MOVEMENT AND THE INTACT HUMAN ELBOW JOINT IN

advertisement
MOVEMENT AND THE MECHANICAL PROPERTIES OF
THE INTACT HUMAN ELBOW JOINT
By Jeremy Malcolm Lanman
B.A.,
Yale University (1972)
SUBMITTED IN PARTIAL FULFILLMENT
OF THE REQUIREMENTS FOR THE
DEGREE OF DOCTOR OF
PHILOSOPHY
AT THE
MASSACHUSETTS INSTITUTE OF TECHNOLOGY
JUNE 1980
(
Massachusetts Institute of Technology, 1980
Signature of Author:
Department of Psychology
April 11, 1980
Certified by:
Emilio Bizzi
Thesis Supervisor
'1-in
Accepted by:
ARCHIVES
MASSACHUSETTS INSTITUTE
OF TECHNOLOGY
JUL 8 1980
LIBRARIES
'Chairman, Department Committee
on Graduate Students
MOVEMENT AND THE MECHANICAL PROPERTIES OF
THE INTACT HUMAN ELBOW JOINT'
by
Jeremy M. Lanman
Submitted to the Department of Psychology on April 11, 1980
in partial fulfillment of the requirements for the
degree of PhoD.
ABSTRACT
Recent studies of motor coordination have emphasized the importance of
the mechanical properties of muscle in both load compensation-and in the
control of movement. Teleologically, different mechanical characteristics
may be optimal for different tasks. For example, a skiier may wish to
adjust the stiffness of his legs in accordance with the terrain and snow
conditions. How, and to what extent can subjects control the mechanical
properties of their joints? To help answer this question, we investigated
short term changes in muscle properties by measuring the response of the
intact human elbow joint to mechanical perturbations. The resistance of the
intact joint to such displacements reflects contributions from the muscles
and passive tissues spanning the elbow joint. In order to explore the
effects of both muscular activation and movement on joint properties we
studied three conditions: (1) voluntary isometric contraction, (2) passive
movement, and (3) voluntary movement.
In the isometric condition, the subject maintained a constant arm position against various loads. In the passive movement condition the subject's
arm was moved at a constant velocity. Finally, the voluntary movement
condition combined muscular activation with movement. During all three
conditions a small amplitude, high frequency (15-30 Hz) sinusoidal perturbation was applied about the elbow joint. Joint impedance was derived from
the resistance of the elbow to this superimposed displacement. We obtained
continuous measures of joint impedance, joint angle, and the biceps and
triceps EMG's.
During isometric contraction, joint impedance increased monotonically
with increasing muscular activation (estimated by EMG). In contrast, during
passive arm movements joint impedance decreased markedly. This was true
even for very slow passive movements when no change in muscular activation
could be detected. Finally, voluntary movements resulted in changes in
joint impedance which were consistent with the combined effects of
muscular activation and movement.
We attribute these changes in joint impedance to variations in the
properties of the muscles acting about the joint. We conclude that length
TABLE OF CONTENTS
Page
ABSTRACT
2
ACKNOWLEDGEMENTS
5
INTRODUCTION
6
Muscle Properties
Joint Properties
Techniques for Measuring Joint Properties
METHODS
Error Analysis
Advantages of High Frequency Measurement
Testing and Calibration
Data Acquisition
Methods Summary
RESULTS
Voluntary Muscle Activation
Passive Movement
Voluntary Movement
Impedance Measurements at Different
Amplitudes and Frequencies
8
11
11
14
18
19
20
22
23
24
24
25
26
28
DISCUSSION
30
MUSCLE MODEL
35
Bond Number
Testing the Muscle Model
Summary of Muscle Model
Applying the Muscle Model to the Intact Joint
The Influence of Passive Tissues
Conclusions from the Muscle Model
36
38
39
40
41
42
CONCLUSION
43
REFERENCES
46
APPENDIX I
52
APPENDIX II
61
FIGURE LEGENDS
63
FIGURES
68
TABLES
changes (whether active or passive) and neural activation have opposing
effects on the mechanical properties of muscle. These results will be
discussed in relation to the well-characterized effects of length change
and activation on the mechanical properties of isolated muscle. Through an
understanding of the factors controlling muscle properties, we can begin to
understand how the nervous system might adjust the mechanical properties of
a joint.
Dr. Emilio Bizzi, Professor of Neurophysiology
Lanman, 5
ACKNOWLEDGEMENTS
The help and support of many people was required for the completion
of this work.
In particular I would like to thank my advisor, Dr. Emilio
Bizzi for his help and support.
I would also like to thank Dr. Parvati
Dev for her gentle guidance and maintained interest, Dr. Neville Hogan
for his many insightful comments, and Dr. Tom Zeffiro for his continual
encouragement and his good sportsmanship in subjecting himself to gooey
cornstarch.
John Swan and Ed Cruz must be thanked for achieving excellence
in constructing the apparatus.
Finally, I would like to thank Terry
Heyward for her cheerfulness while typing so many revisions of this
manuscript.
Lanman, 6
INTRODUCTION
The translation of patterns of neural activity into movement, or more
generally into behavior, is accomplished by the musculo-skeletal system.
This translation can be thought of as occurring in two steps:
(1) motor-
neuronal activity stimulates force generation in muscle and (2) this muscle
force acts through a mechanical linkage to generate torque about one or more
joints.
The second part of this problem can, at least in principle, be
modeled by applying the laws of classical mechanics to the anatomical
relationships of the muscles and joints.
The prior problem of translating
neuronal activity into muscle force is more difficult to understand and
model, and neurophysiologists have all too frequently glossed over this
problem, simply assuming that alpha motorneuronal activity stimulates a more
or less equivalent increase in muscle force.
The dangers of such an over-
simplification can perhaps be best understood by analogy to visual physiology:
suppose that investigators were to assume that retinal ganglion cells simply
reported absolute incident light intensity of one point on the retina, thus
ignoring the vast amount of preprocessing that occurs peripherally, on the
retina itself.
Such an assumption would clearly interfere with progress in
understanding visual physiology.
Ongoing research on isolated muscle preparations has begun to reveal
some of the principles, but also some of the complexities of muscle action.
A particularly striking demonstration of the importance of intrinsic muscle
properties has been seen in the flight muscles of certain insects (Pringle,
1967).
The wings of such a species may beat 120 times/sec while the motor
units fire at only three spikes/sec.
This oscillatory behavior has been
modeled as a non-neural feedback system between muscle and load, the central
Lanman, 7
feature of which is the enhancement of active tension generation by mechanical
stretch.
Clearly, a neurophysiologic investigation searching for a neuronal
oscillator generating the 120 cycle wing beat would be misdirected in such a
species; the oscillator is the muscle itself.
Although this extreme case of
a muscle oscillator is probably restricted to invertebrates, the enhancement
of muscle force generation by mechanical stretch has been repeatedly observed
in vertebrate preparations (for a recent study see Edman et al., 1978).
Could this be a kind of non-neural stretch reflex?
The point illustrated by
these examples is that muscle force is not simply and directly related to its
stimulation rate; the degree and rate of stretch is important as well.
This complex interaction of neural and mechanical factors in determining
muscle force is important in understanding central neural control of movement.
For example, much research (Bizzi et al., 1978; Feldman, 1966a, 1966b;
Grillner, 1972; Polit, 1977) has been devoted to elucidating the role of the
intrinsic length/tension properties of muscle in maintaining a stable posture.
