Exploring Biofeedback with a Tangible Interface Designed for Relaxation
Morgane Hamon, Rémy Ramadour and Jérémy Frey
Ullo, 40 rue Chef de Baie, La Rochelle, France
Keywords:
Biofeedback, Tangible Interface, Relaxation, Ambient Display, Interoception.
Abstract:
Anxiety is a common health issue that can occur throughout one’s existence. In this pilot study we explore an
alternative technique to regulate it: biofeedback. The long-term objective is to offer an ecological device that
could help people cope with anxiety, by exposing their inner state in a comprehensive manner. We propose
a first iteration of this device, “Inner Flower”, that uses heart rate to adapt a breathing guide to the user; and
we investigate its efficiency and usability. Traditionally, such device requires user’s full attention. We propose
an ambient modality during which the device operates in the peripheral vision. Beside comparing “Ambient”
and “Focus” conditions, we also compare the biofeedback with a sham feedback (fixed breathing guide). We
found that the Focus group demonstrated higher relaxation and performance on a cognitive task (N-back).
However, there was no noticeable effect of the Ambient feedback, and the biofeedback condition did not yield
any significant difference when compared to the sham feedback. These results, while promising, highlight
the pitfalls of any research related to biofeedback, where it is difficult to fully comprehend the underlying
mechanisms of such technique.
1 INTRODUCTION
At times, to endure fear might be beneficial, when
one is facing dangerous situations or unknown events.
“Fight or flight” responses were proven important for
survival. However if stress becomes chronic and is
not treated, it can be a factor of sleep disorders or car-
diovascular disease (Kivimäki et al., 2002). It can also
lead to a pathological state of anxiety when the anti-
cipation of stressing stimuli is sufficient to trigger the
same symptoms as with the actual appearance of sti-
muli. Finding effective and lasting solutions to reduce
stress is necessary to alleviate this public health pro-
blem, which impedes the lives of many. Treatments
exist against stress and anxiety, but they might re-
quire a strong and timely involvement (i.e. therapies)
or provoke side effects (i.e. drugs). Studying anx-
iety and offering alternative solutions to drugs (sport,
yoga, mindfulness) is a growing body of research, and
nowadays tools exist to let people autonomously re-
flect on their states and better act upon themselves.
1.1 Biofeedback
Biofeedback is a method that enables users to learn to
control autonomous bodily processes. Biofeedback is
part of a larger notion known as "interoception" which
Figure 1: In this example heart-rate is measured through
a smartwatch; tangible interfaces serve both as breathing
guides for the user and as biofeedback devices (image
c
Inria, photograph C. Morel).
is defined by (Farb et al., 2015) as "the sense of sig-
nals originating within the body" or "the process of
receiving, accessing and appraising internal bodily
signals". Biofeedback relies on physiological measu-
rements and is getting more and more popular thanks
to the increasing availability of non-invasive sensors
(Figure 1). Most of the time, signals originate from
respiration, electrodermal activity (EDA) or heart rate
54
Hamon, M., Ramadour, R. and Frey, J.
Exploring Biofeedback with a Tangible Interface Designed for Relaxation.
DOI: 10.5220/0006961200540063
In Proceedings of the 5th International Conference on Physiological Computing Systems (PhyCS 2018), pages 54-63
ISBN: 978-989-758-329-2
Copyright © 2018 by SCITEPRESS Science and Technology Publications, Lda. All rights reserved
(HR) (McKee, 2008).
While these processes are mostly involuntary and
with weak level of consciousness, through physiolo-
gical sensors it becomes possible to visualize them in
real-time. Users are then able to view effects of a par-
ticular behavior on their physiology . For example if
a person wearing a HR sensor begins to run, HR is
going to increase suddenly. Connected to a biofeed-
back display the increase of HR could be symbolized
by a color change, a blinking light, or simply seen as
a graph. This information can then be used to help
users regulate their state, for example stay within a
specific HR range in to prevent injuries or maximize
the outcomes of the training. Biofeedback applicati-
ons are diverse, from rehabilitation to stress manage-
ment (McKee, 2008).
It is worth noting that despite the fact that bio-
feedback has been investigated and applied for deca-
des, its effectiveness has not been systematically de-
monstrated. Notably, even though a review such as
(Yucha and Montgomery, 2008) presents many stu-
dies across a variety of medical applications, there
are hardly any comparisons between biofeedback and
sham feedback, which is the only way to form a pro-
per control group and pin-point the real efficacy of the
technique.
Still, as a drug-free and a non-invasive approach,
potential risks and side effects of Biofeedback are
small. With a medical opinion it could be a compro-
mise for people who can not or do not want to take
drugs (e.g pregnant women).
