2001_Anderson_Pandy_JB.pdf

(424 KB) Pobierz
PII: S0021-9290(00)00155-X
Journal of Biomechanics 34 (2001) 153}161
1999 ASB Pre-Doctoral Award
Static and dynamic optimization solutions for gait
are practically equivalent
Frank C. Anderson !, * , Marcus G. Pandy !,"
! Department of Mechanical Engineering, The University of Texas at Austin, Austin, Texas, USA
" Department of Kinesiology and Health Education, The University of Texas at Austin, Austin, Texas, USA
Accepted 3 July 2000
Abstract
The proposition that dynamic optimization provides better estimates of muscle forces during gait than static optimization is
examined by comparing a dynamic solution with two static solutions. A 23-degree-of-freedom musculoskeletal model actuated by 54
Hill-type musculotendon units was used to simulate one cycle of normal gait. The dynamic problem was to "nd the muscle excitations
which minimized metabolic energy per unit distance traveled, and which produced a repeatable gait cycle. In the dynamic problem,
activation dynamics was described by a "rst-order di!erential equation. The joint moments predicted by the dynamic solution were
used as input to the static problems. In each static problem, the problem was to "nd the muscle activations which minimized the sum
of muscle activations squared, and which generated the joint moments input from the dynamic solution. In the "rst static problem,
muscles were treated as ideal force generators; in the second, they were constrained by their force}length}velocity properties; and in
both, activation dynamics was neglected. In terms of predicted muscle forces and joint contact forces, the dynamic and static solutions
were remarkably similar. Also, activation dynamics and the force}length}velocity properties of muscle had little in#uence on the static
solutions. Thus, for normal gait, if one can accurately solve the inverse dynamics problem and if one seeks only to estimate muscle
forces, the use of dynamic optimization rather than static optimization is currently not justi"ed. Scenarios in which the use of dynamic
optimization is justi"ed are suggested.
(
2001 Elsevier Science Ltd. All rights reserved.
Keywords: Dynamic optimization; Static optimization; Gait; Articular contact forces; Muscle forces
1. Introduction
and processing of body segmental kinematics (Patriarco
et al., 1981; Davy and Audu, 1987). Second, while it is
purportedly necessary to account for the underlying
physiological properties of muscle in order to obtain
more accurate estimates of muscle force (Hardt, 1978; An
et al., 1989; Kaufman et al., 1991; Prilutsky et al., 1997),
the time-independent nature of static optimization makes
it relatively di$cult to incorporate muscle physiology
properly. Third, the time-independence of the perfor-
mance criterion required by static optimization may not
permit the objectives of the motor task to be properly
characterized (Hardt, 1978; Pandy et al., 1995). Finally,
analyses based on an inverse dynamics approach may
not be appropriate for explaining muscle coordination
principles (Kautz et al., 2000; Zajac, 1993).
Dynamic optimization, a forward dynamics approach,
is not subject to these criticisms, and ostensibly can
provide more realistic estimates of muscle force. Dy-
namic optimization integrates system dynamics into the
Static optimization, an inverse dynamics approach,
has been used extensively to estimate in vivo muscle
forces during gait (Seireg and Arvikar, 1975; Pedotti et
al., 1978; Hardt, 1978; Crowninshield et al., 1978; Crow-
ninshield and Brand, 1981; Patriarco et al., 1981; RoK hrle
et al., 1984; Brand et al., 1986; Pedersen et al., 1997;
Glitsch and Baumann, 1997). Static models are computa-
tionally e$cient, have allowed full three-dimensional
motion, and have generally incorporated 30 or more
muscles per leg. However, static optimization has been
criticized on several points. First, the inverse dynamics
problem is highly dependent on the accurate collection
* Correspondence address. Biomechanical Engineering Division, De-
partment of Mechanical Engineering, Stanford University, Stanford,
California 94305-3030, USA. Fax: 650-725-1587.
E-mail address: fca@stanford.edu (F.C. Anderson).
0021-9290/01/$ - see front matter ( 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 2 1 - 9 2 9 0 ( 0 0 ) 0 0 1 5 5 - X
207207430.005.png
154
F.C. Anderson, M.G. Pandy / Journal of Biomechanics 34 (2001) 153 } 161
solution process: quantities like muscle forces and the
performance criterion are treated as time-dependent state
variables whose behavior is governed by sets of di!eren-
tial equations (Hatze, 1976). Ideally, the di!erential equa-
tions accurately represent the underlying physiological
properties of the system, and so the time histories of
muscle forces predicted are consistent with those forces
which could naturally arise during movement.
