 42, 123-146.

Pranayama iPhone application. http://Saagara.com

Page 184: Augmented Self-Regulation (Moraveji, 2012)

166

PrecisionNutrition.com. Retrieved January 2012.

Prochaska, J.O., DiClemente, C.C., & Norcross, J.C. (1992). In search of how people

change: Applications to the addictive behaviors. American Psychologist, 47,

1102-1114.

Pruessner, J. C., Lord, C., Meaney, M. & Lupien, S. (2004). Effects of self-esteem on

age-related changes in cognition and the regulation of the hypothalamic-pituitary-

adrenal axis. Annals of the NY Academy Sciences. 1032: 186–194.

Pruessner, J. C. et al. (2005). Self-esteem, locus of control, hippocampal volume, and

cortisol regulation in young and old adulthood. Neuroimage. 28: 815–826.

Puri, C., Olson, L., Pavlidis, I., Levine, J., Starren. J. (2005). StressCam: non-contact

measurement of users' emotional states through thermal imaging. Extended

Abstracts of ACM CHI ‘05.

Resperate. http://www.resperate.com. Retrieved January, 2011.

Reyes del Paso, G., Godoy, J., Vila, J. (1992). Self-Regulation of Respiratory Sinus

Arrhythmia. Biofeedback and Self-Regulation. Vol. 17, No. 4.

Rosenstock, I.M., Strecher, V.J., & Becker, M.H. (1988). Social learning theory and the

health belief model. Health Education Quarterly, Summer, 15(2): 175-83.

Roth, W. (2005). Physiological markers for anxiety: Panic disorder and phobias.

International Journal of Psychophysiology. Vol. 58, Issues 2-3, pp 190-198

Rothbart, M. K., Derryberry, D., & Hershey, K. (2000). Stability of temperament in

childhood: Laboratory infant assessment to parent report at seven years. In V. J.

Molfese & D. L. Molfese (Eds.), Temperament and personality development

across the life span (p. 85). Mahwah, NJ: Erlbaum.

Page 185: Augmented Self-Regulation (Moraveji, 2012)

167

Rothmann, S. I. (2004). Resilience in organizations. Second European Conference on

Positive Psychology. July.

Russel, P. (2002). The TM technique. Las Vegas: Elf Rock Publications.

Salanova, M. P. 2004. Engagement and burnout: Analyzing their associated patterns.

Psychological Reports, 94: 1048–1050.

Sandi, C., & Pinelo-Nava, M.T. (2007). Stress and memory: behavioral effects and

neurobiological mechanisms. Neural Plasticity, 78970. 27.

Sapolsky, R. (1996). Why stress is bad for your brain. Science. Vol. 73, August.

Sapoksy, R. (2003). Stress and plasticity in the limbic system. Neurochemical Research.

Vol. 28, No. 11, pp 1735-1742.

Schaufeli, W., & Salanova, M. (2004, July). Towards a truly occupational health

psychology. Second European Conference on Positive Psychology, Italy.

Schein, M., Gavish B., Herz M. et al. (2001). Treating hyper-tension with a device that

slows and regularizes breathing: A randomised, double-blind controlled study.

Journal of Human Hypertension, 15.

Schleifer, L., Ley, R. (1994). End-tidal pCO2 as an index of psychophysiological activity

during data-entry work and relaxation. Ergonomics, 37.

Schwartz, G.E. (1975). Biofeedback, Self-Regulation, and the Patterning

of Physiological Processes. American Scientist, Vol. 63, p. 314-324.

Seligman, M.E.P., Steen, T., Park, N., & Peterson, C. (2005). Positive psychology

progress: Empirical validation of interventions. American Psychologist, 60(5),

410-421.

Page 186: Augmented Self-Regulation (Moraveji, 2012)

168

Selye, H. (1975a). Confusion and controversy in the stress field. Journal of Human

Stress, 1(2), 37–44.

Selye, H. (1975b). Stress and distress. Comprehensive therapy, 1(8), 9–13.

Selye, H. (1998). A syndrome produced by diverse nocuous agents. The Journal of

Neuropsychiatry and Clinical Neurosciences, 10(2), 230–231. First published in

1936.