1.2 Heart Rate Variability: A Marker
of Anxiety
One of the most studied case of biofeedback for stress
management is the cardiac activity. Heart Rate is un-
der control of the autonomous nervous system (ANS)
and is not constant at rest. Heart Rate Variability
(HRV) a marker of the evolution of HR over time
is a representative index of ANS activity. Among
the variety of factors that could influence HR, a low
HRV has been shown to be correlated with impaired
parasympathic activity, higher anxiety, and a variety
of disorders (Vaschillo et al., 2006). (Tan et al., 2011)
showed that veterans with Post Traumatic Stress Dis-
order (PTSD) had a HRV significantly lower than sub-
jects without PTSD. The HRV was compared between
veterans with PTSD who received HRV biofeedback
in addition to their regular treatment and veterans wit-
hout the additional biofeedback treatment. The re-
sults indicated that the group with HRV biofeedback
had significantly increased their HRV while reducing
symptoms of PTSD compared to the other group. Fin-
ding ways to assist people to learn to increase their
own HRV and improve their health is one of the mo-
tivations of our study.
1.3 Increasing Heart Rate Variablity
Deep breathing is a well documented manner to incre-
ase the HRV. “Cardiac Coherence” is a notion accor-
ding to which respiration and cardiovascular functi-
ons are synchronized. That means the HR increases
during the inhalation and decreases during exhalation,
and so periodically (McCraty et al., 2009). Authors
proposed that 6 breaths per minute is the respiratory
frequency that allows one to reach cardiac coherence
and the highest amplitude in heart rate oscillations
(hence the highest HRV) (DeBoer et al., 1987). Car-
diac Coherence is then a priori obtained when brea-
thing at a 0.1Hz frequency which corresponds to six
10-second breathes per minute.
Several studies have investigated a static breathing
guidance at 6 breaths per minute to reduce stress.
For example in (Yu et al., 2015), where authors sho-
wed that HRV was higher after the breathing exer-
cise, while subjective impressions as measured by
the State Trait Anxiety Index questionnaire, or STAI
(Spielberger et al., 1970) were not different. In
(Dijk and Weffers, 2011) authors designed an im-
mersive system which is composed of a blanket con-
taining small vibrating motors (haptic stimulus) and
headphones (audio stimulus). Haptic and audio sti-
muli are synchronized and generated at such a fre-
quency that the user can follow as a breathing gui-
dance. One of the studied frequency modality was
6 breaths/min. Some participants reported the fre-
quency as too fast or too slow. The Authors highlig-
hted that a poorly adapted frequency can potentially
create hypo/hyperventilation.
Even if a static guidance might, on average be op-
timal for the population, it is not the best way to max-
imize HRV for each individual. As a matter of fact,
(Vaschillo et al., 2006) found that resonant frequency
(equivalent to the cardiac coherence) differs accor-
ding to each person. By exposing a breathing gui-
dance from 4.5 to 6.5 breaths per minute to their par-
ticipants, they managed to define the best frequency
for each subject.
As such, we opted for an adapted breathing gui-
dance to create our biofeedback; in our current study
we compare it with a fixed breathing feedback (i.e.
“Dynamic” vs “Static” feedback) in order to have a
better grasp on the effect of an adapted biofeedback.
Exploring Biofeedback with a Tangible Interface Designed for Relaxation
55
1.4 Shaping the Best Biofeedback
(Muench, 2008) proposed a certain type of biofeed-
back to achieve higher HRV and reach a resonant fre-
quency. In this study, a graph based on HR fluctua-
tions was created, having users inhaling until the HR
reaches a “peak” and exhaling until it reaches a “pit”.
Because of the improved HRV as compared to bre-
athing exercises agnostic to HR, we also designed a
biofeedback device that would use HR measurements
to propose an adapted breathing guide.
However, the feedback modality used to convey
the information back to the user is important, and
graphs are not the only way to present a breathing
guidance. Actually, such kind of feedback might even
impede acceptability because it could appear as being
too judgmental due to its close relationship to a me-
tric. In order to craft a more “organic” and yet infor-
mative biofeedback, we decided instead to rely on a
physical object to present the feedback.
Indeed, thanks to recent advances in human-
computer interaction, is is now possible to directly
integrate digital information within users’ surroun-
dings, for example through the use of tangible inter-
faces. These interfaces have been proven effective to
help people learn about bodily processes see e.g.
(Fleck et al., 2018). Through tangible interfaces, bi-
ofeedback can then be more easily integrated in the
natural settings of users, and become part of a speci-
fic scenario.
1.5 Ambient Feedback
Ambient computing contrasts with disrupting notifi-
cations. In this context “ambient” refers to informa-
tion that is being presented in the peripheral attention
of users (MacLean, 2009). Ambient devices do not
mobilize attention. They require minimal efforts from
the user and yet they provide informative feedback;
rather than “pushing” a notification it is up to users to
“pull” information when they require it.
Along those lines, (Moraveji et al., 2011) high-
lights two ways to train respiration: consciously thin-
king about it or learning to follow an external stimulus
such as a pacing light or an auditory guide. While the
former method implies focus of attention, the latter, if
proven effective, could alleviate the required amount
of cognitive resources (Hazlewood et al., 2011).