Unfortunately, dynamic optimization incurs so much
computational expense that relatively few dynamic solu-
tions for gait have been found (Chow and Jacobson,
1971; Davy and Audu, 1987; Yamaguchi and Zajac, 1990;
Tashman et al., 1995). Further, for gait, this approach has
required that the dynamic models be simpli"ed to such
an extent that it has been di$cult to ascertain whether its
computational expense is justi"ed. While Davy and
Audu (1987) solved static and dynamic optimization
problems for gait as part of the same study and found
di!erences between the two solutions, it is di$cult to
draw conclusions based on these di!erences because (1)
the authors simulated only the swing phase of gait, (2) the
model was con"ned to the sagittal plane and included
only nine muscles, and (3) the joint moments used as
input to the static optimization problem were di!erent
from those predicted by the dynamic optimization solu-
tion.
The aim of this work is to assess whether or not
dynamic optimization provides more realistic estimates
of muscle force than static optimization. In particular, we
address the following questions:
When estimating in vivo muscle and joint contact
forces during gait
2. Methods
2.1. Musculoskeletal model
The body was modeled as a 10-segment, 23-degree-of-
freedom linkage (Fig. 1). The inertial properties of the
segments were based on anthropometric measures ob-
tained from "ve healthy adult males (age 26$3 yr,
height 177$3 cm, and mass 70.1$7.8 kg) (McConville
et al., 1980). Interactions of the feet with the ground
were modeled using a series of spring-damper units
distributed under the sole of each foot (Anderson and
Pandy, 1999).
The model was controlled by a total of 54 musculoten-
dinous actuators (Fig. 2). Each leg was controlled by 24
actuators. Relative movements of the pelvis and upper
body were controlled by six abdominal and back ac-
tuators. Each actuator was modeled as a three-element,
Hill-type muscle in series with tendon (Zajac, 1989). The
muscle parameters, as well as the origin and insertion
sites, were based on data reported by Delp (1990). The
actions of ligaments were used to prevent anatomically
infeasible joint angles from arising during a simulation.
Each ligament was modeled as an exponential curve
which provided a restoring torque about a joint axis as
a function of joint angle (Davy and Audu, 1987). For
(1) Is the large computational cost of dynamic optimiza-
tion justi"ed?
(2) Is it important to account for muscle physiology,
namely activation dynamics and the force}length}
velocity properties of muscle?
The basis for the assessment is a comparison of a
previously attained dynamic optimization solution for
gait (Anderson, 1999; Anderson and Pandy, submitted)
with two analogous static optimization solutions, one in
which the force}length}velocity properties of muscle
were taken into account and one in which they were
neglected. The musculoskeletal model used allowed
three-dimensional motion of the body parts and simula-
tion of a full gait cycle. Furthermore, the model was
actuated by a number of muscles which approaches that
used in earlier static optimization models. Signi"cantly,
the joint torques predicted by the dynamic optimization
solution were used as input to the static optimization
problems. In this way, the experimental errors endemic
to solving the inverse dynamics problem were eliminated,
and the results of the static and dynamic optimization
solutions could be compared more directly.
Fig. 1. Model of the body. The "rst six-degrees-of-freedom were used to
de"ne the position and orientation of the pelvis relative to the ground.
The remaining nine segments branched out in an open chain from the
pelvis. The head, arms, and torso were represented as a single rigid body
which articulated with the pelvis via a three-dof ball-and-socket joint
located at approximately the third lumbar vertebra. Each hip was
modeled as a three-dof ball-and-socket joint, each knee as a one-dof
hinge joint, each ankle-subtalar joint as a universal joint with a single
joint center, and each metatarsal joint as a hinge joint. The directions of
the knee, ankle, subtalar, and metatarsal joint axes were anatomical
and based on in vivo and cadaveric measurements (Inman, 1976;
Anderson, 1999; Anderson and Pandy, 1999).
 
207207430.006.png
F.C. Anderson, M.G. Pandy / Journal of Biomechanics 34 (2001) 153 } 161
155
served as controls in the dynamic optimization problem:
u
m
, 2 , u
1 m
,
a
m
(0), m "1, 2 , 54,
(1)
where u i m
is the i th muscle excitation node for muscle
(0) is the activation level of muscle m at t "0in
the simulation. The dependence of muscle activation
level on excitation level was described by a "rst-order
di!erential equation
m
a
5 "( u
2 ! ua )/
q 3*4%
#( u ! a )/
q &!--
, a " a
m
( t ), u " u
m
( t ),
(2)
where a
m
( t ) is the activation level of muscle m at time t ,
( t ) is the excitation level of muscle m at time t , and
q 3*4%
m
(200 ms) are the rise and decay
constants of muscle activation (Pandy et al., 1992).