The length/tension characteristics of muscle can be modeled, somewhat metaphorically, as a spring (Houk and Henneman, 1974) with parameters such as
stiffness and rest length that can be controlled by innervation.
In turn,
this spring-like property of muscle provides for a kind of non-neural load
compensation:
when an external load is applied to a muscle-spring, the muscle
is stretched and consequently provides more tension.
If a muscle-spring is
very stiff, i.e., if tension increases rapidly for a small increase in
length, then the intrinsic properties of muscle will be very effective in
load compensation.
Recent measurements indicate that this sort of intrinsic
muscle-spring type of load compensation is more important than any such
compensation mediated by reflex pathways (Bizzi et al., 1978; Grillner, 1972;
Lanman, 8
Merton, 1953).
Thus, the mechanical properties of muscle are important in
understanding control of movement.
Unfortunately, however, there is a lack
of data on muscle and joint properties under natural conditions,
The goal of
this study is to measure the mechanical properties of the intact human elbow
joint under natural conditions of innervation and movement.
Since the
properties of a joint are in large part determined by the muscles, we will
first review the main features of the structure and mechanical properties of
muscle.
Muscle Properties
In order to provide a conceptual framework in which to understand the
properties of muscle, it is important to review what is known about the
structure and functional properties of muscle.
Studies of the structure of
muscle have revealed that muscle is made up of groups of muscle fibers which
are in turn made up of repeated parallel arrays of interdigitating actin and
myosin filaments (Fig. 1).
The filaments themselves are quite rigid and
inextensible (Huxley and Simmons, 1971), so when muscle is stretched or
shortened the filaments must slide past one another, thereby changing their
degree of interdigitation or overlap.
Tension in muscle is generated by
molecular bond bridges between the two types of filaments.
Since bonds
can only be formed between adjacent overlapping portions of the actin and
myosin filaments, tetanic muscle tension varies as a function of overlap, and
therefore also as a function of muscle length (Gordon et al., 1966).
In this
way the molecular structure of muscle can give insight into the static length/
tension properties of muscle discussed in the previous paragraph.
Muscle tension depends not only on the degree of overlap of filaments,
but also on the proportion of overlapping bonds that are actually attached
Lanman, 9
and pulling, generating tension.
Neural excitation of muscle stimulates
increased bond formation, by increasing the availability of Ca++ to the
troponintropomyosin system (Murray and Webber, 1974).
excitation contraction coupling.
This is the basis of
When muscle stimulation is stopped, bond
formation stops and tension drops as bonds break.
Under rigorously isometric
conditions the fall in tension can be surprisingly slow (seconds), indicating
that spontaneous bond breakage is quite slow (Huxley and Simmons, 1972; Webber
and Murray, 1973).
When muscle length is allowed to change, however, there
is good evidence that bonds are broken much more rapidly.
This is reasonable
since as muscle length changes, the actin and myosin filaments slide past one
another.
Since the bond bridges cannot stretch indefinitely, they must
eventually break loose.
Thus neural stimulation of muscle increases bond
formation and therefore produces a rise in muscle tension, while mechanical
shortening or lengthening results in bond breakage and a decrease in muscle
tension (Joyce et al., 1969).
The balance between these two opposing
factors, electro-chemical activation and mechanical disruption, may
explain the rather complex dependence of muscle force on the velocity of
contraction or stretch, i.e., the force/velocity relationship of muscle.
This relationship is of direct significance in understanding neural control
of movement, as opposed to posture.
Having briefly sketched the molecular mechanisms thought to underlie the
length/tension and the force/velocity relationships of muscle, it is now
important to consider how muscle responds to rapid stretch.
Figure 2 from
Rack and Westbury (1969) shows the static length/tension characteristics of
the cat soleus muscle at different rates of stimulation (dotted lines).
Super-
imposed on these static characteristics are a series of solid lines showing the
Lanman, 10
response of muscle to dynamic (ramp) stretch.
Note that the dynamic charac-
teristics are much steeper than the isometric length/tension properties.
Muscle resists small rapid stretches much more strongly than expected on the
basis of static characteristics, and this high resistance is called shortrange stiffness (SRS).
Thus the mechanical properties of muscle appear
differently, depending on how far and how fast the muscle is stretched.
The static length/tension characteristics of muscle appear to fit
molecular structural data quite well, and structural analysis of muscle can
contribute to our understanding of the SRS phenomenon as well.
When muscle
is rapidly and incrementally stretched, the actin and myosin filaments are
slightly displaced, thus stretching any molecular bridges between the filaments.
Since each individual bond bridge resists this stretch, the greater
the number of bonds the greater this resistance will be.
If bond bridges
generate overall muscle tension as well as SRS, then both of these parameters
should increase together as "bonds" are formed.
This hypothesis has been
repeatedly verified not only under steady-state conditions (Ford et al., 1977;
Morgan, 1977; Rack and Westbury, 1974), but also during isotonic shortening
(Julian and Sollins, 1975), and during changes in stimulation rate (Mason,
1978).
In all of these diverse paradigms resistance to incremental stretch,
or SRS, was directly proportional to overall muscle tension, indicating that
each bond contributes a constant ratio of both tension and stiffness to whole
muscle.
To summarize, the resistance of muscle to small rapid stretch reflects
the stiffness of individual bond bridges, while the changes in muscle tension
seen with large slow displacements reflect (among other factors) the changes
in overlap between actin and myosin filaments.
The mechanical properties of
Lanman, 11
muscle applicable during voluntary movement probably reflect elements of both
the static length/tension characteristics and the more "dynamic" SRS
properties.
Joint Properties
In the previous section we discussed the structure and mechanical
properties of skeletal muscle.
Because of the linkage of muscles about the
It is
joint, these muscle properties contribute to total joint properties.
total joint characteristics that determine how our limbs respond to
environmental loads.
As discussed earlier, muscle properties vary with level of activation
and with extent and rate of stretch.
Consequently, we expect joint properties
to vary with electromyographic (EMG) activity and with joint rotation.
The
goal of this study is to measure the mechanical properties of the intact human
elbow joint with natural innervation.
We will look specifically for (1) the
effects of EMG activation on joint properties, (2) the effects of passive
movement on joint properties and (3) the variations in joint properties
during voluntary movements.
Previously, intact joint properties have only
been investigated under isometric conditions.
Before going into the
details of this study, we will briefly discuss the history and techniques
for measuring joint properties.
Techniques for Measuring Joint Properties
A wide variety of different types of perturbations have been used to
probe the mechanical properties of joint.
These include torque steps,
gaussian noise and continuous sinusoids (for recent examples see:
Bizzi et al., 1978; gaussian noise:
uous sinusoids:
steps:
Agarwal and Gottlieb, 1977b; and contin-
Agarwal and Gottlieb, 1977a; Joyce et al., 1974).
If muscle
Lanman, 12
and joint could be treated as a simple linear system, all these different
methods would give the same results; however, in fact the result differ.
This is not surprising when you consider the complexity of the system.
For
example, recall the very different response of muscle seen under static
versus dynamic conditions (isometric length/tension characteristics versus
short range stiffness).
The resistance of the intact joint to rapid small-
amplitude perturbation should reflect the SRS of muscle, while the response
to maintained displacement should depend primarily on the isometric length/
tension properties.