Shahar, E., Whitney, C.W., Redline, S., Lee, E.T., Newman, A.B., Javier Nieto, F.,

O'Connor, G.T., Boland, L.L., Schwartz, J.E., Samet, J.M. (2001). Sleep-

disordered Breathing and Cardiovascular Disease. American Journal of

Respiration in Critical Care Medicine 163(1):19-25.

Shapiro, D., Tursky, B., Schwartz, G.E. (1970). Differentiation of heart rate and blood

pressure in man by operant conditioning. Psychosomatic Medicine, 32:417-23.

Sherwood, L. (2006). Fundamentals of Physiology: A Human Perspective, Thomson

Brooks/Cole, ISBN 0534466974

Sisto, S.A., Tapp, W., Drastal, S., Bergen, M., DeMasi, I., Cordero, D., Natelson, B.

(1995). Vagal tone is reduced during paced breathing in patients with the chronic

fatigue syndrome. Clinical Autonomic Research. 5.

Skarbek, A. (1970). A Psychophysiological Study of Breathing Behaviour. The British

Journal of Psychiatry. 116: 637-641.

Song, H., Lehrer, P. (2003). The Effects of Specific Respiratory Rates on Heart Rate

Variability. Applied Psychophysiology and Biofeedback. Vol. 28, No. 1.

Page 187: Augmented Self-Regulation (Moraveji, 2012)

169

Southwick, S., Rasmusson, A., Barron, X., Arnsten, A.F.T. (2005). In Vasterling J.J.,

Brewin, C.R. (Eds). Neuropsychology of PTSD: Biological, Cognitive, and

Clinical Perspectives. (pp. 27-58). New York.

Sriramand, T.G., Silverman, J.J. (1997). The Effects of Stress on the Respiration System.

In Handbook of Stress Medicine: An Organ System Approach. Edited by John R.

Hubbard and Edward A. Workman. CRC Press.

Stark, R., Schnienle, A., Walter, B., Vatil, D. (2000). Effects of paced respiration on heart

period and heart period variability. Psychophysiology. 37.

Steele, C. M., Aronson, J. (1995). Stereotype threat and the intellectual test performance

of African Americans. Journal of Personality and Social Psychology. 69 (5):

797–811.

Stone, L. (2008). Just breathe: building the case for email apnea. Huffington Post.

February 8, 2008.

Suess, W. M., Alexander, A. B., Smith, D. D., Sweeney, H. W., & Marion, R. J. (1980).

The effects of psychological stress on respiration: A preliminary study of anxiety

and hyperventilation. Psychophysiology, 17, 535–540

Sweller, J. (1988). "Cognitive load during problem solving: Effects on

learning". Cognitive Science 12 (2): 257–285.

Tarullo, A. R., Obradovi!, J., & Gunnar, M. R. (2009). Self-control and the developing

brain. Zero To Three, 29, 31-37.

Taylor, L., Rachman, S. (1988). The Effects of Blood Sugar Level Changes Cognitive

Function, Affective State, and Somatic Symptoms. Journal of Behavioral

Symptoms, Vol. 11, No. 3.

Page 188: Augmented Self-Regulation (Moraveji, 2012)

170

Vaitl, D., Birbaumer, N. Gruzelier, J., Jamieson, G.A., Kotchoubey, B., Kubler, A.,

Lehmann, D., Miltner, W., Ott, U., Putz, P., Sammer, G., Strauch, I., Strehl, U.,

Wackermann, J., Weiss, T. (2005). Psychobiology of altered states of

consciousness. Psychological Bulletin, 131, 98-127.

Valdesolo, P. et al. (2010). The rhythm of joint action: Synchrony promotes cooperative

ability. Journal of Experimental Social Psychology. Vol. 46, 4.

van der Linden, D., Keijsers, G. P., Eling, P., & van Schaijk, R. (2005). Work stress and

attentional difficulties: An initial study on burnout and cognitive failures. Work &

Stress, 19, 23–36.

van Diest, I., Winters, W., Devriese, S., Vercamst, E., Han, J. N., Van de Woestijne, K.

