scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Muscle Synergies Facilitate Computational Prediction of Subject-Specific Walking Motions.

TL;DR: A patient-specific synergy-controlled neuromusculoskeletal simulation framework that can predict walking motions for an individual post-stroke and suggest that the framework may be able to bridge the gap between patient- specific muscle synergy information and resulting functional capabilities and limitations.
Abstract: Researchers have explored a variety of neurorehabilitation approaches to restore normal walking function following a stroke. However, there is currently no objective means for prescribing and implementing treatments that are likely to maximize recovery of walking function for any particular patient. As a first step toward optimizing neurorehabilitation effectiveness, this study develops and evaluates a patient-specific synergy-controlled neuromusculoskeletal simulation framework that can predict walking motions for an individual post-stroke. The main question we addressed was whether driving a subject-specific neuromusculoskeletal model with muscle synergy controls (5 per leg) facilitates generation of accurate walking predictions compared to a model driven by muscle activation controls (35 per leg) or joint torque controls (5 per leg). To explore this question, we developed a subject-specific neuromusculoskeletal model of a single high-functioning hemiparetic subject using instrumented treadmill walking data collected at the subject’s self-selected speed of 0.5 m/s. The model included subject-specific representations of lower body kinematic structure, foot-ground contact behavior, electromyography-driven muscle force generation, and neural control limitations and remaining capabilities. Using direct collocation optimal control and the subject-specific model, we evaluated the ability of the three control approaches to predict the subject’s walking kinematics and kinetics at two speeds (0.5 and 0.8 m/s) for which experimental data were available from the subject. We also evaluated whether synergy controls could predict a physically realistic gait period at one speed (1.1 m/s) for which no experimental data were available. All three control approaches predicted the subject’s walking kinematics and kinetics (including ground reaction forces) well for the model calibration speed of 0.5 m/s. However, only activation and synergy controls could predict the subject’s walking kinematics and kinetics well for the faster non-calibration speed of 0.8 m/s, with synergy controls predicting the new gait period the most accurately. When used to predict how the subject would walk at 1.1 m/s, synergy controls predicted a gait period close to that estimated from the linear relationship between gait speed and stride length. These findings suggest that our neuromusculoskeletal simulation framework may be able to bridge the gap between patient-specific muscle synergy information and resulting functional capabilities and limitations.

Content maybe subject to copyright    Report

UC Davis
UC Davis Previously Published Works
Title
Muscle Synergies Facilitate Computational Prediction of Subject-Specific Walking Motions.
Permalink
https://escholarship.org/uc/item/84h3r42d
Journal
Frontiers in bioengineering and biotechnology, 4(OCT)
ISSN
2296-4185
Authors
Meyer, Andrew J
Eskinazi, Ilan
Jackson, Jennifer N
et al.
Publication Date
2016
DOI
10.3389/fbioe.2016.00077
Peer reviewed
eScholarship.org Powered by the California Digital Library
University of California