Because of these differences, when reporting the
mechanical properties of muscle and/or joint, it is important to specify how
these properties are measured, i.e., what waveshape, amplitude and rate of
perturbation was used for measurement.
In this study, small amplitude,
high frequency sinusoidal disturbances were applied to the forearm.
The
resistance of the joint to these perturbations gives a measure of the
mechanical properties of the joint.
This approach was chosen for two reasons.
First, by using continuous sinusoidal perturbations, impedance can be monitored continuously and any changes in mechanical properties can be detected as
they happen.
Second, by using high frequency disturbances, joint properties
can be monitored during natural voluntary movements.
This is done by using
filters to separate out the relatively low frequency components of voluntary
movement from the higher frequency superimposed perturbations.
Thus this
method of using continuous sinusoids to measure joint properties is particularly well suited to detecting changes in the mechanical properties of the
intact joint during different conditions of posture and movement,
Thus far we have discussed joint properties in relation to intrinsic
muscle properties, yet measurements on the intact joint may be influenced by
Lanman, 13
reflex activity as well.
Previous investigators have seen effects of reflex
activity on joint mechanical properties (Agarwal and Gottlieb, 1977a; Goodwin
et al., 1978; Joyce et al., 1974), but such effects were only seen at
measurement frequencies below about 15 Hz.
This is because of the low-pass
filter characteristics of muscle (Bawa et al., 1976a; Goodwin et al., 1978).
Muscle simply cannot contract and relax quickly enough to generate bursts of
force as fast as 15 times a second.
Most of the measurements reported here
are taken at frequencies above 15 Hz and, therefore, reflect intrinsic
muscle properties rather than reflex modulation.
To summarize, by using small amplitude, high frequency perturbations,
it is possible to measure changes in the mechanical properties of a joint,
These measurements can be made under different conditions of load and movement, with a minimum of interference with voluntary arm movements.
The
results of these measurements can be compared to similar measurements made
on isolated muscle.
Lanman, 14
METHODS
The elbow joint was chosen for this study because of its relative anatomical simplicity.
A torque motor was used to apply accurately known sinusoidally
varying torques to the forearm, and the resulting acceleration was monitored.
After allowing for inertial effects, the resistance of the elbow joint to this
perturbation can be measured and represented on a phasor diagram as a combination of two components:
an in-phase component, or an "equivalent stiffness"
(K ), and a quadrature component, or an "equivalent viscosity" (B ). These two
components of resistance to imposed movement can be measured continuously both
during posture and during movement.
This will be done in order to detect
changes in the mechanical properties of the elbow joint.
In order to perform these measurements, the forearm was attached to the
shaft of a powerful torque motor with an individually fitted fiberglass cast
(Fig. 3).
The torque motor was anchored to a concrete footing and the upper
arm was held steady both by a brace and by the large inertia of the upper body.
Thus, the only movement was a rotation of the forearm in the horizontal plane,
corresponding to flexion or extension of the elbow joint.
Sinusoidal torque
disturbances (3 to 30 Hz) were applied to the forearm by the torque motor,
and this torque was monitored by a strain gauge bridge.
The muscles and
tissues of the elbow also transmitted some torque to the forearm.
the torque on the shaft ( s) and the torque in the elbow joint (j)
The sum of
caused
the forearm to accelerate:
(1)
0e=T
s
+ t. .
I = moment of inertia of arm/cast
.. system
e = angular acceleration of
arm/cast
Ts
= measured torque on shaft
rj = torque transmitted through
elbow joint
Lanman, 15
The torque of the elbow joint cannot be directly measured, but it can be
computed by rearranging equation (1):
(2)
. = I
- T
When the forearm was driven back and forth with a sinusoidal acceleration (O),
the joint torque (rj) represents the resistance of the muscles and tissues of
the joint.
In order to remove non-sinusoidal components, such as contribu-
tions from voluntary muscular contractions, both torque and acceleration
measurements were narrowband filtered
(m)0
at the frequency of torque motor drive
The filtered signals are denoted by subscripts (T
'I
).. The output
of the filters can be described as sinusoids with slowly varying amplitude
t = f
and phase:
O = A sin (wt)
t
TjW = T sin (wt + #)
+
I
frequency of torque motor drive
e = filtered instantaneous angular
W acceleration of forearm
W=
A = peak amplitude of Oe
rJ = filtered instantaneous torque,
T , as defined on previous page
T = peak amplitude of Tj
4 = phase difference between Tj
and O
t = time
1 Bandwidth
of 10 Hz.
Lanman, 16
can be used to compute the equivalent stiff-
The resistance of the elbow (.j)
ness and equivalent viscosity of the joint.
This computation can be best
represented on a phasor diagram:
velocity axis
T
quadrature component
T sin
displacement axis
in-phase component = T cos f
2
(3) KW
T o Acos #
(4)
T w sin 4
A
BW
The equivalent stiffness and viscosity mimic the response of the joint under
the conditions of measurement.
Since it is anticipated that equivalent
stiffness and viscosity will depend on the frequency of measurement, these
values are subscripted:
K and B .
The amplitude of perturbation used were
necessarily always small (limited by the power of the torque motor, see
Table II).
There was no significant difference in measurements made at the
different amplitudes tested.
To further illustrate the relationship between sinusoidally varying
resistance and the mechanical properties of a system, a simple linear system
will be given as an example.
Example:
Let us consider a hypothetical system consisting of ideal
inertial, viscous and elastic elements, M, B, and K,
respectively. When this system is perturbed by a sinusoidal
acceleration (*) it resists with a force (f):
Lanman, 17
B
M
f, x
x = X sin wt
=X
K
measured
here
sin (wt + 900)
X sin (wt + 1800)
2
f = Mx + Bx + Kx
The instantaneous driving force (f) can be represented as the sum of
sinusoidal components, each with a constant amplitude and phase:
f = [M] X sin (wt) +
inertial component
[B] X sin (wt + 900) +
viscous component
[K] X2- sin (wt + 1800)
elastic component
The addition of these sinusoidal components can be represented
vectorally on a phasor diagram. In such a diagram a sinusoid is
represented by a vector; the length of the vector represents the
amplitude and the direction of the vector represents the phase
angle. (Throughout this study, phase angle is measured with
respect to the acceleration.) The velocity ( ) is shifted 900
with respect to acceleration, while the displacement (x) is 1800,
i.e., opposite to the acceleration.
MX
I velocity axis
BX/w
acceleration
axis
F
I
2 ---
displacement axis
KX/w
f = Mx + Bx + Kx
= F sin (Wt +
)
This diagram shows graphically the relationship between the applied
force (F) and the forces of the mechanical elements, K, B and M. ..
K and B can be computed from the applied force, the acceleration (X)
and the mass:
Lanman, 18
K
M 2 - F w Xcos
B= f
B=
sin
x 4
The phase angle between the applied force and acceleration ( ) allows
the total force to be divided into in-phase and quadrature components
which reflect stiffness and viscosity, respectively.
Error Analysis
The method outlined above is subject to some constant or slowly varying
errors, and for this reason only short term changes in joint properties will
be reported.
Such changes could be readily detected during muscle activation
and passive joint rotation.
The pattern of changes seen was quite repeatable
from day to day and subject to subject (three subjects were used).
Thus,
changes in joint impedance can be reliably detected even though the absolute
values of both K and B are not certain.