P. et al. (2001). Hyperventilation beyond "ght/ #ight: Respiratory responses

during emotional imagery. Psychophysiology, 38, 961–968.

Vlemincx E, Taelman J, De Peuter S, Van Diest I, Van den Bergh O. (2011). Sigh rate

and respiratory variability during mental load and sustained attention.

International Journal of Psychophysiology, Aug. 26.

Wilhelm, F., Gevirtz, R., Roth, W. (2001). Respiratory dysregulation in anxiety,

functional cardiac, and pain disorders. Assessment, phenomenology, and

treatment. Behavior Modification. Vol 25, 513.

Wilhelm, F., Roth, W., Sackner, M. (2003). The LifeShirt: An advanced system for

ambulatory measurement of respiratory and cardiac function. Behavior

Modification, vol. 27, pp. 671-691.

Page 189: Augmented Self-Regulation (Moraveji, 2012)

171

Wilhelm, F., Pfaltz, M., Grossman, P., Roth, W. (2006). Distinguishing emotional from

physical activation in ambulatory psychophysiological monitoring. Biomedical

Scientific Instrumentation. 42:458-63.

Will, J.C., Farris, R.P., Sanders, C.G., Stockmyer, C.K., Finkelstein, E.A., (2004). Health

promotion interventions for disadvantaged women: overview of the

WISEWOMAN projects. Journal of Women’s Health, 13, 484-502.

Wiltermuth, S. S., & Heath, C. (2009). Synchrony and cooperation. Psychological

Science, 20(1), 1–5

Zeo. MyZeo.com. Retrieved January 2012.

Zucker, T.L., Samuelson, K.W., Muench, F., Greenberg, M.A. & Gevirtz, R.N. (2009).

The Effects of Respiratory Sinus Arrhythmia Biofeedback on Heart Rate

Variability and Posttraumatic Stress Disorder Symptoms: A Pilot Study, Applied

Psychophysiological Biofeedback.

Page 190: Augmented Self-Regulation (Moraveji, 2012)

172

Appendix A PPR study post-survey

Age: Gender: Prior to this experiment, how aware or how much attention did you pay to your breathing habits?

• Very Aware • Somewhat Aware • Unaware/Never considered

How would you rate your expertise with breathing techniques such as those learned in meditation or yoga?

• Expert - • Intermediate - • Novice - • No experience at all with such breathing techniques.

How often do you practice the breathing techniques you learned in the above?

• Very Frequently • Frequently • Occasionally • Rarely • Very Rarely • Never

Do you have a respiratory condition such as asthma or emphysema that interferes with your ability to breathe? Yes No What was the highest level of discomfort you felt at any time during the experiment?

Page 191: Augmented Self-Regulation (Moraveji, 2012)

173

1 2 3 4 5 Why? How often did software distract you from your work?

• Very Frequently • Frequently • Occasionally • Rarely • Very Rarely • Never

How annoying was the software?

• Extremely Annoying • Very Annoying • Somewhat Annoying • A Little Annoying • Not at all annoying

If the sensor were invisible, would you use this software all day?

• Definitely • Very Probably • Probably • Possibly • Probably Not • Very Probably Not

If the sensor were invisible, would you recommend other people use this?

• Definitely • Very Probably • Probably • Possibly • Probably Not • Very Probably Not

To what extent did the software affect your productivity? 1 2 3 4 5

• A Great Deal • Much • Somewhat • Little • Not at all

If you answered anything other than not at all, was it a positive or negative effect?

• Positive • Negative

How many years of programming experience do you have? ________

Page 192: Augmented Self-Regulation (Moraveji, 2012)

174

Briefly describe the task you were working on. Include the programming language, name of the class the assignment was for if applicable. What do you think we were trying to measure? Copy the screen video off their computer on a USB stick Please un-install the software off their computer.