In (Moraveji et al., 2011), authors investigated
whether a breathing guidance at the periphery of the
screen during work had a influence on breath rate. As
this type of guide does not require full attention they
called it Peripheral Paced Respiration (PPR). This
ambient guide can be very useful by allowing the user
to fully commit to another task. The guide rate setting
was 20% below the user baseline. They showed signi-
ficant difference on breath rate depending on the acti-
vation of the PPR. The same thinking led (Azevedo
et al., 2017) to develop a wearable device that deli-
vers tiny vibrations on the wrist with a frequency 20%
slower than the participant resting HR. Users didn’t
know the function of the device and were preparing an
oral presentation (to induce stress). Results showed
that the control group had significantly higher anxiety
according to questionnaires (STAI) and physiological
data (EDA).
(Schein et al., 2001) investigated in a longitudinal
study if a device called BIM (for Breathe with Inte-
ractive Music) had a positive influence on Blood Pres-
sure (BP). The BIM device creates a musical pattern
which is related to the user breathing rate. A 10 minu-
tes long quiet synthesized music recording was used
as an active control. Results showed BP reduction
was greater in the experimental group compared to
the control, and seem to have a long-term effect (sig-
nificantly different 6 months after).
Based on these various findings, we decided to in-
vestigate not only a “Focus” but an Ambient” utiliza-
tion of a tangible biofeedback device as well. In order
to better frame our experimental design, we turned
to previous work from psychology and physiological
computing that aimed at investigating various dimen-
sions of stress.
1.6 Evaluating and Inducing Stress
Inducing stress to evaluate short-term effects of bio-
feedback devices is common in the literature. To arti-
ficially put participants in a more stressed state incre-
ases the range of measurements and help to reduce va-
riability between subjects. Different type of stress can
be induced, physical (e.g. extreme heat or cold), psy-
chological (e.g. increase in mental workload) or psy-
chosocial (e.g. public speaking) (Mühl et al., 2014).
Psychosocial stress can be induced by faking an in-
terview – Trier Social Stress Test (Kirschbaum et al.,
1993) or simply asking participants to prepare a
speech, as in (Azevedo et al., 2017).
In (Yu et al., 2015) authors induced psychologi-
cal stress with a mathematical task for a ten-minutes
period. Several manners exists to check if the stress
inducing task had an effect. First, it can be detected by
recording physiological data, for example with HRV
as a marker of anxiety. EDA is also a common re-
corded measure e.g. (Roseway et al., 2015) to
detect arousal. Because physiological signals might
have poor specificity, those indicators do not repre-
sent how participants are feeling, what is their state
PhyCS 2018 - 5th International Conference on Physiological Computing Systems
56
of mind. Questionnaires can compensate for this si-
tuation. To measure anxiety the most frequently used
questionnaire is the STAI as in (Yu et al., 2015), (Aze-
vedo et al., 2017), or (Mühl et al., 2014). In the latter
work authors compared the robustness of various phy-
siological signals to detect stress. They crossed psy-
chosocial stress (TSS) with psychological stress, by
manipulating the amount of cognitive workload en-
dured by participants. To do so, they employed the
N-back task, which leverage on memory load (Owen
et al., 2005). We chose to use the N-back task as well
to induce stress in our study since it is effective in
doing so and since we can easily compute a perfor-
mance metric to sense how participants might be af-
fected by the exposition to a tangible biofeedback.
1.7 Objectives
In the present study, our first objective is to investigate
if an ambient modality could have positive effects on
stress. To explore this issue we compared two moda-
lities: Ambient vs Focus. Ambient” stands for the
presence of the device during the whole experiment
(30min) in the peripheral visual field of the partici-
pant. “Focus” refers to a short (6 min) utilization in
the middle of the experiment requiring full-attention.
Our second objective is to check the relevance
of breathing guidance based on HR . Indeed, as we
propose a new device it is important to consider that
any observable effect might be due to a “novelty ef-
fect”. Moreover, we wanted to assess the usefulness
of actually measuring physiological activity. Hence
we compare a true biofeedback (“Dynamic” condi-
tion, where the breathing guide is adapted using HR)
with a sham biofeedback, or “pseudofeedback” (“Sta-
tic” condition, where the breathing guide is set to a
fixed rate). As explained by (Health et al., 1982), a
pseudofeedback is defined as a non contingent stimu-
lus presented in exactly the same manner as the true
biofeedback and with the intention of having subjects
believe that it is true biofeedback”.
2 STUDY
This study possesses a 2 (Attention: Ambient vs
Focus) × 2 (Biofeedback: Dynamic vs Static)
between-subject experimental plan. As a result
we split our participants into four distinct groups:
Ambient-Dynamic, Ambient-Static, Focus-Dynamic
and Focus-Static.
Because in the Focus group the device is used only
mid-experiment, the first part of the experiment in Fo-
cus serves as a control group (no device) for the inves-
tigation of attention.