October 2016 | Volume 4 | Article 771
METHODS
published: 13 October 2016
doi: 10.3389/fbioe.2016.00077
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org
Edited by:
Ramana Vinjamuri,
Stevens Institute of
Technology, USA
Reviewed by:
Laurent Simon,
New Jersey Institute of
Technology, USA
Jonathan B. Shute,
University of Florida, USA
*Correspondence:
Benjamin J. Fregly
fregly@u.edu
Specialty section:
This article was submitted
to Bionics and Biomimetics,
a section of the journal
Frontiers in Bioengineering
and Biotechnology
Received: 01June2016
Accepted: 21September2016
Published: 13October2016
Citation:
MeyerAJ, EskinaziI, JacksonJN,
RaoAV, PattenC and FreglyBJ
(2016) Muscle Synergies Facilitate
Computational Prediction of
Subject-Specic Walking Motions.
Front. Bioeng. Biotechnol. 4:77.
doi: 10.3389/fbioe.2016.00077
Muscle Synergies Facilitate
Computational Prediction of
Subject-Specic Walking Motions
Andrew J. Meyer
1
, Ilan Eskinazi
1
, Jennifer N. Jackson
2
, Anil V. Rao
1
, Carolynn Patten
3,4
and
Benjamin J. Fregly
1,2
*
1
Department of Mechanical and Aerospace Engineering, University of Florida, Gainesville, FL, USA,
2
Department of
Biomedical Engineering, University of Florida, Gainesville, FL, USA,
3
Department of Physical Therapy, University of Florida,
Gainesville, FL, USA,
4
Neural Control of Movement Lab, Malcom-Randall VA Medical Center, Gainesville, FL, USA
Researchers have explored a variety of neurorehabilitation approaches to restore normal
walking function following a stroke. However, there is currently no objective means for
prescribing and implementing treatments that are likely to maximize recovery of walking
function for any particular patient. As a rst step toward optimizing neurorehabilitation
effectiveness, this study develops and evaluates a patient-specic synergy-controlled
neuro musculoskeletal simulation framework that can predict walking motions for
an individual post-stroke. The main question we addressed was whether driving a
subject-specic neuromusculoskeletal model with muscle synergy controls (5 per leg) facil-
itates generation of accurate walking predictions compared to a model driven by muscle
activation controls (35 per leg) or joint torque controls (5 per leg). To explore this question,
we developed a subject-specic neuromusculoskeletal model of a single high-functioning
hemiparetic subject using instrumented treadmill walking data collected at the subject’s
self-selected speed of 0.5m/s. The model included subject-specic representations of
lower-body kinematic structure, foot–ground contact behavior, electromyography-driven
muscle force generation, and neural control limitations and remaining capabilities. Using
direct collocation optimal control and the subject-specic model, we evaluated the ability
of the three control approaches to predict the subject’s walking kinematics and kinetics at
two speeds (0.5 and 0.8m/s) for which experimental data were available from the subject.
We also evaluated whether synergy controls could predict a physically realistic gait period
at one speed (1.1m/s) for which no experimental data were available. All three control
approaches predicted the subject’s walking kinematics and kinetics (including ground
reaction forces) well for the model calibration speed of 0.5m/s. However, only activation
and synergy controls could predict the subject’s walking kinematics and kinetics well for
the faster non-calibration speed of 0.8m/s, with synergy controls predicting the new gait
period the most accurately. When used to predict how the subject would walk at 1.1m/s,
synergy controls predicted a gait period close to that estimated from the linear relationship
between gait speed and stride length. These ndings suggest that our neuromusculo-
skeletal simulation framework may be able to bridge the gap between patient-specic
muscle synergy information and resulting functional capabilities and limitations.
Keywords: biomechanics, computational neurorehabilitation, direct collocation optimal control, muscle synergy
analysis, neuromusculoskeletal modeling, predictive gait optimization