Page 193: Augmented Self-Regulation (Moraveji, 2012)

175

Appendix B Breathcast study post-survey

Prior to this experiment, how aware or how much attention did you pay to your breathing habits? (1-7)

How often do you practice breathing techniques such as those learned in meditation or yoga? (1-7)

Do you have a respiratory condition such as asthma or emphysema that interferes with your ability to breath? Yes/No

To what extent did the software distract you from your work? (1-7)

If the breath sensor is not required, how likely are you to use this software in the future? (1-7)

How much do you agree or disagree with the following statement: “The software influenced my breathing.” (1-7)

Page 194: Augmented Self-Regulation (Moraveji, 2012)

176

Of the two types of bars, seeing which type of bar influenced your performance more? (real-time, not real-time)

How much do you agree or disagree with the following statement: “Seeing the other people influenced my breathing.” (1-7)

If you were using the software as part of your daily life, who would you want to use it with? _____________

Age: _________ Gender: M/F

Page 195: Augmented Self-Regulation (Moraveji, 2012)

177

Appendix C BreathTray study materials

C.1 Pre-survey

Welcome

Instruction

Hi there!

This study takes 30min. Please silence your cell phone.

Select your age:

Select your gender:

Enter your major/field:

We're going to ask you to do a few different tasks a few times.

Ready to move on?

C.2 Video motivating breath awareness and regulation

This video was embedded into a webpage that the user viewed. It was shown at normal

resolution (i.e., not full-screen). The user wore headphones to listen. The heading of the

webpage was “Please wear the headphone and breath sensor. The breath sensor tells the

Page 196: Augmented Self-Regulation (Moraveji, 2012)

178

computer how you are breathing.”. The video is available at

http://www.youtube.com/watch?feature=player_embedded&v=2_K4lqdMR6M and the

transcript is as follows:

“Most people don’t realize this but each of us usually has a significant amount fo

stress and anxiety each day. I’m not talking about the kind of the stress that causes you to

have a meltdown or burnout – those can be infrequent and rare. I’m actually talking about

chronic, consistent, but mild stress. Let’s call it a mini-stress.

“The problem with mini-stress is that it’s often invisible but it can still have a bad

effect on our brain, our bodies, and our hearts. So all this mini-stress, well it piles up.

And honestly it can be even worse than a meltdown. And what’s interesting is that stress

and emotions actually cause changes in the way that we breathe. That’s right, it’s pretty

incredible. And what’s useful about that is that cause effortless breathing has been shown

by a number of studies to have a calming effect and reduce stress and anxiety. Bonus: it

also helps you focus and be productive.

“So the best way to do this is through effortless diaphragmatic breathing that makes

your entire chest and abdomen move subtlety together in an effortless manner – it’s

breathing in a natural way. You’re not under any pressure, and that’s really how you want

to be breathing. So you don’t have to do yoga to relieve stress. You can just breathe

calmly while you work, while you walk, while you play. So don’t forget, effortless

diaphragmatic breathing is the key. Good luck.”

Page 197: Augmented Self-Regulation (Moraveji, 2012)

179

C.3 Textual Explanation of BreathTray

This text was shown to the user on a web page before they began work on the two tasks

in the BreathTray condition:

“During these next tasks, you can see your current breath rate and calm points in the top-

right of the screen.

The first number refers to how many 'calm points' you have received. You receive points

when you breathe at or below your personal resting breath rate.

You will get more points the more calmly you breathe.

The second number refers to current breath rate.

For example: '21.8bpm 143%' means you are currently breathing at 21.8bpm, 43% above

your resting rate.

The text is RED when you are breathing above your resting rate and BLUE when you are

at or below.”

C.4 Text explanation of Serial Sevens task

Counting Backwards : Practice

We will ask you to count backwards by 7's. Just type the number and press [Enter].

The number which you input will clear, and enter the next number.

For example:

428 (we give you)

421 (you write - correct)

Page 198: Augmented Self-Regulation (Moraveji, 2012)

180

414 (you write - correct)

408 (you wrote - incorrect!)

401 (you write - correct)

394 (you write - correct)

Go for speed and accuracy.

Do not press Back, Reload, Forward, or any other keys.

Ready for a 90 seconds practice round?

Please press start button when you are ready.

C.5 Text explanation of Problem-Solving with Audio Distractors task

Problem-solving while hearing sounds : Practice

You'll do your best to solve simple math problems.