Our hypotheses are:
H1. Exposition to an Ambient feedback (H1a)
or to a Focus feedback (H1b) reduces psychological
stress.
H2. An adapted breathing guidance with a bio-
feedback device (Dynamic) will reduce stress as com-
pared to a pseudofeeback (Static).
H3. The usability resulting from the use of a
biofeedback device is improved as compared to a
pseudofeedback (Focus-Dynamic vs Focus-Static).
2.1 Inner Flower
Figure 2: Left: Examples of prototypes toward an tangible
biofeedback. Right: Design of the Inner Flower used in the
study.
In order to test our hypothesis, we iterated over
various form factors to craft a tangible biofeedback
that could both act as an ambient feedback and as a
breathing guide. Among the biofeedback modalities
that are traditionally employed – visual, audio or hap-
tics (Frey et al., 2018) suggests that there is little
difference regarding the effectiveness of conveying a
breathing feedback. We chose to rely on visuals with
color LEDs since they offered many degrees of free-
dom to work with (e.g. location of the lights, inten-
sity, color, speed of the animation) and since it was
easy to implement.
We employed laser cutting, Arduino-based micro-
contollers (Arduino mini), and Adafruit
1
Neopixel
LEDs to prototype several artifacts (Figure 2). It
quickly became apparent that frosted plastic was the
best material to work with in order to improve the
LEDs’ diffusion, smooth the light and prevent an im-
pression of a pixelated display, which would be too
much reminiscence of a desktop display. Regarding
1
https://www.adafruit.com/
Exploring Biofeedback with a Tangible Interface Designed for Relaxation
57
the actual appearance of the object, after having ex-
plored more abstract shapes, we decided to create a
somewhat stereotypical flower in order to convey a
sense of well-being. Additionally, as one might want
to smell the scent of a flower, it is in a way a valid
proxy for breathing.
Then we had to design the feedback that would
guide the breathing. While it was straightforward to
pick movements as a guidance, several adjustments
were required in order to smooth the animation of the
LEDs that were disposed on the “petals” of the flower.
In the final version (Figure 2, right), the user would
breathe in when the lights from of the Inner Flower
move outwards and breathe out when the lights move
inward. In the Static condition this animation is set
to 6 breaths/min. In the Dynamic condition, the bre-
athing rate (BR) is coupled with the heart rate of the
participant. We start our investigation with a simple
coupling: BR = HR/. The divider has a default
value of 15. It can be set to adapt the breathing gui-
dance to each user during a calibration phase. During
the Dynamic condition, the breathing guidance varies
in real-time according to the HR of the user so as to
reach cardiac coherence. For example with a HR at 90
beats per minute and a divider set to 12, the resulting
breathing guide pulses 7.5 times per minute.
There is no direct representation of the HR (e.g.
no light blinking at the pace of the heart rate) so as
not to overwhelm users with information not related
to current usage scenario.
2.2 Measures
During the study we were interested in collecting two
types of data: the (psychological) stress level of par-
ticipants, and a usability index related to the tangible
biofeedback. Stress was assessed along three dimen-
sions: physiological activity (HR), behavioral measu-
res (performance in a N-back task) and questionnai-
res (STAI). Usability was assessed through a questi-
onnaire (USE) administrated at the end of the experi-
ment for those groups that explicitly used the device
(i.e. Focus groups).
2.2.1 Heart Rate
Each participant was equipped with a smartwatch me-
asuring heart rate (“Link” from Mio
2
), placed on their
non-dominant hand. Mio smartwatchs employ pho-
toplethysmography (PPG) to compute heart-rate, a
technique which basically detects variations in skin’s
color to assess heartbeats. This solution was prefer-
red over electrocardiography (ECG) to improve com-
2
https://www.mioglobal.com/
fort and acceptance of the system, which is meant to
be used in ecological settings, outside the laboratory.
Even though PPG is less robust than ECG, being that
the participants are steady and seated, i.e. without the
risk of creating motion artifacts throughout this study,
PPG is a sufficiently good sensor. During pilot stu-
dies, we validated that the instantaneous HR measu-
red by this particular smartwatch was accurate enough
to detect changes in HRV associated with deep relax-
ation.
Data was collected over Bluetooth, processed in
real-time in the Dynamic condition and stored for
further analysis. From HR measurements we focu-
sed our investigation on one index of HRV: RMSSD
– root mean sum of the squared differences. RMSSD
takes as input the inter-beat interval (the inverse of
the instantaneous HR measured by the smartwatch).
It is one of the best indicator of cognitive workload
(Mehler et al., 2011), a specific type of psychological
stress induced during the experiment.
2.2.2 N-back
Figure 3: 2-back task. Participants have to click left when
the displayed letter is the same as two steps before.