2
Meyer et al. Muscle Synergies Facilitate Walking Predictions
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org October 2016 | Volume 4 | Article 77
INTRODUCTION
Roughly one in six people worldwide will suer a stroke at some
point in their lifetime, with ~15 million people experiencing
a stroke each year (
World Stroke Organization, 2016). Due to
improvements in acute stroke management, the majority of these
individuals will survive their initial stroke, which has helped
make stroke a leading cause of serious, long-term disability in
adults worldwide (
Go etal., 2013; World Stroke Organization,
2016
). More than a third of stroke survivors experience signi-
cant physical disability (
Lloyd-Jones etal., 2010), with walking
dysfunction being among the greatest stroke-related limitations
contributing to disability. While the majority of persons who
suer a stroke regain some level of ambulatory function, their
gait is typically slow, asymmetrical, and metabolically inecient
(
Olney etal., 1986; Roth etal., 1997). Diminished walking ability
is tied to decreased quality of life, increased risk of depression,
and increased risk of serious secondary health conditions
(
Blair et al., 1989; Mutikainen et al., 2011; Ostir et al., 2013).
Restoration of walking function following a stroke is therefore
both a high priority for rehabilitation and an important public
health problem.
Despite recognition of the problem, current clinic-based
neurorehabilitation methods produce only modest improve-
ments in walking function for persons post-stroke (
States etal.,
2009
; Bogey and Hornby, 2014; Winstein etal., 2016). For this
reason, researchers and clinicians have explored a variety of
neurorehabilitation approaches in search of an eective means
to restore post-stroke walking function. ese approaches
include functional electrical stimulation (FES) (
Popovic et al.,
1999
; Kesar etal., 2009, 2010; Sabut etal., 2013; Chung etal.,
2014
; O’Dell etal., 2014; Pilkar etal., 2014; Auchstaetter etal.,
2015
; Chantraine et al., 2016), ankle–foot orthoses (AFOs)
(
Ferreira etal., 2013; Tyson etal., 2013; Kobayashi etal., 2016),
exoskeletons (
Nilsson etal., 2014; Bortole etal., 2015; Buesing
et al., 2015
), partial body weight support (Ng et al., 2008; Lee
etal., 2013
) and split-belt treadmill training systems (Reisman
etal., 2007
; Malone and Bastian, 2014), and robotic gait trainers
(
Pennycott et al., 2012; Mehrholz etal., 2013; Bae et al., 2014;
Hussain, 2014; Dundar et al., 2015). Each of these approaches
has shown varying levels of promise for improving post-stroke
walking function. However, a critical challenge is determin-
ing the treatment prescription – which approach to apply, how
much of the approach to apply, and how the approach should be
applied – that will maximize recovery of walking function for
any particular individual. Furthermore, there is currently no way
to identify whether a small amount of treatment provided by a
combination of approaches might be dramatically more eective
than a large amount of treatment provided by a single approach
(
Belda-Lois et al., 2011). Current treatment design methods
based on trial-and-error and subjective clinical judgment cannot
address these challenges, since they do not provide an objective
means for predicting a patients walking function following a
specied treatment or treatment combination.
One possible approach for overcoming this challenge is to
use patient-specic neuromusculoskeletal models to predict
post-treatment walking function for dierent neurorehabilitation
technologies (alone or combined) under consideration. Such
models should account for how the patient interacts with the
treatment approach (
Mooney and Herr, 2016) so that the optimal
prescription can be recommended based on objective predictions
of walking improvement. A number of studies have pursued
such modeling eorts by simulating the eects of FES (
Riener,
1999
; Heilman and Kirsch, 2003; Zhang and Zhu, 2007; Shao and
Buchanan, 2008
; Nekoukar and Erfanian, 2013; Sharma et al.,
2014
; Alibeji etal., 2015), exoskeletons (Fleischer and Hommel,
2008
; Afschri etal., 2014; Farris etal., 2014; Sawicki and Khan,
2015
), orthoses (Zmitrewicz etal., 2007; Crabtree and Higginson,
2009
; Silverman and Neptune, 2012), and strength training
(
Goldberg and Neptune, 2007; Knarr et al., 2014) on lower
extremity function and walking ability in the context of stroke,
spinal cord injury, amputee, and general rehabilitation. Few of
these studies focused on stroke (
Goldberg and Neptune, 2007;
Shao and Buchanan, 2008; Knarr etal., 2014), few used three-
dimensional models (
Fleischer and Hommel, 2008; Afschri
etal., 2014
; Farris etal., 2014; Knarr etal., 2014; Sawicki and Khan,
2015
), few used subject-specic models created by calibrating
critical neuromusculoskeletal model parameters to movement
data collected from an individual (Fleischer and Hommel, 2008;
Shao and Buchanan, 2008; Knarr et al., 2014), and only one
included modeling elements that accounted for subject-specic
neural control capabilities and limitations (Alibeji etal., 2015).
No study to date has predicted a stroke patients complete post-
treatment walking motion and speed resulting from application
of a specic neurorehabilitation intervention.
As a rst step toward optimizing patient interaction with
stroke neurorehabilitation technologies, this study describes
the development and evaluation of a subject-specic synergy-
controlled neuromusculoskeletal simulation framework that can
predict three-dimensional walking motions for an individual
post-stroke. e main question we address is whether actuating
a subject-specic neuromusculoskeletal model with muscle syn-
ergy controls (5 per leg) facilitates generation of accurate walking
predictions compared to actuating the model with muscle activa-
tion controls (35 per leg) or joint torque controls (5 per leg). We
hypothesize that synergy controls will work the best since they
combine a low number of control signals with a subject-specic
representation of the coupling between muscle activations within
each leg. We collect gait data from a stroke subject walking at
0.4, 0.5, 0.6, 0.7, and 0.8m/s on an instrumented treadmill and
use data from his self-selected speeds of 0.4–0.6m/s to develop
a subject-specic neuromusculoskeletal model. We incorporate
the subject-specic full-body model into a direct collocation
optimal control framework to predict new walking motions for
the subject. To evaluate the framework and the potential benets
of using synergy controls, we predict how the individual will walk
(including cadence and stride length) at 0.5 and 0.8m/s (condi-
tions for which experimental data are available for comparison)
using joint torque, muscle activation, or muscle synergy controls
and at 1.1m/s (a condition for which no experimental data are
available) using only synergy controls. With future simulation
of dierent neurorehabilitation approaches, our subject-specic
synergy-controlled neuromusculoskeletal simulation framework
may help identify optimal neurorehabilitation prescriptions that