The dynamic optimization problem was minimally
constrained: only those constraints necessary to ensure
a repeatable gait cycle were imposed. Speci"cally, ter-
minal boundary constraints were applied to the joint
angular displacements, joint angular velocities, muscle
activation levels, and muscle excitation levels so that the
states at the end of the simulation were approximately
equal to those at t "0. Because only one-half of the gait
cycle was simulated, terminal states for the left-hand side
of the body were constrained to equal the initial states for
the right-hand side of the body, and vice-versa:
(22 ms) and
q &!--
Fig. 2. Muscles in the model. Abbreviations used for the muscles are as
follows: (ERCSPN) erector spinae; (EXTOBL) external abdominal ob-
liques; (INTOBL) internal abdominal obliques; (ILPSO) iliopsoas;
(ADLB) adductor longus brevis; (ADM) adductor magnus; (GMEDA)
anterior gluteus medius and anterior gluteus minimus; (GMEDP) pos-
terior gluteus medius and posterior gluteus minimus; (GMAXM) me-
dial gluteus maximus; (GMAXL) lateral gluteus maximus; (TFL) tensor
fasciae latae; (SAR) sartorius; (GRA) gracilis; (HAMS) semimem-
branosus, semitendinosus, and biceps femoris long head; (RF) rectus
femoris; (VAS) vastus medialis, vastus intermedius, and vastus lateralis;
(BFSH) biceps femoris short head; (GAS) gastrocnemius; (SOL) soleus;
(PFEV) peroneus brevis and peroneus longus; (DFEV) peroneus tertius
and extensor digitorum; (DFIN) tibialis anterior and extensor hallucis
longus; (PFIN) tibialis posterior, #exor digitorum longus, and #exor
hallucis longus. Muscles included in the model but not shown in the
diagram are: (PIRI) piriformis; (PECT) pectinius; (FDH) #exor
digitorum longus/brevis and #exor hallucis longus/brevis and; (EDH)
extensor digitorum longus/brevis and extensor hallucis longus/brevis.
q
#-
i
( t
&
)" q
i
(0), i "4, 2 , 23,
(3)
q
5 #-
i
( t
&
)" q
5 i
(0), i "4, 2 , 23,
(4)
a
# m
( t
&
)" a
m
(0), m "1, 2 , 54,
(5)
u 15,#-
m
" u m
, m "1, 2 , 54,
(6)
( t ) are the i th generalized coordinate
and generalized velocity of the model, respectively, a
i
( t ) and q
5 i
details concerning the model refer to Anderson (1999)
and Anderson and Pandy (1999).
( t )
m
are the
15th and 1st control node for muscle m , respectively. The
superscript cl indicates the contralateral quantity. For
example, in Eq. (5) the contralateral quantity is the ac-
tivation level of the contralateral muscle. Note that con-
straints were not placed on the generalized coordinates
or velocities relating to the translation of the pelvis (i.e,
i "1, 2 , 3 in Eqs. (3) and (4)).
There is much experimental evidence to suggest that
people select walking speeds which minimize the meta-
bolic energy expended per unit distance traveled (Ral-
ston, 1958,1976; Corcoran and Brengelmann, 1970;
Finley and Cody, 1970). The performance criterion for
the dynamic optimization problem was therefore ex-
pressed as follows
1 m
and u
m
2.2. Dynamic optimization problem
A "xed-"nal-time dynamic optimization problem was
solved for one cycle of normal gait. Because the gait cycle
was assumed to be bilaterally symmetric, it was only
necessary to simulate one-half of the gait cycle. The initial
states were obtained by averaging kinematic and force
platform data obtained from "ve subjects. The "nal time
for the dynamic optimization problem was "xed to the
average time taken by the "ve subjects to complete half of
a gait cycle, which was 0.56 s.
The excitation history for each muscle was para-
meterized by 15 nodal points which were allowed to vary
continuously between zero and one, zero indicating no
excitation and one indicating maximal excitation. The
nodal points representing the muscle excitations, to-
gether with the initial activation level for each muscle,
P
t &
A
B Q # 5 +
m /0
( A Q
# M Q
# S Q
m
#=Q
)
B
d t
m
m
m
J "
0
,
(7)
X
( t
)! X
(0)
#.