In order to better understand the nature of these errors, one must
consider the computations involved.
joint torque (Tj).
Joint properties are determined from
This quantity can not be directly measured and so must be
computed by subtraction (see equation #2).
At higher frequencies, this
difference is relatively small and so even small measurement errors may be
significant.
Figure 4 shows how small errors in amplitude or phase measure-
ments can effect the computation of equivalent stiffness and viscosity.
Errors of the magnitude shown are likely, due to inevitable errors in calibrating the transducers; however, these errors are constant or slowly varying.
Furthermore, the inertia of the arm-cast system is constant.2
Therefore, any
changes in the torque and acceleration measurements must indicate a change in
the mechanical properties of the elbow joint.
To show that measurements of changes are independent of any of these
Lanman, 19
constant errors, errors were deliberately introduced.
weights to the arm/cast system.
This was done by adding
As expected, such mass loads shifted the
position of the in-phase impedance (K ) trace, but had no effect on the
pattern of changes seen under different conditions of load and movement.
Advantages of High Frequency Measurement
The difficulties encountered when using high frequency perturbations to
measure joint impedance were pointed out in the section on Error Analysis.
In essence, the problem is that at high frequencies the large inertial
reaction forces mask the much smaller forces generated by the muscles and
passive tissues spanning the elbow joint.
All the same, high frequency
measurements have compensatory advantages that make it worthwhile to overcome
these difficulties,
First, during voluntary movements impedance can only be
measured at high frequency.
The ability to measure impedance during active
movements depends on the use of narrowband filters to separate out accelerations due to voluntary muscular activation (relatively low frequency) from
accelerations due to the high frequency perturbation of the torque motor:
2To assure that the inertia of the arm/cast system could not change, the
forearm was fixed in an individually fitted fiberglass cast and the residual
spaces between arm and cast were filled with a viscous solution of cornstarch and water. The system was sealed so the arm could not be pushed into
or withdrawn from the cast without creating a bubble. Even if the forearm
should move slightly inside the cast, an equivalent volumn (mass) of solution
would be displaced, thus conserving the total moment of inertia of the arm/
cast system. Only the appearance of a bubble (larger than 10 ml at the tip
of the cast) could alter the moment of inertia enough to significantly
influence the computation of K and B . Such bubbles were not observed.
Since the inertia of the arm/cast system could not change, any change in
measured torque and acceleration must reflect changes in the muscles and
passive tissues spanning the joint.
Lanman, 20
superimposed oscillation
narrowband filter
voluntary movement with
superimposed sinusoidal
perturbation
low pass filter
voluntary movement
(see Fig. 5 for a demonstration of the effectiveness of this separation.)
It
is not possible to separate a low frequency oscillation from voluntary
movement.
There is another advantage in using high frequency disturbances to probe
joint properties.
When the elbow joint is oscillated, even at high
frequencies, reflex modulation of EMG activity can be observed (Rack and Ross,
1974).
However, because of the low pass properties of muscle, very little
high frequency modulation of muscle force is produced.
When muscle is
stimulated with say 30 Hz bursts of activation, a rather well fused tetanus
is observed with only a few percent modulation at the drive frequency
et al., 1976a, 1976b; Goodwin et al., 1978).
(Bawa
Therefore, any 30 Hz modulation
measured in joint torque cannot be attributed to reflex modulation.
High
frequency modulation of torque must reflect the intrinsic mechanical
properties of muscle itself.
Testing and Calibration
Before making any physiologic measurements the system was extensively
tested with artificial loads.
To do this the arm/cast was filled with water,
thus constituting a purely inertial load.
This arrangement is intended to be
Lanman, 21
equivalent to the system with the arm in the cast except that there can be
Any departure of the measured joint torque (Tj) from zero
no joint torque.
must be a measurement error:
= applied torque measured on
S-
s -est
e
= error
=
..
motor shaft
e = measured acceleration
Iest = estimated moment of inertia
By adjusting Iest it was possible to reduce this error to less than 0.1%, but
only after eliminating all extraneous movement and vibration.
This was done
by (1) anchoring the torque motor body to a concrete footing, (2) counterbalancing the one-sided inertia of the cast, and (3) using a very rigid
(1" diameter) shaft between the motor and the cast.
With these modifications
it was possible to achieve small errors at all frequencies and at all shaft
angles.
Using this same water filled cast arrangement, the transient response of
the system was tested.
This was done by attaching a small steel weight to the
arm/cast with an electromagnet.
When the magnet was shut off the weight
dropped, thus creating an instantaneous change in load.
The in-phase
impedance trace (K ) responded to this step change with an exponential (time
constant - 100 msec) (see Fig. 6).
Along with showing the transient response
of the system, this figure shows how the apparatus can be calibrated.
If the
moment of inertia (mass times radius squared) of the dropped weight is known,
then the change in the computed value of K can be calibrated in Nt-m/rad:
AK
=
I w2
This technique of adding known inertial loads was in fact used to calibrate
the entire system at the start of each experimental run.
Lanman, 22
Data Acquisition
The resistance of the elbow joint to sinusoidal acceleration was
recorded along with:
(1) the surface EMG over the biceps and triceps muscles
(2) the angular position and velocity of the elbow joint
(3) the magnitude of any steady bias force load on the forearm.
These measurements were made using standard techniques and were reccrded in
real time on a Honeywell visicorder and on magnetic tape for subsequent
computer analysis.
Data were collected under three conditions:
(1) Voluntary muscle activation:
In this condition the subject was
instructed to maintain a constant arm position while holding bias
loads ranging from 2.5 to 40 ft-lbs.
These loads were generated by
suspending lead weights, with a lever and pulley linkage, from the
motor shaft.
(2) Passive movement:
By using the torque motor as a position servo,
it was possible to rotate the arm passively (i.e., without any
voluntary muscular activation or EMG activity).
Position ramps
corresponding to velocities of .5 to 10 degrees per second were
used.
(3) Voluntary movement:
As discussed earlier, in the high frequency
range it is possible to measure variation in joint properties during
active voluntary movements.
Such movements were studied with and
without an additional viscous load.
The effect of the viscous load
was to oppose movement velocity so that greater muscle force (and
greater EMG activity) was required to achieve the same velocity of
movement.
This viscous load was generated by feeding the output of
Lanman, 23
a tachometer (velocity transducer) into the torque motor driver.
Methods Summary
When the elbow joint is driven with a small amplitude sinusoidal torque
it responds with very nearly sinusoidal acceleration.
The relationship
between this torque and acceleration tells us something about the mechanical
properties of the joint.
For example:
if the muscles about the elbow become
stiffer, or more resistant to movement for any reason, then the arm will move
less, i.e., accelerate less for a given amplitude of torque drive.
relationship is important as well.
For example:
The phase
if the joint should resist
movement velocity in a viscous-like fashion, then the resisting force would
be in-phase with movement velocity, while an elastic type of resistance to
displacement would result in a force in-phase with the joint displacement.
The sum of viscous and elastic-like forces gives a net force with a phase
angle intermediate between the velocity and displacement, i.e., between 900
and 1800.
Thus, to describe and quantify the resistance of the elbow joint to
movement the amplitude and phase relationships between torque and acceleration
will be measured over a range of amplitudes and frequencies.
These measure-
ments will be made during posture maintained against various bias loads and
also during active and passive movements.