3
Meyer et al. Muscle Synergies Facilitate Walking Predictions
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org October 2016 | Volume 4 | Article 77
maximize recovery of walking function on an individual patient
basis.
METHODS
Experimental Data Collection
To assist with development and evaluation of our subject-
specic synergy-controlled neuromusculoskeletal simulation
framework, we collected experimental walking data from one
high-functioning hemiparetic male individual with chronic
stroke-related walking dysfunction (age 79years, LE Fugl-Meyer
Motor Assessment 32/34 pts, right-sided hemiparesis, height
1.7 m, mass 80.5 kg). All study procedures were approved by
the University of Florida Health Science Center Institutional
Review Board (IRB-01) and the Malcom Randall VA Medical
Center Research and Development Committee and included
approval to study individuals with stroke-related disability. Study
personnel obtained written informed consent prior to participant
enrollment and involvement in study activities. Study procedures
were conducted in accordance with the Declaration of Helsinki.
Motion capture (Vicon Corp., Oxford, UK), ground reaction
(Bertec Corp., Columbus, OH, USA), and electromyography
(EMG) data (Motion Lab Systems, Baton Rouge, LA, USA)
were collected simultaneously while the participant walked on
a split-belt instrumented treadmill (Bertec Corp., Columbus,
OH, USA) at speeds ranging from 0.4 to 0.8m/s in increments of
0.1m/s. 0.8m/s was the fastest speed at which the subject could
walk safely without assistance. is range of speeds included the
participants self-selected walking speed of 0.5m/s. More than 50
gait cycles were recorded at each walking speed. A static stand-
ing trial was collected for model scaling purposes. To facilitate
subsequent creation of subject-specic foot–ground contact
models, the participant wore Adidas Samba Classic sneakers,
which have a at sole and neutral midsole with no cushioning,
and we collected additional static trials where we used a marker
wand to trace the outline of each sneaker sole on the ground.
Motion capture data were obtained using a modied Cleveland
Clinic full-body marker set with additional markers added to the
feet (
Reinbolt etal., 2005). Marker motion and ground reaction
data were ltered at a variable cut-o frequency of 7/tfHz, where
tf is the period of the gait cycle being processed, using a fourth-
order zero phase lag Butterworth lter (
Hug, 2011). is variable
cut-o frequency would cause data collected at a normal walking
speed to be ltered at ~6Hz.
Electromyography data were collected from 16 muscles in each
leg and processed using standard methods (Lloyd and Besier,
2003
). A combination of surface and ne-wire EMG electrodes
was used. Surface EMG data were collected for gluteus maximus
and medius, semimembranosus, biceps femoris long head, rectus
femoris, vastus medialis and lateralis, medial gastrocnemius, tibi-
alis anterior, peroneus longus, and soleus. Fine-wire EMG data
were collected for adductor longus, iliopsoas, tibialis posterior,
extensor digitorum longus, and exor digitorum longus. All
EMG data were high-pass ltered at 40Hz (
Lloyd and Besier,
2003), demeaned, rectied, and then low-pass ltered at a vari-
able cut-o frequency 3.5/tfHz. Filtering was performed using a
fourth-order zero phase lag Butterworth lter. EMG data from
each muscle were normalized to the maximum value over all trials
and resampled to 101 time points per gait cycle (heel strike to heel
strike for the less involved le side) while keeping an additional
20 time points before the start of the cycle to permit modeling of
electromechanical delay. In addition, each processed EMG signal
was oset on a cycle-by-cycle basis so that the minimum value
was zero.
Neuromusculoskeletal Model
Development
e subject-specic neuromusculoskeletal model that served as
the foundation for our simulation framework incorporated four
modeling components to account for the unique neurophysio-
logical and musculoskeletal characteristics of the subject: (1) a
subject-specic lower-body kinematic model to simulate the
subjects skeletal motion, (2) subject-specic foot–ground con-
tact models to simulate how the subjects feet interact with the
ground, (3) subject-specic EMG-driven muscle moment models
to simulate the subject’s lower extremity joint moments, and (4)
a subject-specic muscle synergy control model to simulate the
subjects neural control system. Below we describe each of these
four components in further detail. Unless otherwise noted, we
calibrated model parameters in each component using a single
representative walking trial collected at the subjects self-selected
speed of 0.5m/s.
Subject-Specic Lower-Body Kinematic Model
Our neuromusculoskeletal model creation process started with
a generic full-body musculoskeletal model (
Arnold etal., 2010;
Hamner etal., 2010) developed in OpenSim (Delp etal., 2007).
e original model included 37 degrees of freedom (DOFs) and
44 Hill-type muscle-tendon actuators in each leg. We locked the
wrist and forearm pronation–supination angles to anatomically
reasonable values for walking, leaving the following 31 DOFs:
6 DOF ground-to-pelvis joint, 3 DOF hip joints, 1 DOF knee
joints, 1 DOF ankle joints, 1 DOF subtalar joints, 1 DOF
toejoints connecting rear foot and toe segments, 3 DOF back
joint, 3 DOF shoulder joints, and 1 DOF elbow joints. We also
eliminated nine muscle-tendon actuators without related EMG
data (extensor hallucis longus, exor hallucis longus, gemelli,
gracilis, pectineus, piriformis, quadratus femoris, sartorius,
tensor fascia latae), leaving 35 muscles per leg that actuated
hip exion-extension, hip adduction-abduction, knee exion-
extension, ankle exion-extension, and ankle inversion–ever-
sion on each leg. We then scaled the modied model using the
standing static trial marker data and the OpenSim “Scale Model”
tool, where distances between markers placed over identiable
landmarks were averaged between the two sides to maintain
bilateral symmetry following scaling.
Once the model was scaled, we calibrated joint and marker
positions in the torso, pelvis, and lower-body portion of the
OpenSim model using marker data from a representative walking
trial. e calibration approach was similar to one described pre-
viously (
Reinbolt etal., 2005, 2008) except that it was performed
on the scaled OpenSim model using the OpenSim-MATLAB
Application Programing Interface and included modications
to maintain correct bone geometry positions within the body