The N-back task served both to induce psycholo-
gical (cognitive) stress, and to evaluate the cognitive
load of participants. The latter is revealed by calcu-
lating participants’ performances. During this task,
as in (Mühl et al., 2014), each letter appears on the
screen for 0.5 second every 2 seconds. Participants
have to determine whether the current letter is the
same as the one they saw N steps before (left click
on a mouse) or not (right click). We employed a 2-
back task (Figure 3), which showed to induce a high
workload level (Mühl et al., 2014). Typically, when
workload increases the performance during the task
decreases.
An immediate feedback is provided to inform par-
ticipants whether their answers are correct or not.
Two N-back tasks were presented to each participant,
denoted as “N-back1” and “N-back2”. Each task is
PhyCS 2018 - 5th International Conference on Physiological Computing Systems
58
comprised out of three sequences of one minute (30
letters). The first task (N-back1) includes an additi-
onal sequence at the beginning that acts as a training
session. Over each N-back task we measured partici-
pants’ performance (percentage of correct answers).
2.2.3 STAI
The State Trait Anxiety Index (STAI) enables the me-
asurement of self-reported anxiety levels (Spielberger
et al., 1970). The STAI Y-A version of the question-
naire measures anxiety as a state (about one’s current
endeavor). The classical version is composed of 20
items (e.g. “I am tense”; “I feel content”) that parti-
cipants rate on a 4-point Likert scale. A higher score
reflects a higher anxiety state. Because we adminis-
trated multiple times the STAI during the experiment,
we favored a six-item short-form (Marteau and Bek-
ker, 1992). It was faster to fill by participants and
still produced similar scores to those obtained with
the full-form.
2.2.4 Usability
In order to assess the usability of the device, we adap-
ted the USE questionnaire (Lund, 2001). We remo-
ved the “Usefulness” subscale because it was not rele-
vant to our study – participants did not use the device
long enough. Questions were translated to French.
The resulting questionnaire comprises three subsca-
les: “Ease of use”, “Ease of learning” and “Satis-
faction”. Each is composed of three items (e.g. “I
easily remember how to use it”, “It is simple to use”).
Participants had to state their agreement to each sen-
tence on a 4-point Likert scale (“Not at all” ... “Very
much”). Since such usability questionnaire is poorly
suited to assess an ambient device, with which users
merely interact directly and/or consciously, the USE
was only administrated to participants of the Focus
groups.
2.3 Participants
A total of 36 volunteers (18 females) took part in the
study. Overall the mean age was 23.80 years old (SD
= 4.82). The demographics per group is depicted in
Figure 4.
2.4 Procedure
The timeline of the experiment is presented Figure 5.
The experiment took place in a quiet room, deprived
of distractions, and we detail it step by step as follows.
Figure 4: Demographics of our groups.
Figure 5: Timeline of the experiment.
Upon entering the room and sitting at a table, par-
ticipants signed a consent form. Afterwards we pro-
ceeded to explain briefly how the experiment would
take place and equipped participants with the smart-
watch. The Inner Flower was then switched on and its
core functionalities were explained to participants. In
the Focus-Dynamic group, a calibration phase occur-
red in order to determine a pace that would suit users
(i.e. obtaining a breathing guide that would not appear
too fast nor too slow). After the calibration, the device
was switched off. On the other hand, the Focus-Static
group had no calibration, hence the device was sim-
ply switched off. Lastly in the Ambient groups the
device was left active on the side of the table, in the
peripheral vision of participants.
Afterwards, to induce psychological stress and
control for the effect of the N-back task participants
fulfilled a first STAI (STAI-1), performed a first N-
back task (N-back1) and finally fulfilled a second
STAI (STAI-2). At this point of the experiment, by
comparing Ambient groups (active device on the side)
with Focus groups (devices turned off) we are able to
investigate the effect of an ambient breathing guide.
Exploring Biofeedback with a Tangible Interface Designed for Relaxation
59
In order to assess how much the Inner Flower
would affect HRV while employed as an explicit brea-
thing guide, further in the experiment we switched-on
the device in the Focus groups and participants car-
ried out the breathing exercise. In the Ambient groups
participants performed a substitute task instead; they
read the short story “The Oval Portrait” by Edgar Al-
lan Poe. Both tasks lasted 6 minutes and were de-
signed to equally involve users’ attention. At the end
of the task, the device was switched-off in the Focus
groups.
Then, all participants performed a second N-back
task (N-back2). Their performance, during this test
would enable us to assess the efficacy of the Inner
Flower as a tool to reduce psychological stress and
improve cognitive availability.
Finally, participants filled out a third STAI (STAI-
3). Additionally, in the Focus groups, participants
answered the USE questionnaire.
3 RESULTS
Due to the nature of the data and the modest number
of participants per group, we used non-parametric sta-
tistical tests to analyze the data. For studying HRV,
N-back and STAI, we performed resampling (per-
mutation) statistics using the Minque
3
package from
R. Answers to the USE questionnaire were analyzed
with a Wilcoxon rank sum test. When applicable, we
tested for the effect of each main factor (Attention,
Biofeedback and moment of measurement) as well as
for the interaction of thereof.