An important advantage of this
technique of impedance measurement is that the measurement does not interfere
with voluntary movement.
Because of the small amplitude of the perturbations
used, measurements can be made under essentially isometric conditions, or
during the trajectory of a voluntary movement.
Lanman, 24
RESULTS
We measured the resistance presented by the muscles and passive tissues
spanning the elbow joint in response to imposed sinusoidal oscillation
(3-30 Hz).
Measurements were made under three conditions (see Methods):
(1) voluntary muscle activation (isometric)
(2)
passive movement
(3) voluntary movement.
The resistance of the total elbow joint (joint impedance) increased with
muscle activation and decreased during passive movement.
The third condition
of voluntary movement is a combination of the first two conditions.
Impedance
changes measured during voluntary movements reflect a combination of effects
due to muscle activation and to movement.
Voluntary Muscle Activation
Isometric muscle activation was achieved by asking the subject to hold
a constant arm position while various steady torque loads were applied
(2.5 to 40 ft-lbs).
EMG measurements showed the increasing muscle activation
required to maintain these loads.
A sinusoidal perturbation was superimposed
on the steady arm position in order to measure the equivalent stiffness and
viscosity of the elbow joint.
Such measurements showed that joint impedance
(both K and B ) increased with increasing bias load (Fig. 7).
Furthermore,
when the subject co-activated maximally, isometrically activating both flexor
and extensor muscles simultaneously,. then impedance increased still further.
This increase in equivalent stiffness (K ) could be as large as several
hundren N-m/rad, hundreds of times stiffer than passive tissues.
muscles can account for such a large increase in joint stiffness.
Only the
When
muscle is activated more cross bridges are formed between the filaments and
Lanman, 25
consequently stiffness increases greatly.
Theoretically, muscle stiffness should increase linearly with increasing
muscle force; however, our measurements of joint stiffness level off at
higher torques (Fig. 8).
This saturation effect may be due to the compliance
of the tendons in series with the muscles (see Discussion).
If so, our
results imply that the maximum stiffness of the flexor muscle tendons must
be about a few hundred Nt-m/rad,
The measurements reported above were all made with the elbow joint at
about 900, intermediate between flexion and extension.
Joint impedance did
not change significantly when measured at different joint angles, except in
extreme flexion or extension.
In these extreme positions impedance rose
markedly.
Passive Movement
Next, impedance measurements were made during passive arm movements.
This was done to test if movement, without any change in EMG activity, could
influence joint properties.
position servo.
To do this the torque motor was used as a
When the forearm was passively rotated by the servo, joint
impedance decreased (Fig. 9).
This was true even for very slow movements
(> 30 /sec), movements which were much too slow to elicit any stretch reflex.
The decrease in joint impedance depends on the velocity of joint rotation
as shown in Figure 10.
The stiffness decrease was the same for either
direction of joint rotation.
This effect of movement on joint properties was initially puzzling;
however, the molecular structure of muscle appears to give a reasonable
explanation.
When the forearm moves, the muscles change length and actin
and myosin filaments slide past one and another.
Any bond bridges between
Lanman, 26
the filaments will be pulled off by this movement and, consequently, muscle
stiffness will drop.
Such a decrease in stiffness as a result of movement
has been observed in isolated muscle studies (Julian and Sollins, 1975; Moss
et al., 1976), and there is every reason to expect such a stiffness decrease
in the intact joint as well.
A model of how joint stiffness should vary with
movement velocity will be presented in the Discussion.
Getting back to the data, there is significant scatter in the results
from different subjects.
tone.
This could be due to differences in resting muscle
Subjects with higher muscle tone will have a higher stiffness with
the arm stationary and thus stiffness can decrease more when the arm moves.
Vol untary Movement
We have seen that joint impedance increases with muscle activation but
decreases with passive movement.
What happens when activation and movement
are combined during a voluntary movement?
Figure 11 is a comparison of
impedance measurements made during three movements with similar trajectories
but with different levels of muscle activation:
(1) On the left is a passive movement driven by the position servo.
There is no detectable EMG activity.
Stiffness (K ) decreased
during this movement, as with any passive movement.
(2) In the center is a similar amplitude voluntary movement.
This
movement differs from the previous one by a burst of EMG activity.
During this voluntary movement stiffness again decreased, but
during the period of EMG activity stiffness recovered partially.
(3) Finally, on the right is a voluntary movement working against a
viscous load.
This movement required a large burst of EMG activity
to overcome the load.
During this burst of EMG activity stiffness
Lanman, 27
increased markedly.
The repeatability of these measurements is shown in Figure 12.
For all these
movements, the equivalent viscosity changes were similar to, but much smaller
than the changes in equivalent stiffness.
For this reason the ensuing
discussion will focus on stiffness changes, although the same arguments
should hold for the viscosity changes as well.
For all three of these movements the pattern of stiffness changes can be
understood as a simple combination of a decrease in stiffness due to movement
with a superimposed increase related to muscle activation.
The larger the burst
of EMG activity, the larger the superimposed increase in stiffness (Fig. 13).
The results shown in Figure 11 can be compared to results previously
found for passive movement and for isometric muscle activation.
The left hand
figure is simply another passive movement, directly comparable to the passive
In the other two records there is some EMG
movements previously discussed.
activity during the movement.
The effect of this muscle activation on stiff-
ness measurements during movement can be directly compared to the increase in
stiffness seen with isometric muscle activation.
For example, during the
loaded movement the muscles generated some 5 ft-lbs of torque (see torque trace
in Fig. 11C) and joint stiffness increased by as much as 75 Nt-m/rad.
Under
isometric conditions with the muscles holding a 5 ft-lb load the same increase
in stiffness was seen (Fig. 15).
Apparently, stiffness increases with torque
in much the same way either with the arm stationary or moving.
The difference
is that when the arm is moving the increase is superimposed on the stiffness
decrease due to the movement itself.
The total stiffness change is the sum of
the increase due to muscle activation and the decrease due to movement.
For all three movements stiffness initially decreases.
This seems
Lanman, 28
counterintuitive, how can stiffness decrease start the arm moving?
To under-
stand this, let us look closely at the sequence of events in Figure 14.
Initially, the arm is held in extension by a low level of EMG activity in the
triceps muscle.
This EMG activity decreases slightly (just before one second)
and shortly thereafter the arm starts moving slowly, at about 2 0 /sec.
At this
point in time the arm is drifting towards the primary or equilibrium position,
with little or no EMG activity (Bizzi and Polit, 1978).
During this period of
arm drift stiffness decreases, as expected for a passive movement.
The next
change is the large burst of EMG activity in the agonist (biceps) muscle.
During this burst of EMG activity stiffness increases markedly, as expected
for a muscle actively generating tension.
into two phases:
Thus the movement may be divided
(1) an initial slow drift initiated by a decrease in antago-
nist activity followed by (2) a second fast phase of movement initiated by a
large burst of agonist activity.
The movement is initiated by a decrease in
antagonist EMG, and this accounts for the initial decrease in joint stiffness.
Impedance Measurements at Different Amplitudes and Frequencies
The amplitude and frequency of perturbation used to measure joint
impedance was systematically varied.
Extensive measurements were made at
frequencies of 15, 20, 25 and 30 Hz.
The amplitude of the perturbation was
always less than one degree.