4
Meyer et al. Muscle Synergies Facilitate Walking Predictions
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org October 2016 | Volume 4 | Article 77
segments (Charlton et al., 2004). To facilitate the calibration
process, we created marker plates on the torso, pelvis, thighs,
and shanks to which all markers on the respective OpenSim
body segments were attached. To perform the actual calibration,
we used non-linear least squares optimization (lsqnonlin) in
MATLAB to adjust joint (knee, ankle, and subtalar in both legs)
and marker plate (torso, pelvis, thighs, and shanks) positions
and orientations in their respective body segments such that
marker errors from repeated OpenSim inverse kinematic analy-
ses were minimized. e optimization cost function included
penalty terms that prevented large changes in joint and marker
plate positions and orientations that would produce only small
improvements in marker tracking. Modication of the two hip
joint center locations was achieved by modifying the position
and orientation of the rigid marker plate on the pelvis. For joint
centers and orientations, symmetry between le and right sides
of the body was maintained during the kinematic calibration
process. Markers on the feet were not adjusted since their loca-
tions were well dened. e position and orientation of the toe
axis in each foot and of the back, shoulder, and elbow joints was
maintained from the scaled OpenSim model.
Subject-Specic Foot–Ground Contact Models
Following kinematic calibration, we created a subject-specic
foot–ground contact model for each foot of the OpenSim model
using recently developed methods (
Jackson et al., 2016). e
elastic foundation contact models were developed in MATLAB
and used a grid of contact elements that spanned the rear foot
and toes segments of each foot. To create the element grid, we
started with the shoe outlines obtained from the static trial
marker data and used principal component analysis to identify
the principal axes of each foot (rear foot and toes segments
together). Using these axes, we constructed an 11 (anterior-
posterior) × 8 (medial-lateral) grid of rectangular contact
elements for the le foot, where the edges of the grid extended
2.5mm beyond the edge of the shoe outline in both directions.
Forty-seven elements whose centers fell within the shoe outline
were retained in the contact model, while 41 elements whose
centers fell outside the shoe outline were removed. Given the
locations of the MATLAB contact element centers relative to the
foot markers from the static le shoe outline trial, we calculated
the locations of the element centers on the OpenSim rear foot
and toes segments. We then projected the le toes axis of the
OpenSim model onto the contact element grid. Elements whose
centers were posterior to the axis were assigned to the rear foot
segment, while elements whose centers were anterior to the axis
were assigned to the toes segment. e complete MATLAB/
OpenSim contact element grid for the le foot was mirrored to
the right foot by aligning the principal axes of the mirrored grid
with those of the right foot.
Each contact element in the foot–ground contact models
generated normal force using a linear spring with non-linear
damping and shear force using a continuous stick-slip friction
model. For any contact element i, the required time-varying
inputs for contact force calculations performed in MATLAB were
the penetration into the oor y
i
, the normal penetration rate
y
i
,
and the shear slip rate
v
s
i
of the element center in the laboratory
coordinate system as calculated by OpenSim. e normal contact
force F
i
for element i was calculated as (
Hunt and Crossley, 1975)
Fkyy cy
iiii
=−+()
()
0
1
(1)
where k
i
is the spring stiness unique to each spring, y
0
is the
spring resting length common to all springs on the same foot
and essentially adjusts the height of the oor, and c (=1e–2) is a
non-linear damping coecient common to all springs. e linear
spring also generates a small amount of force when the foot is o
the oor, and this negligible force transitions in a smooth and
continuous way to the large force produced when the spring is in
contact with the ground (
Anderson and Pandy, 2001; Ackermann
and van den Bogert, 2010
). e non-linear damping ensures
that the normal contact force does not exhibit damping-related
discontinuities when a spring enters or leaves contact. e shear
contact force f
i
for element i was calculated using a simple con-
tinuous and dierentiable friction model (
Ackermann and van
den Bogert, 2010
)
fF
v
v
ii
s
l
i
=
µtanh
(2)
where μ [=1 (
Ackermann and van den Bogert, 2010)] is a dynamic
friction coecient common to all springs and v
l
(=5cm/s) is a
latching speed common to all springs that denes the edge of a
linear transition region between zero slip rate and the start of
dynamic friction. Shear contact force f
i
was applied to the element
center in the direction opposite to the slip velocity vector. e
direction calculation included a small constant value of 1e4 in
the denominator to avoid division by a small number when the
slip speed was near zero. Once F
i
and f
i
were calculated for each
contact element, the net contact force and torque due to all contact
elements in the rear foot segment were calculated with respect to
the rear foot origin, and similarly for the toes segment using the
toes origin (
Kane and Levinson, 1985). ese net contact forces
and torques were then applied to their corresponding segments
in the OpenSim model. is approach allowed rear foot and toes
contributions to total ground reaction force to be predicted by
the model.
We calibrated the spring stiness k
i
of each contact element
in both feet and the spring resting length y
0
for all contact ele-
ments in each foot using marker and ground reaction data from
a representative walking trial. We made two assumptions about
the spring stiness distribution across the bottom of the shoe to
simplify the calibration process. First, we assumed that the mir-
rored stiness distribution was the same for both feet. Second,
we assumed that the stiness distribution across the entire shoe
bottom could be approximated by a three-dimensional parabolic
surface, which possesses only six unknown coecients rather
than 47 unknown independent spring stiness values and pre-
vents neighboring springs from having dramatically dierent
stinesses. To calibrate these six coecients and the two resting
lengths, we formulated a direct collocation optimal control
problem that tracked experimental marker, ground reaction, and
inverse dynamic joint torque data for the entire body with higher
weight placed on matching marker position and ground reaction
data for the two feet. Tracked ground reaction quantities for each