You need headphones because tones will play while you work. Please wear the

headphones now.

Continue working to solve the math problems. Go for speed and accuracy.

Some of the problems are difficult, don't worry and just try your best. :-)

Do not press Back, Reload, Forward, or any other keys.

Ready for a 90 seconds practice round?

Please press start button when you are ready.

C.6 Post-survey

This survey was completed on a webpage.

Page 199: Augmented Self-Regulation (Moraveji, 2012)

181

How distracting was the breath feedback? (1=not at all, 7=very much so) How much did you think breathing differently helped your performance? (1=not at all, 7=dramatic difference) How much did you think the breath feedback influenced your breathing? (1=not at all, 7=dramatic difference) How much did you think the breath feedback influenced your performance? (1=not at all, 7=dramatic difference) How likely would you be to use such computer-based feedback while you work on a normal day? (1=no chance, 7=definitely would use) How much experience do you have with conscious breathing and/or meditation? (1=none, 7=daily practice)

Page 200: Augmented Self-Regulation (Moraveji, 2012)

182

Appendix D Breathwear Instructions and Feedback Form

The URL for this form was given to participants to send feedback and

troubleshoot any issues that would arise when using the system. The live form is

available at http://bit.ly/bwstudy.

Breathwear FAQ & feedback

INSTALL APP

1. give us your iphone's udid by installing http://itunes.apple.com/us/app/udid-

sender/id306603975?mt=8 and emailing it to us

2. we will email you with some file attachments

3. save the .ipa & .mobileprovision files to your desktop

4. drag .mobileprovision file to your computer's itunes library: http://bit.ly/wEMbRf

5. drag the .ipa file to your computer's itunes library's apps folder (same way)

6. sync your iphone to itunes

Page 201: Augmented Self-Regulation (Moraveji, 2012)

183

UPDATE APP (when we send it to you)

1. save the .ipa file to your desktop

2. drag it to the “apps” section of your computer's itunes library: http://bit.ly/wEMbRf

3. drag it from your itunes library to your iphone's app library.

WEAR THE SENSOR

1. adjust the size first, make it very loose

2. put it on, then tighten the belt

3. it should not be tight like a belt - it should be loose

4. if any connections come off you can slide them back on

CONNECT SENSOR TO IPHONE

1. make sure app is not running (i.e., kill the process).

2. start the breathwear “bc” app.

3. press the belt's left red button. you should hear a beep. if it's a long beep you toggled it

off.

4. on the app, press "connect to sensor".

5. belt should do 3 short beeps when it is connects. app takes 5 sec to load.

6. "connect" button switches to "disconnect". data shows after 30sec. (7).

troubleshoot: in iphone's settings go to “general” then “bluetooth”. toggle bluetooth

off/on.

FIRST 24 HOURS

Page 202: Augmented Self-Regulation (Moraveji, 2012)

184

1. in the settings, set “show notifications” and “show data on home screen” to OFF.

2. after 24 hours, set them to ON.

EACH NIGHT: SEND DATA, CLEAR, DISCONNECT

1. press the SEND DATA button and send the email to [email protected].

2. press CLEAR DATABASE

3. press “disconnect” in the app. no need to press any sensor button.

ANY TIME

Note: change your baseline to your desired breath rate any time you want.

Having a sensor connection issue? See directions above.

• Not sure if my belt is on.

• My belt was working but it stopped.

• I can't get my belt working at all.

• Other:

Having trouble installing the app? See directions above.

• I can't seem to do the first install.

• I can't seem to update the app.

• Other:

Page 203: Augmented Self-Regulation (Moraveji, 2012)

185

Having trouble with the iPhone app? See directions above.

• Want to turn off notifications.

• Uses too much battery.

• My breath rate is not correctly detected.

• My baseline needs to be changed.

• Other:

I have a question, comment, idea, reflection, critique, or feedback not addressed

above.

Breathwear helps me increase my self-awareness.

1 2 3 4 5 6 7 8 9 10

Not at all Dramatically

Email address *

Mobile phone number *