3.1 HRV
HRV was computed during both N-back tasks (time1
and time3) and during the breathing exercise (time2).
There was a significant interaction between the At-
tention factor and time, with a difference between
Focus during the breathing task (M=1.47×10
2
,
SD=0.67×10
2
) and the rest of the conditions
(M=1.07×10
2
, SD=0.48×10
2
, p < 0.05, Figure 6).
Across other factors and interactions there were no
significant differences in HRV.
Note that due to technical issues the HR data is
incomplete for 5 participants (out of 36).
3.2 N-back
We investigated the evolution of the percentage of
accuracy during the N-back between the first and
3
https://cran.r-project.org/web/packages/minque/
minque.pdf
Figure 6: HRV across Attention factor and time (i.e. during
N-back1, exercise, N-back2). “*”: p-value < 0.05.
Figure 7: Evolution of the performance between the first
and the second N-back task. Performance increased further
in the Focus groups. “**”: p-value < 0.01.
the second test (i.e. N-back2 - N-back1). Over-
all performance increased between those two tests;
with a significant effect of the Attention factor. Per-
formance increased more sharply in Focus groups
(M=+11.23pp, SD= 8.36) as compared to Ambient
groups (M=+4.21pp, SD= 7.03, p<0.01, Figure 7).
Across other factors and interactions there was no sig-
nificant differences in the evolution of N-back perfor-
mance.
3.3 STAI
There was a significant effect of time. STAI-2 sco-
res (after the first N-back task) were higher (M=42.4,
SD=9.08, p < 0.01) when compared to the other sco-
res (M=34.8, SD=8.57, Figure 8, top). There was
a significant interaction between Attention and time,
with lower scores in the Focus groups in STAI-3 at
the end of the experiment (M=30.8, SD=5.92), when
compared to the rest of the conditions. (M, 38.5,
SD=9.48, p < 0.01, Figure 8, bottom). Across other
PhyCS 2018 - 5th International Conference on Physiological Computing Systems
60
Figure 8: Top: STAI scores across time (lower score: less
anxiety). Bottom: STAI scores across Attention factor and
time. “**”: p-value < 0.01.
factors and interactions there was no significant diffe-
rences.
3.4 USE Questionnaire
As mentioned earlier the USE questionnaires were
fulfilled only in Focus groups. There was no signi-
ficant difference between Static (M=32.8, SD= 2.92)
and Dynamic (M=28.9, SD=6.79, p=0.26) groups.
4 DISCUSSION
The N-back task had a significant effect on the percei-
ved anxiety, as the STAI-2 was higher than the STAI-
1. This replicates the validity of such task to induce
stress.
As opposed to previous work e.g. (Yu et al.,
2015), which employed arithmetic tasks between bre-
athing exercises – we did not measure a noticeable ef-
fect of the Ambient condition and could not validate
H1a. There was no significant difference within any
marker of stress (HRV, N-back performance, STAI)
between participants with an active ambient device
and participants with no device (beginning of the
study). We would need longer experiments, or even
longitudinal studies, to better grasp the influence and
the dynamic of an ambient (and subtle) biofeedback.
Interestingly, when asked after the experiment du-
ring informal interviews, participants in the Ambient
groups reported that they did not pay attention to the
device an indicator that at the very least it was not
disruptive. Still, it is possible that such an ambient bi-
ofeedback might reduce stress even more than a short
explicit breathing exercise when users are exposed for
a couple hours, or over the course of several days.
In contrast, over the 6-minute breathing exercise
participants of the Focus group demonstrated a lower
level of stress on all three markers – H1b is validated.
Not only did the breathing exercise increased HRV
and helped participants to feel less anxious, but it also
induced a higher performance during the N-back task.
This scenario mirrors the (negative) effects that might
arise when a psychological test is administrated to a
sensitive population in a stressful environment (e.g.
elderly in a hospital). In similar situations, the use of a
device alike the Inner Flower might increase patients’
cognitive availability before a test and prevent biases
during a psychological evaluation.
Since there was no effect of the Feedback moda-
lity over the course of the experiment, we cannot vali-
date H2. An actual biofeedback did not perform better
(or worse) than a simple breathing exercise based on a
fixed breathing rate. There are however multiple fac-
tors that could explain this outcome and that would
need further investigations.
First, by trying to give more freedom to users and
finely adapt the breathing guide (i.e. calibration phase
in Focus-Dynamic group) we might have introduced
a higher between-subject variability, as we observed
more variability in the Focus-Dynamic group.
Second, the N-back task might have resulted in a
ceiling effect with some participants. While overall
participants tended to improve their scores between
N-back1 and N-back2, those who already had a good
score (i.e. >80%, n=17) had difficulties to improve af-
terward. This plateau is among the confounding fac-
tors that we would need to control more finely du-
ring recruitment for future work, alongside persona-
lity traits.