Citations
More filters
Journal ArticleDOI
TL;DR: OpenSim is an extensible and user-friendly software package built on decades of knowledge about computational modeling and simulation of biomechanical systems that enables computational scientists to create new state-of-the-art software tools and empowers others to use these tools in research and clinical applications.
Abstract: Movement is fundamental to human and animal life, emerging through interaction of complex neural, muscular, and skeletal systems. Study of movement draws from and contributes to diverse fields, including biology, neuroscience, mechanics, and robotics. OpenSim unites methods from these fields to create fast and accurate simulations of movement, enabling two fundamental tasks. First, the software can calculate variables that are difficult to measure experimentally, such as the forces generated by muscles and the stretch and recoil of tendons during movement. Second, OpenSim can predict novel movements from models of motor control, such as kinematic adaptations of human gait during loaded or inclined walking. Changes in musculoskeletal dynamics following surgery or due to human–device interaction can also be simulated; these simulations have played a vital role in several applications, including the design of implantable mechanical devices to improve human grasping in individuals with paralysis. OpenSim is an extensible and user-friendly software package built on decades of knowledge about computational modeling and simulation of biomechanical systems. OpenSim’s design enables computational scientists to create new state-of-the-art software tools and empowers others to use these tools in research and clinical applications. OpenSim supports a large and growing community of biomechanics and rehabilitation researchers, facilitating exchange of models and simulations for reproducing and extending discoveries. Examples, tutorials, documentation, and an active user forum support this community. The OpenSim software is covered by the Apache License 2.0, which permits its use for any purpose including both nonprofit and commercial applications. The source code is freely and anonymously accessible on GitHub, where the community is welcomed to make contributions. Platform-specific installers of OpenSim include a GUI and are available on simtk.org.

642 citations

Journal ArticleDOI
TL;DR: OpenSim Moco as mentioned in this paper is a software toolkit for optimizing the motion and control of musculoskeletal models built in the OpenSim modeling and simulation package, which can handle a wide range of problems that interest biomechanists, including motion tracking, motion prediction, model fitting, electromyography-driven simulation, and device design.
Abstract: Musculoskeletal simulations are used in many different applications, ranging from the design of wearable robots that interact with humans to the analysis of patients with impaired movement. Here, we introduce OpenSim Moco, a software toolkit for optimizing the motion and control of musculoskeletal models built in the OpenSim modeling and simulation package. OpenSim Moco uses the direct collocation method, which is often faster and can handle more diverse problems than other methods for musculoskeletal simulation. Moco frees researchers from implementing direct collocation themselves-which typically requires extensive technical expertise-and allows them to focus on their scientific questions. The software can handle a wide range of problems that interest biomechanists, including motion tracking, motion prediction, parameter optimization, model fitting, electromyography-driven simulation, and device design. Moco is the first musculoskeletal direct collocation tool to handle kinematic constraints, which enable modeling of kinematic loops (e.g., cycling models) and complex anatomy (e.g., patellar motion). To show the abilities of Moco, we first solved for muscle activity that produced an observed walking motion while minimizing squared muscle excitations and knee joint loading. Next, we predicted how muscle weakness may cause deviations from a normal walking motion. Lastly, we predicted a squat-to-stand motion and optimized the stiffness of an assistive device placed at the knee. We designed Moco to be easy to use, customizable, and extensible, thereby accelerating the use of simulations to understand the movement of humans and other animals.