(The actual amplitude range used at each
frequency is shown in Table I. The maximum amplitude is limited by the
power of the torque motor and the minimum is limited by noise.)
Within these
limits there was no consistent dependence of impedance on amplitude of perturbation.
Even the largest amplitude perturbations were not sufficient to
cause muscle stiffness to yield.
Measurements made at different frequencies were also quite similar.
The
Lanman, 29
patterns of impedance changes reported in previous paragraphs were seen at
all frequencies and amplitudes studied; however, there were some quantitative
differences.
At higher frequencies equivalent stiffness (K ) changed more
and viscosity (B ) changed less than equivalent measurements made at lower
frequencies (Fig.
15A, B).
The quantitative differences in impedance measurements made at different
frequencies is consistent with the idea that joint impedance reflects the
short range stiffness of muscle,
In fact, measurements on isolated muscle
show that muscle impedance varies with measurement frequency in the same
However, the intact
general way as seen in the intact joint (Kawai, 1978).
human elbow joint is a rather awkward system to study these effects and
probably not too much should be made of the small differences in impedance
seen at frequencies between 15 and 30 Hz.
Finally, a few measurements were made in the low frequency range
(3-12 Hz).
Such measurements are difficult to interpret because of possible
reflex effects.
All the same, these lower frequency measurements did show
the major effects reported for high frequency measurements.
Even with 3 Hz
perturbations, equivalent stiffness and viscosity decreased during passive
movement and increased with isometric muscle activation.
Lanman, 30
DISCUSSION
Measurements of joint impedance (resistance to sinusoidal stretch) show
an increase during voluntary muscle activation and a decrease during passive
movement.
A third condition of voluntary movement combines muscle
activation with movement.
During voluntary movements impedance measurements
reflect the combined effects of an increase due to muscle activation and a
decrease due to movement,
These measurements of total joint impedance must
reflect, at least in part, the properties of the muscles acting about the
elbow joint.
In this section we will discuss the relationship between the
properties of muscle and the intact elbow joint
neural activation
stimulation
r = moment arm
length = X
joint angle
r
'-
=
e
force = F
external torque
=T
(1) AX = r sin Ae
(2)
T =rF
In order to compare joint properties to muscle properties, several
mechanical relationships must be derived.
The bones of the forearm act as
a lever so that when the arm flexes through some angle, the flexor muscles
shorten and the extensors lengthen.
The quantitative relationship
Lanman, 31
between change in angle (AO) and change in muscle length (AX) is given in
equation #1. For simplicity, this equation is an approximation.
For large
angles (extreme flexion or extension), a more complex relation would be
A further consequence of the lever
necessary (Lestienne and Pertuzon, 1973).
like arrangement of the joint is that the force in a muscle, say the biceps,
contributes to the total joint torque in proportion to its moment arm (the
moment arm is essentially determined by the radius of insertion of the
muscle (see equation #2)).
The total muscular torque about the joint is the
sum of the contributions from each muscle:
T = E ri fi.
Using these
relationships, it is possible to convert from muscle length and force to
joint angle and torque, i.e., from a linear to an angular coordinate system.
Given these mechanical relationships between muscle and joint, what
comparisons can be made between measurements on muscle and joint?
As a basis
for comparison, a quantitative relationship has been found between muscle
SRS (short range stiffness) and steady muscle force (Schoneberg et al., 1974):
(3) SRS = F1
L
E
SRS = stiffness
F = steady muscle force
L = muscle rest length
e = empirical constant
Theoretically, this relationship should hold even for human muscle, and the
value of the constant of proportionality (E) computed from this study can
be compared to values of e measured in isolated muscle.
To understand
why the value of e should be similar for all vertebrates, consider the
structural basis of the stiffness/force relationship, i.e., the bond
bridge.
The more bonds that are formed (per half sarcomere) the greater
the net muscle force.
In addition, each bond bridge is stiff, like a small
spring, and therefore resists displacement.
The greater the number of bond
Lanman, 32
bridges the greater the net stiffness of the muscle.
Thus, one expects
muscle force and muscle stiffness to vary together, each depending on the
number of bond bridges.
This relationship is for short range stiffness
measurements because any larger changes in length (greater than about 100A
per 1/2 sarcomere) would severely disrupt the bond bridges.
The constant of
proportionality (e) reflects the ratio of force to stiffness contributed by a
single bond bridge between the actin and myosin filaments.
Since the
structures of actin and myosin are very similar across vertebrates (Lazarides
and Rivel, 1979; Walker and Schrodt, 1974) one would expect e to be very
nearly constant between species.
Table II is a comparison of measurements of E made in widely different
preparations.
The principle source of variability in these values appears to
be the speed (interval or frequency) of measurement.
is stretched, the stiffer it appears.
The faster the muscle
Once an allowance is made for speed
of measurement, the different studies give remarkably consistent estimates
of e.
For comparison to the intact human arm, Rack and Westbury's (1974)
data is useful because it was collected from intact mammalian muscle, under
relatively natural conditions of stimulation, at body temperature.
Their
measurements of the ratio of stiffness to force (pp. 340) give:
1 = SRS L
F
=
(.6 to 1) Nt/mm 30 mm
Nt
= 3% to 5%
Unfortunately, this data was collected using ramp stretch of muscle, while
the data on the human elbow is for sinusoidal stretch.
It is not clear which
single frequency of sinusoid corresponds best to ramp stretch.
All the same,
this is probably the best isolated muscle data to use for comparison to
Lanman, 33
intact joint.
Can this quantitative relationship between muscle stiffness and force
account for measurements on the intact joint?
To make this comparison, it is
first necessary to convert to angular coordinates:
SRS =1
Le
Kangular
=
F
rL Tm
(muscle coordinates)
(joint coordinates)
Substituting into this equation Rack and Westbury's (1974) value for e, the
rest length (L)and moment arm (r)of the biceps muscle (Lestienne and
Pertuzon, 1973) gives a prediction for how fast joint stiffness should increase with increasing bias load:
(4)
Kangular
4 cm
(0.05 to .03) 12 cm
= (6.6 to 1.1) T
T = Nt-m
angular
Nt-m
rad
The angular stiffness of the elbow joint, measured at 15 Hz, does in fact
increase by about 10 Nt-m/rad for each Nt-m increase in bias load.
Figure 8
shows the stiffness/torque relationship as measured (@15 Hz) and as predicted
(equation #4).
Sources of error in this comparison include:
(1)extrapolat-
ing from Rack and Westbury's (1974) ramp stretch to sinusoidal stretch,
(2)considering only the biceps muscle when there are actually three major
flexors each with slightly different lengths and insertions and (3)the
values of (L) and (r)can only be roughtly estimated in the intact preparation.
Because of these approximations only a rough comparison can be made between
isolated muscle and intact joint measurements.
All the same, it is hard to
imagine any non-muscle tissue being stiff enough to account for these
Lanman, 34
measurements.
(Hayes and Haytze (1977) estimated passive tissue stiffness to
be only 1 to 2 Nt-m/rado)
This stiffness/torque relationship can also be used to evaluate the
stiffness decrease seen during passive movement (Fig.
10).
If we attribute
this stiffness decrease to the yielding of the elbow muscles, then it follows
that at rest these muscles must have a total stiffness of at least 50 Nt-m/rad.