&
#.
m and a
u
where q
is the activation level of muscle m , and u
 
207207430.001.png
156
F.C. Anderson, M.G. Pandy / Journal of Biomechanics 34 (2001) 153 } 161
where the numerator is a computation of the total meta-
bolic energy consumed and the denominator is the
change in position of the center of mass in the direction of
progression. The equations used to estimate metabolic
energy from the internal states of the muscles were de-
veloped based largely on the work of Davy and Audu
(1987), Hatze and Buys (1977), and Mommaerts (1969).
B Q is the basal metabolic rate of the whole body. For
muscle m , A Q
) is the activation level of muscle m at a dis-
crete time step designated by t
m
( t
i
. The simulation interval
(i.e., 0.0) t )0.56) was broken into 172 segments, and
a static optimization problem was solved at each t
i
i
.
Activation dynamics was neglected.
To assess the importance of incorporating muscle
physiology, one static optimization problem was solved
neglecting the force}length}velocity properties of muscle,
and a second was solved with these properties included.
In both cases, at each discrete time step, t
m
is the activation heat rate, M Q
m
is the
maintenance heat rate, S Q
is the shortening heat rate, and
, the force
output of each muscle was computed from its activation
level, a
m
i
is the mechanical work performed (Anderson, 1999).
Hyper-extension of the joints was penalized by appen-
ding the following function to the performance criterion:
m
). In the `non-physiologicala case, each muscle
was assumed to be an ideal force generator
( t
m
i
/
( t
)" w
P
t &
A
1 +
j /1
¹
-*' j
( t )
2
B
d t ,
(8)
F
m
( t
i
)" a
m
( t
i
) F
o m
,
(11)
&
0
is
its maximum isometric force. In the `physiologicala case,
the force generated by a muscle was constrained by its
force}length}velocity properties:
( t
) is the force generated by muscle m and F
o m
i
( t ) is the torque applied by the ligaments
about joint axis j at time t , and w (0.001) is a weight
factor. The performance criterion was therefore aug-
mented
-*' j
F
( t
)" a
( t
) f ( F
o m
, l
m
, v
),
(12)
m
i
m
i
m
@ " P
t &
A
B Q # 5 +
m /0
B
( A Q
# M Q
# S Q
m
#=Q
)
d t
) is the general
force}length}velocity surface for muscle assumed in the
musculoskeletal model, and l
o m
, l
m
, v
m
m
m
m
0
# /
( t
).
X
( t
)! X
(0)
&
#.
&
#.
the
shortening velocity of muscle m (Zajac, 1989). Muscle
lengths and shortening velocities were computed from
the joint angular displacements and velocities predicted
by the dynamic optimization solution, with tendon com-
pliance and muscle pennation angle taken into account.
The force outputs of the muscles were constrained to
produce the muscular joint moments predicted by the
dynamic optimization solution
is the length and v
J
(9)
m
m
Thus, the dynamic optimization problem was to "nd
values of the controls (Eq. (1)) such that the constraints
(Eqs. (3)}(6)) were satis"ed and the performance criterion
(Eq. (9)) was minimized. Note that the dynamic equations
for the skeleton (not shown), for the muscle forces (not
shown), and for the muscle activations (Eq. (2)) were also
constraints in the dynamic problem and were enforced
implicitly by numerically integrating these equations for-
ward in time during the simulation. The dynamic optim-
ization problem was solved using a parallel parameter
optimization algorithm (Pandy et al., 1992; Anderson
et al., 1995).
5 +
m /1
F
m
( t
i
) r
m , j
( t
i
)"¹
j
( t
i
), j "7, 2 , 23, i "1, 2 , 172,
(13)
)is
the moment arm of muscle m about the j th joint axis, and
¹
m
( t
) is the force generated by muscle m , r
m , j
( t
i
2.3. Static optimization problems
) is the muscular joint moment acting about the j th
joint axis as predicted by the dynamic optimization solu-
tion. Note that Eq. (13) begins with j "7, the
#exion}extension axis of the back. Because the "rst six
degrees-of-freedom concerned the displacement and ori-
entation of the pelvis relative to the ground (see Fig. 1),
no muscular moments were exerted about these general-
ized coordinates.