70 citations

Posted ContentDOI
12 Nov 2019-bioRxiv
TL;DR: Open-Sim Moco, a software toolkit for optimizing the motion and control of musculoskeletal models built in the OpenSim modeling and simulation package, is introduced, which is the first musculOSkeletal direct collocation tool to handle kinematic constraints, which are common in musculo-kinematic models.
Abstract: Musculoskeletal simulations of movement can provide insights needed to help humans regain mobility after injuries and design robots that interact with humans. Here, we introduce Open-Sim Moco, a software toolkit for optimizing the motion and control of musculoskeletal models built in the OpenSim modeling and simulation package. OpenSim Moco uses the direct collocation method, which is often faster and can handle more diverse problems than other methods for musculoskeletal simulation but requires extensive technical expertise to implement. Moco frees researchers from implementing direct collocation themselves, allowing them to focus on their scientific questions. The software can handle the wide range of problems that interest biomechanists, including motion tracking, motion prediction, parameter optimization, model fitting, electromyography-driven simulation, and device design. Moco is the first musculoskeletal direct collocation tool to handle kinematic constraints, which are common in musculoskeletal models. To show Moco’s abilities, we first solve for muscle activity that produces an observed walking motion while minimizing muscle excitations and knee joint loading. Then, we predict a squat-to-stand motion and optimize the stiffness of a passive assistive knee device. We designed Moco to be easy to use, customizable, and extensible, thereby accelerating the use of simulations to understand human and animal movement.

69 citations


Cites methods from "Muscle Synergies Facilitate Computa..."

  • ...The advantages of direct collocation have led biomechanists to use the method for tracking motions [16,23], predicting motions [24–33], fitting muscle properties [34], and optimizing design parameters [35]....

    [...]

Journal ArticleDOI
TL;DR: A set of polynomial expressions that can be used as regression equations to estimate length and three-dimensional moment arms of 43 lower-limb musculotendon actuators allow one to find, at a low computational cost, the musculOTendon geometric parameters required for numerical simulation of large musculoskeletal models.

58 citations

Journal ArticleDOI
11 Jul 2017-PLOS ONE
TL;DR: This study presents a novel EMG-driven modeling method that automatically adjusts surrogate representations of the patient’s musculoskeletal geometry to improve prediction of lower extremity joint moments during walking.
Abstract: Neuromusculoskeletal disorders affecting walking ability are often difficult to manage, in part due to limited understanding of how a patient's lower extremity muscle excitations contribute to the patient's lower extremity joint moments. To assist in the study of these disorders, researchers have developed electromyography (EMG) driven neuromusculoskeletal models utilizing scaled generic musculoskeletal geometry. While these models can predict individual muscle contributions to lower extremity joint moments during walking, the accuracy of the predictions can be hindered by errors in the scaled geometry. This study presents a novel EMG-driven modeling method that automatically adjusts surrogate representations of the patient's musculoskeletal geometry to improve prediction of lower extremity joint moments during walking. In addition to commonly adjusted neuromusculoskeletal model parameters, the proposed method adjusts model parameters defining muscle-tendon lengths, velocities, and moment arms. We evaluated our EMG-driven modeling method using data collected from a high-functioning hemiparetic subject walking on an instrumented treadmill at speeds ranging from 0.4 to 0.8 m/s. EMG-driven model parameter values were calibrated to match inverse dynamic moments for five degrees of freedom in each leg while keeping musculoskeletal geometry close to that of an initial scaled musculoskeletal model. We found that our EMG-driven modeling method incorporating automated adjustment of musculoskeletal geometry predicted net joint moments during walking more accurately than did the same method without geometric adjustments. Geometric adjustments improved moment prediction errors by 25% on average and up to 52%, with the largest improvements occurring at the hip. Predicted adjustments to musculoskeletal geometry were comparable to errors reported in the literature between scaled generic geometric models and measurements made from imaging data. Our results demonstrate that with appropriate experimental data, joint moment predictions for walking generated by an EMG-driven model can be improved significantly when automated adjustment of musculoskeletal geometry is included in the model calibration process.

51 citations

References
More filters
Journal ArticleDOI
21 Oct 1999-Nature
TL;DR: An algorithm for non-negative matrix factorization is demonstrated that is able to learn parts of faces and semantic features of text and is in contrast to other methods that learn holistic, not parts-based, representations.
Abstract: Is perception of the whole based on perception of its parts? There is psychological and physiological evidence for parts-based representations in the brain, and certain computational theories of object recognition rely on such representations. But little is known about how brains or computers might learn the parts of objects. Here we demonstrate an algorithm for non-negative matrix factorization that is able to learn parts of faces and semantic features of text. This is in contrast to other methods, such as principal components analysis and vector quantization, that learn holistic, not parts-based, representations. Non-negative matrix factorization is distinguished from the other methods by its use of non-negativity constraints. These constraints lead to a parts-based representation because they allow only additive, not subtractive, combinations. When non-negative matrix factorization is implemented as a neural network, parts-based representations emerge by virtue of two properties: the firing rates of neurons are never negative and synaptic strengths do not change sign.