Third, and maybe most importantly, our choice of
biofeedback might not have been perceived as an ac-
tual manifestation of user’s physiology. While we did
not want to overwhelm users with too many stimuli,
the absence of a dedicated HR biofeedback might
have impeded their sense of agency, i.e. users did not
realize that they were actually “connected” with the
Exploring Biofeedback with a Tangible Interface Designed for Relaxation
61
device.
This latter interpretation would be on par with the
fact that there was no significant difference in the usa-
bility questionnaire. With current results we cannot
validate H3, since Focus-Dynamic and Focus-Static
groups rated equally high the device, between 80%
and 90%. We would expect a higher sense of agency
(i.e. in Focus-Dynamic) to be reflected on “Ease of
use” and “Ease of learning” subscales of USE.
Despite the encouraging effect on relaxation and
cognitive availability of the Focus groups, when users
are presented with an explicit breathing guide, the ab-
sence of results between a coupled (Dynamic) and
a fixed (Static) breathing guidance highlights once
again how much rigor is needed when assessing the
effect of physiological measures and the resulting be-
nefit of biofeedback.
5 LIMITATIONS AND FUTURE
WORK
As shown by (Hazlewood et al., 2011) it is difficult to
evaluate ambient technology on a one-time basis be-
cause it has to be by definition blended into the envi-
ronment. In the present study, conditions were maybe
not ecological enough in the sense that experiments
took place in a small and impersonal room. To solve
these issues we are working with a designer to bet-
ter script the overall interaction and bring the study to
a home studio replica. Our next experimental design
will include as well a control group without any de-
vice, for a better control of the confounding variables.
While it is harder to bring the technology outside the
laboratory, longitudinal studies, over several days or
weeks, would inform us about the required amount of
time for an ambient biofeedback to become effective.
Concerning the form factor, we conducted infor-
mal interviews at the end of the experiment. Some of
the participants showed interest in multi-modal feed-
back, for example an audio feedback to let them conti-
nue the breathing exercise while closing their eyes. In
order to enable such usage we will complement ex-
isting feedback, for example with audio waves as in
(Dijk and Weffers, 2011).
To determine the importance of physiological me-
asurements, we hypothesize that a social usage of the
device might be a way to reinforce the association
with an actual biofeedback. (Roseway et al., 2015)
designed an ambient device embedded in the work-
place that informs those around about one’s state.
Thanks to that colleagues were more sensitive to emo-
tional states of others. Similar examples inspired us
to create a collaborative scenario where users would
try to synchronize their heart rate through the device,
with a color-based HR feedback. We envision such
exercise as a way to establish trust between people.
The Inner Flower could then become both a tool to
manage stress and a proxy for communication, for ex-
ample in situations involving care takers and care gi-
vers, or when people suffer from communicative dis-
orders.
Over the course of the study we demonstrated how
a tangible device could help reduce perceived anxiety
(measured with STAI) and alleviate a physiological
symptom of stress (increase in HRV). Moreover, it
enabled participants to improve performance in a cog-
nitive task. Increasing cognitive availability before a
psychological test is one of the most promising ap-
plications of this technology. With a usability score
between 80% and 90% we can state that the device
was appealing to users. In the future we will further
investigate the influence of both attention and type of
biofeedback.
REFERENCES
Azevedo, R. T., Bennett, N., Bilicki, A., Hooper, J., Mar-
kopoulou, F., and Tsakiris, M. (2017). The calming
effect of a new wearable device during the anticipa-
tion of public speech. Scientific Reports, 7(1):1–7.
DeBoer, R. W., Karemaker, J. M., and Strackee, J. (1987).
Hemodynamic fluctuations and baroreflex sensitivity
in humans: a beat-to-beat model. American Jour-
nal of Physiology-Heart and Circulatory Physiology,
253(3):H680–H689.
Dijk, E. O. and Weffers, A. (2011). Breathe with the ocean:
a system for relaxation using audio, haptic and visual
stimuli. In Special Symposium at EuroHaptics 2010:
Haptic and Audio-Visual Stimuli: Enhancing Expe-
riences and Interaction, Amsterdam, The Netherlands.
Univ. Twente.
Farb, N., Daubenmier, J., Price, C. J., Gard, T., Kerr, C.,
Dunn, B. D., Klein, A. C., Paulus, M. P., and Mehling,
W. E. (2015). Interoception, contemplative practice,
and health. Frontiers in Psychology, 6(June):763.
Fleck, S., Baraudon, C., Frey, J., Lainé, T., and Hachet, M.
(2018). “Teegi’s so Cute!”: Assessing the Pedagogi-
cal Potential of an Interactive Tangible Interface for
Schoolchildren. In IDC 2018 - 17th Interaction De-
sign and Children Conference, Trondheim, Norway.
Frey, J., Grabli, M., Slyper, R., and Cauchard, J. R.
(2018). Breeze: Sharing Biofeedback Through We-
arable Technologies. In CHI ’18.