The performance criterion was to minimize activation
squared, summed across all muscles (Kaufman et al.,
1991; Crowninshield and Brand, 1981):
j
( t
i
So that the static and dynamic solutions could be
compared more directly, the joint angular displacements,
joint angular velocities, and muscular joint moments
predicted by the dynamic optimization solution were
input into the static optimization problem, as though
these quantities came from experimental measures. In
this way, imprecision in the collection and processing of
experimental data and errors in the estimation of body
anthropometry were eliminated.
Following the example of Kaufman et al. (1991), in
order to account for the force}length}velocity properties
of each muscle, the muscle activation levels were used as
the controls in the static optimization problems:
J
i
" 5 +
m /1
( a
m
( t
i
))
2
, i "1, 2 , 172.
(14)
In the non-physiological case (Eq. (11)), note that
muscle activation is equal to muscle stress multiplied by
a
m
( t
i
), m "1, 2 , 54, i "1, 2 , 172,
(10)
where a
=Q
where F
m
where ¹
where the function f ( F
where F
i
F.C. Anderson, M.G. Pandy / Journal of Biomechanics 34 (2001) 153 } 161
157
some proportionality constant
a
m
"
F
F
0 m
" k
PCSA ,
F
m
(15)
/PCSA is muscle stress, PCSA is physiological
cross-sectional area, and k is a constant. In the physiolo-
gical case (Eq. (12)), muscle activation is approximately
equal to muscle stress multiplied by some propor-
tionality constant, provided a muscle operates on the #at
portion of its force-length curve and at small contraction
velocities:
a
"
F
m
) + k
m
PCSA
F
when l
+ l
0 m
and
m
f ( F
0 m
, l
, v
m
m
m
v
m
+0.0.
(16)
is the optimal muscle-"ber length.
Thus, the non-physiological static optimization prob-
lem was to "nd an activation level for each muscle in the
model such that the performance criterion (Eq. (14)) was
minimized, the joint torques predicted by the dynamic
optimization solution (Eq. (13)) were generated at each
joint, and each muscle was treated as an ideal force
generator (Eq. (11)). The physiological static optimiza-
tion problem was the same, except that each muscle was
constrained to act in accordance with its
force}length}velocity properties (Eq. (12)). Both static
optimization problems were solved using a gradient-
based sequential quadratic programming algorithm
(Pandy et al., 1992).
0 m
Fig. 3. The net muscle moments applied about the axes of the hip joint
are shown as predicted by the dynamic optimization solution (thick
black lines) and as measured by Crowninshield et al. (1978) (circles),
Cappozzo et al. (1975) (triangles), Patriarco et al. (1981) (crosses), and
Inman et al. (1981) (squares).
3. Results
The joint moments predicted by the dynamic optim-
ization solution (and used as input to the static optimiza-
tion problem) were similar to those reported in the
literature (Figs. 3 and 4) (Cappozzo et al., 1975; Crownin-
shield et al., 1978; Patriarco et al., 1981; Inman et al.,
1981). Joint moments about the axes of the back, subta-
lar, and metatarsal joints were also predicted by the
dynamic solution but are not reported here. Discontinui-
ties in the joint torques are present because the terminal
boundary constraints imposed in the dynamic optimiza-
tion solution were not precisely met.
In general, the forces predicted by the dynamic and
static solutions were very similar for many muscles (Figs
5 and 6: ILPSO, GMAXL, GMEDA, VAS, GAS, SOL,
TFL, BFSH, PFEV). However, there were two notable
exceptions. Following heelstrike, the dynamic solution
predicted signi"cant activity in the medial portion of
gluteus maximus, whereas the static solutions predicted
very little activity in this muscle (Fig. 5: GMAXM,
0}30%). On the other hand, the static solutions predicted
signi"cant activity in hamstrings, whereas the dynamic
Fig. 4. The net muscle moments applied about the axes of the knee and
ankle joints are shown as predicted by the dynamic optimization
solution (thick black lines) and as measured by Cappozzo et al. (1975)
(triangles) and Inman et al. (1981) (squares).
solution predicted much less activity in this muscle
(Fig. 5: HAMS, 0}30%).
The resultant joint contact forces at the hip, knee, and
ankle predicted by the dynamic and both static optimiza-
tion solutions were very similar to each other and some-
what similar to contact forces reported by Crowninshield
and Brand (1981) and Hardt (1978) (Fig. 7). For both the
where F
m
where l
207207430.002.png 207207430.003.png 207207430.004.png
Zgłoś jeśli naruszono regulamin