11,500 citations

01 Jan 1999
TL;DR: In this article, non-negative matrix factorization is used to learn parts of faces and semantic features of text, which is in contrast to principal components analysis and vector quantization that learn holistic, not parts-based, representations.
Abstract: Is perception of the whole based on perception of its parts? There is psychological and physiological evidence for parts-based representations in the brain, and certain computational theories of object recognition rely on such representations. But little is known about how brains or computers might learn the parts of objects. Here we demonstrate an algorithm for non-negative matrix factorization that is able to learn parts of faces and semantic features of text. This is in contrast to other methods, such as principal components analysis and vector quantization, that learn holistic, not parts-based, representations. Non-negative matrix factorization is distinguished from the other methods by its use of non-negativity constraints. These constraints lead to a parts-based representation because they allow only additive, not subtractive, combinations. When non-negative matrix factorization is implemented as a neural network, parts-based representations emerge by virtue of two properties: the firing rates of neurons are never negative and synaptic strengths do not change sign.

9,604 citations

Journal ArticleDOI
TL;DR: The Statistical Update brings together the most up-to-date statistics on heart disease, stroke, other vascular diseases, and their risk factors and presents them in its Heart Disease and Stroke Statistical Update each year.
Abstract: Appendix I: List of Statistical Fact Sheets. URL: http://www.americanheart.org/presenter.jhtml?identifier=2007 We wish to thank Drs Brian Eigel and Michael Wolz for their valuable comments and contributions. We would like to acknowledge Tim Anderson and Tom Schneider for their editorial contributions and Karen Modesitt for her administrative assistance. Disclosures View this table: View this table: View this table: # Summary {#article-title-2} Each year, the American Heart Association, in conjunction with the Centers for Disease Control and Prevention, the National Institutes of Health, and other government agencies, brings together the most up-to-date statistics on heart disease, stroke, other vascular diseases, and their risk factors and presents them in its Heart Disease and Stroke Statistical Update. The Statistical Update is a valuable resource for researchers, clinicians, healthcare policy makers, media professionals, the lay public, and many others who seek the best national data available on disease …

6,176 citations

Journal ArticleDOI
03 Nov 1989-JAMA
TL;DR: Higher levels of physical fitness appear to delay all-cause mortality primarily due to lowered rates of cardiovascular disease and cancer, and lower mortality rates in higher fitness categories also were seen for cardiovascular Disease and cancer of combined sites.
Abstract: We studied physical fitness and risk of all-cause and cause-specific mortality in 10 224 men and 3120 women who were given a preventive medical examination. Physical fitness was measured by a maximal treadmill exercise test. Average follow-up was slightly more than 8 years, for a total of 110 482 person-years of observation. There were 240 deaths in men and 43 deaths in women. Age-adjusted all-cause mortality rates declined across physical fitness quintiles from 64.0 per 10 000 person-years in the least-fit men to 18.6 per 10 000 person-years in the most-fit men (slope, —4.5). Corresponding values for women were 39.5 per 10 000 person-years to 8.5 per 10 000 person-years (slope, —5.5). These trends remained after statistical adjustment for age, smoking habit, cholesterol level, systolic blood pressure, fasting blood glucose level, parental history of coronary heart disease, and follow-up interval. Lower mortality rates in higher fitness categories also were seen for cardiovascular disease and cancer of combined sites. Attributable risk estimates for all-cause mortality indicated that low physical fitness was an important risk factor in both men and women. Higher levels of physical fitness appear to delay all-cause mortality primarily due to lowered rates of cardiovascular disease and cancer. ( JAMA . 1989;262:2395-2401)

3,957 citations


"Muscle Synergies Facilitate Computa..." refers background in this paper

  • ...Diminished walking ability is tied to decreased quality of life, increased risk of depression, and increased risk of serious secondary health conditions (Blair et  al., 1989; Mutikainen et  al., 2011; Ostir et  al., 2013)....

    [...]

Journal ArticleDOI
TL;DR: OpenSim is developed, a freely available, open-source software system that lets users develop models of musculoskeletal structures and create dynamic simulations of a wide variety of movements to simulate the dynamics of individuals with pathological gait and to explore the biomechanical effects of treatments.
Abstract: Dynamic simulations of movement allow one to study neuromuscular coordination, analyze athletic performance, and estimate internal loading of the musculoskeletal system. Simulations can also be used to identify the sources of pathological movement and establish a scientific basis for treatment planning. We have developed a freely available, open-source software system (OpenSim) that lets users develop models of musculoskeletal structures and create dynamic simulations of a wide variety of movements. We are using this system to simulate the dynamics of individuals with pathological gait and to explore the biomechanical effects of treatments. OpenSim provides a platform on which the biomechanics community can build a library of simulations that can be exchanged, tested, analyzed, and improved through a multi-institutional collaboration. Developing software that enables a concerted effort from many investigators poses technical and sociological challenges. Meeting those challenges will accelerate the discovery of principles that govern movement control and improve treatments for individuals with movement pathologies.

3,621 citations