According to the stiffness/torque relationship (equation #4), this muscle
stiffness corresponds to a resting muscle tone of some 5 to 7 ft-lbs or
about 3 ft-lbs in the flexors and 3 ft-lbs in the extensors.
few percent of the maximum torque of the elbow joint.
This is only a
In other words, we
hypothesize that the stiffness decrease seen during passive movement is due
to disruption of bond bridges.
This hypothesis leads to an estimate of
resting muscle tone which, although very approximate, is of the correct
magnitude (Moss et al., 1976).
In conclusion, the resistance of the intact human elbow to small
amplitude high frequency perturbations appears to reflect the SRS of muscle
rather than the resistance of passive tissues.
Next, a model will be
developed to extend this comparison of muscle and joint properties.
Lanman, 35
MUSCLE MODEL
Huxley (1957) introduced the crossbridge model of muscle in order to
explain the mechanism of force generation.
Crossbridge models have sub-
sequently been developed to a high level of refinement (and complexity);
however, for the purpose of this study a simpler and more empirical model
seems appropriate.
The model presented here makes no attempt to explain the
mechanism of force generation, but instead is based on the emphirical
relationship between muscle force and resistance to rapid small amplitude
displacement (SRS) (see Introduction).
The concept of actin-myosin bond
bridges is used since it accounts for this relationship so neatly.
A bond not
only contributes one unit to overall muscle force, but also contributes a unit
of mechanical resistance to displacement.
The model resulting from these
assumptions is as follows:
Muscle Model
B
K
AW
(1)
F
(2)
B
mo
(3)
KmMW
,,
fb n
b
= bn
b n
F = steady muscle force
Fb = average force of a single bond
n = number of bonds per 1/2 sarcomere
BmW = equivalent viscosity of muscle
at frequency w
bbw = viscosity of a single bond
bridge at w
z = length of muscle in 1/2
sarcomeres
K = equivalent stiffness of muscle
at frequency w
k
= stiffness of a single bond
bridge at w
Lanman, 36
Muscle force, stiffness and viscosity are all proportional to the number of
bonds.
Also, since the resistance of muscle to stretch depends on the
frequency of stretch, the stiffness and viscosity parameters are subscripted
with frequency (i.e., Bm,
bb,
Km
kbw).
The variation of these parameters
with frequency was measured in isolated muscle (Kawai, 1978) and represented
on a Nyquist plot:
60
quadrature
30
component
(B)
100 Hz
1 Hz
%,
-
A
A
0
-20
30
60
90
in-phase component
(K )
30
60
90
120
in-phase component
(K )
The data points correspond to measurements made at 0.25, 0.5,
1, 2, 3.13, 5, 7, 10, 16.7, 25, 33, 50, 80, 100, 133, 167 Hz. The
quadrature component of resistance is plotted vertically and the
in-phase component horizontally. The scale is in units of 1/e.
Stretch at any frequency encounters resistance with both an in-phase component
(K ) and a quadrature component (B ).
Both of these components of impedance
vary directly with the number of bonds.
Bond Number
Muscle force, stiffness and viscosity all depend on bond number (n).
The number of bonds depends on their rate of formation (Rf) and their rate of
breakage (Rb) (just as world population depends on birth rate and death rate).
Lanman, 37
The rate of bond formation depends on the level of muscle activation
(represented rather abstractly as c), and on the degree of overlap (N) of
actin and myosin filaments:
Rf = rate of bond formation
R = c (N-n)
a = activation
N = number of potential bond sites
(determined by overlap)
n = number of active bond bridges
The rate of bond breakage depends critically on the velocity of movement of
actin and myosin.
This is reasonable since as movement proceeds the bond
bridges, which are only some 100 A long, are stretched and eventually must
break loose.
The faster the filaments slide past one another the faster bond
bridges will be stretched and broken.
In addition to bonds broken by
mechanical disruption, some bonds may break by spontaneous dissociation.
Thus, the total rate of bond breakage (Rb) is:
R =
b
v
T
n + Sn
v = steady rate of displacement
= maximum bond stretch, or
threshold for bond breakage
S = reaction rate for spontaneous
dissociation
Note that reciprocal stretch and release of muscle, such as occurs during
high frequency sinusoidal stretch, will not necessarily dirsupt bond bridges,
so long as the peak to peak amplitude of stretch is small compared to the
threshold (T).
The rate v indicates only the maintained velocity of movement,
movement that is sustained long enough to break bonds.
The rate of bond formation and of bond breakage together determine the
number of bond bridges.
The relationship is particularly simple in the steady
Lanman, 38
state where the rate of formation exactly balances the rate of breakage (zero
population growth means the birth rate equals death rate):
Rf = Rb
a (N-n) = v
n
(4)
n + S n
TT/T + a + Ss
This equation relates the number of bonds (n) to the level of activation (a)
and the rate of movement (v).
When activation is high (a >> v/T + S) all
overlapping bonds are formed (n = N).
When activation is low and velocity
high (a << v/T + S) then bond number approaches zero (n = 0).
Testing the Muscle Model
Now that we have an expression for bond number, it is a simple matter to
compute muscle force, stiffness and viscosity from equations #1, #2 and #3.
For example, the expression for muscle force is:
F =f F nfb=ffb IvI/T Na
+a + S
Force
Force
= high
Fma
max
a= low
a 1/2
(5) Fmax = fb N
v1/2
activation
(6) v1/
2
velocity
= (a + S) T
The force/activation and force/velocity relationships implicit in this
expression are plotted above, and are in good agreement with physiological
Lanman, 39
data (Joyce et al., 1969).
Note that the force/velocity curve has a different
shape at different levels of x. This same dependence is seen in Rack and
Westbury's (1974) data (pp. 467).
To fit physiologic data quantitatively, the
choice of parameter values must be constrained by equations #5 and #6. The
dependence of the overlap parameter (N) on muscle length can be established
from the length/tension characteristic of muscle, although, for the purposes
of this study N can be treated as constant.
The model not only fits the force/velocity and force/activation properties of real muscle, but also accounts for the proportionality between
muscle force and SRS.
This can be seen by combining equation #1 for muscle
force and equation #3 for muscle stiffness giving:
k
kb
k
kbw F
Sfb
This relationship compares well with Schoneberg's emphirical relationship
(Schoneberg et al., 1974) (pp. 23).
The constant 1/e corresponds to kbw/fb'
the ratio of bond tension to bond stiffness (stiffness measured at frequency
W).
Summary of Muscle Model
The model described above is based on the assumption that the actin-myosin
bond bridge is the basic unit generating both overall muscle force and also
resistance to stretch.
Whenever the number of bonds changes, both muscle force
and muscle stiffness change together.
Bond number, in turn, depends on the
level of muscle activation and on velocity of movement.
kenetic model is used to describe this dependence.
A simplified chemical
The muscle model derived
from these assumptions can correctly account for the force/activation and
force/velocity characteristics of real muscle.
This model will now be used
Lanman, 40
to predict how muscle, and consequently joint properties should vary with
activation and movement.
Applying the Muscle Model to the Intact Joint
Both muscles and passive tissues contribute to the mechanical properties
of the intact joint; however, because passive tissue stiffness is small (a
few Nt-m/rad, Hayes and Haytze, 1977) compared to the effects measured in
this study, we will at first consider only the contributions of the muscles.
Thus, the predictions of the muscle model will be directly compared to
measurements on the total joint.