Hazlewood, W. R., Stolterman, E., and Connelly, K. (2011).
Issues in evaluating ambient displays in the wild: two
case studies. In Proceedings of the SIGCHI confe-
rence on Human factors in computing systems, pages
877–886. ACM.
Health, T., The, S. A., and Hatch, J. P. (1982). Controlled
Group Designs in Biofeedback Research: Ask, "What
PhyCS 2018 - 5th International Conference on Physiological Computing Systems
62
Does the Control Group Control For?". Biofeedback
and Self-regulation, 7(3).
Kirschbaum, C., Pirke, K.-M., and Hellhammer, D. H.
(1993). The "Trier Social Stress Test" - A Tool for
Investigating Psychobiological Stress Responses in a
Laboratory Setting. Neuropsychobiology, 28(1-2):76–
81.
Kivimäki, M., Leino-Arjas, P., Luukkonen, R., Riihimäi,
H., Vahtera, J., and Kirjonen, J. (2002). Work
stress and risk of cardiovascular mortality: pro-
spective cohort study of industrial employees. BMJ,
325(7369):857.
Lund, A. M. (2001). Measuring usability with the use que-
stionnaire12. Usability interface, 8(2):3–6.
MacLean, K. E. (2009). Putting haptics into the ambience.
IEEE Transactions on Haptics, 2(3):123–135.
Marteau, T. M. and Bekker, H. (1992). The development
of a six-item short-form of the state scale of the Spiel-
berger State-Trait Anxiety Inventory (STAI). British
Journal of Clinical Psychology, 31(3):301–306.
McCraty, R., Atkinson, M., Tomasino, D., and Bradley,
R. T. (2009). The coherent heart: Heart-brain in-
teractions, psychophysiological coherence, and the
emergence of system-wide order. Integral Review,
5(2):10–115.
McKee, M. G. (2008). Biofeedback: An overview in the
context of heart-brain medicine. Cleveland Clinic
Journal of Medicine, 75(SUPPL.2):31–34.
Mehler, B., Reimer, B., and Wang, Y. (2011). A comparison
of heart rate and heart rate variability indices in dis-
tinguishing single task driving and driving under se-
condary cognitive workload. Proceedings of the Sixth
International Driving Symposium on Human Factors
in Driver Assessment, Training and Vehicle Design,
pages 590–597.
Mühl, C., Jeunet, C., and Lotte, F. (2014). Eeg-based wor-
kload estimation across affective contexts. Frontiers
in Neuroscience, 8:114.
Moraveji, N., Olson, B., Nguyen, T., Saadat, M., Khalighi,
Y., Pea, R., and Heer, J. (2011). Peripheral Paced
Respiration: Influencing User Physiology during In-
formation Work. the 24th annual ACM symposium
on User interface software and technolog, pages 423–
428.
Muench, F. (2008). HRV: The manufacturers and vendors
speak. Association for Applied Psychophysiology &
Biofeedback, 36(1):35–39.
Owen, A. M., McMillan, K. M., Laird, A. R., and Bullmore,
E. (2005). N-back working memory paradigm: A
meta-analysis of normative functional neuroimaging
studies. Human Brain Mapping, 25(1):46–59.
Roseway, A., Lutchyn, Y., Johns, P., Mynatt, E., and Czer-
winski, M. (2015). BioCrystal : An Ambient Tool for
Emotion and Communication. International Journal
of Mobile Human Computer Interaction, 7(3):20–41.
Schein, M., Gavish, B., Herz, M., Rosner-Kahana, D., Na-
veh, P., Knishkowy, B., Zlotnikov, E., Ben-Zvi, N.,
and Melmed, R. (2001). Treating hypertension with a
device that slows and regularises breathing: a rando-
mised, double-blind controlled study. Journal of Hu-
man Hypertension, 15(4):271–278.
Spielberger, C. D., Gorsuch, R. L., and Lushene, R. E.
(1970). Manual for the state-trait anxiety inventory.
Palo Alto, CA: consulting psychologists press.
Tan, G., Dao, T. K., Farmer, L., Sutherland, R. J., and Ge-
virtz, R. (2011). Heart rate variability (hrv) and post-
traumatic stress disorder (ptsd): A pilot study. Applied
Psychophysiology and Biofeedback, 36(1):27–35.
Vaschillo, E. G., Vaschillo, B., and Lehrer, P. M. (2006).
Characteristics of Resonance in Heart Rate Variability
Stimulated by Biofeedback. Applied Psychophysio-
logy and Biofeedback, 31(2).
Yu, B., Feijs, L., Funk, M., and Hu, J. (2015). Breathe
with touch: a tactile interface for breathing assistance
system. In Human-Computer Interaction, pages 45–
52. Springer.
Yucha, C. B. and Montgomery, D. (2008). Evidence-based
practice in biofeedback and neurofeedback. AAPB
Wheat Ridge, CO.
Exploring Biofeedback with a Tangible Interface Designed for Relaxation
63