This will be done for conditions of (1)
isometric activation and (2) passive movement.
The condition of isometric activation has, in fact, already been covered
earlier in the Discussion.
By setting the model parameters fb/kb
= E, the
predictions of the model (see equation #7) are exactly as shown in Figure 8.
What is important is that this same model can now be used to predict stiffness
changes during passive movement.
When muscle length changes bonds are disrupted and stiffness decreases.
The relationship between muscle stiffness and velocity can be derived by
combining equations #3 and #4:
K
m
-
kbo
bw
a
Na
Na
Ivl/T + a + S
Converting to angular coordinates and summing all agonist and antagonist
muscles gives total joint stiffness:
k bw r 2 N
K)
iii
Similarly, for joint viscosity:
a.
ri/
+ a + S
i
Lanman,
bbw r
B
B
i6i
2
N
1
These equations are plotted in Figure 15.
41
.
1T +
.+ S
Two curves are shown, for slightly
different levels of resting muscle activity (a).
Possibly the slightly
different results seen on different days can be accounted for by such
uncontrolled changes in resting muscle force.
The approximate fit of the model to the data is perhaps not too surprising
since there are many parameters which may be adjusted.
What the model does
show is that a relatively simple model of muscle which correctly accounts for
the force/activation and force/velocity characteristics of muscle can also
account for the changes in joint properties measured in this study.
The Influence of Passive Tissues
So far we have compared measurements of the mechanical properties of the
intact joint to the predictions of a mathematical model.
effects were ignored.
Passive tissue
The model is in good agreement with the data except
under conditions of isometric muscle activation.
There the model predicts
that muscle stiffness should increase linearly with increasing bias loads,
while measurements on the joint show stiffness saturating at higher torque
levels (Fig. 8).
A possible explanation for this saturation is the com-
pliance of tendon and/or any possible compliance in the coupling between the
torque motor and the bones of the forearm.
The mechanical properties of
tendon and coupling may be quite complex, however, for simplicity these
properties are lumped together in a single visco-elastic element.
Such a
simplified model is adequate to show that a series coupling element can
account for the saturation seen in the stiffness/torque plot.
Lanman, 42
B
Kc
visco-elastic coupling
B
m
K
m
active muscle
The stiffness and viscosity of the coupling element (Kc and Bc) are treated as
constant, while muscle properties (Km and Bm
)
vary according to the muscle model.
All flexor muscles are lumped together and treated as a single equivalent flexor
muscle, and similarly for the three heads of the triceps.
The mathematical
details of the model are covered in Appendix II. The fit of the model to the
data is shown in Figure 16.
Most of the saturation seen at high torque levels
can be accounted for by including the visco-elastic coupling element.
Conclusions from the Muscle Model
No amount of modeling can ever conclusively prove that the joint properties measured in this study are indeed a reflection of the short range
stiffness of muscle; however, this seems a good working hypothesis in view of
the success of this relatively simple model of muscle.
Ultimately it may be
possible to explain and predict all the mechanical properties of the intact
joint from the results of isolated muscle studies.
To rationalize joint
properties in this way would contribute greatly to our growing understanding
of neural control of movement.
this goal.
The present study is only a first step towards
Lanman, 43
CONCLUSION
The mechanical properties of the intact human elbow joint were measured
using small amplitude high frequency sinusoidal perturbations.
The resistance
of the elbow to such perturbation changed systematically as a function of
both the level of muscle activation and velocity of movement.
These are the
first reported measurements of how joint properties change during movement.
These measurements on joint are compared to predictions based on a mathematical model of muscle.
Joint properties appear to reflect the short range
stiffness of muscle.
The significance of short range stiffness in the control of movement is
only beginning to be studied, and any discussion is necessarily somewhat
speculative.
Still, the marked transitions seen in muscle properties are
likely to be important in neural control of movement.
Furthermore, there is
indirect evidence that the central nervous system receives afferent information about very small amplitude disturbances.
This evidence comes from
recent studies of the non-linear properties of the muscle spindle (Goodwin
et al., 1975; Hasan and Houk, 1975a, 1975b; Hulliger et al., 1977a, 1977b).
The spindle receptor is extremely sensitive to small amplitude stretch, and
much less sensitive to larger amplitudes.
Hulliger et al. (1977a, 1977b)
suggest that this transition in sensitivity may be related to the disruption
of bond bridges in the intrafusal muscle fibers.
In any case, the stretch
receptor responds optimally to just those disturbances which stretch and
threaten to break actin-myosin bond bridges.
This tuning of the stretch
receptor suggests that its role may be to report the degree of distortion of
muscle bond bridges.
Nichols and Houk (1976) have proposed an interesting hypothesis relating
Lanman, 44
the non-linear properties of muscle and stretch receptor.
These investigators
compared the response of deafferented muscle to the response of muscle with
the stretch reflex operating.
They found that when deafferented muscle is
stretched it responds with a brief increase in tension and then yields to a
lower tension.
On the other hand, in a preparation in which the reflex is
operating, muscle tension continues to increase with increasing stretch.
In
other words, with reflex pathways intact, muscle appears far more linear in
its response to stretch.
Apparently, the non-linear properties of muscle and
the stretch reflex interact in such a manner that the total system appears
relatively linear.
The function of the stretch reflex may be this lineariza-
tion of muscle properties.
Goodwin et al (1978) propose yet another function for the stretch reflex.
These investigators found that muscle tremor increases after deafferentation.
They suggest that one role of the stretch reflex may be this reduction of
muscle tremor.
The amplitude of physiologic tremor is comparable to the
range of muscle short range stiffness and.also comparable to the high sensitivity range of the stretch receptor.
The measurements reported in this study show that joint stiffness
decreases markedly during passive movement.
This is indirect evidence for
yielding of muscle, and this effect is seen even for very low velocities of
movement.
So low, in fact, that even the movements of tremor, or internal
fasciculations of muscle could cause yielding.
Yielding of muscle implies a
destruction of its ability to generate tension and also a waste of metabolic
energy.
Thus, it would be advantageous for the nervous system to stabilize
muscle against tremor and internal movements.
The stretch receptor, as
discussed earlier, is sensitive to just those amplitudes of stretch which
Lanman, 45
distort and threaten to break bond bridges.
Thus, stretch receptor activity
could be used to reflexively stabilize bond bridges.
For example, if a
fascicle of muscle is stretched to the point that bonds are distorted near to
breaking, then adjacent stretch receptors will be strongly activated.
The
myotatic stretch reflex will cause activation of the stretch muscle fascicle
and thus help stabilize the population of bonds.
In this way the stretch
reflex may help to assure a uniform graded contraction of muscle with a
minimum of fasciculation and tremor.
This would help muscle generate tension
smoothly and efficiently.
Whatever the role of the stretch reflex, it is likely that the reflex is
intimately related to the short range stiffness properties of muscle.
This is
because both the stretch receptor and muscle itself show marked transitions
in their properties whenever the small amplitude range of the SRS is
exceeded.
The role of the stretch reflex will be revealed when we understand
how reflex action interacts with and complements the intrinsic properties of
muscle.
The properties of muscle are important not only in understanding the role
of the stretch reflex but also in understanding more general problems of
neural control of movement (see Introduction).
The purpose of this study
has been to measure and to try to understand the properties of the intact
working elbow joint.
Existing ideas of muscle structure and function are
drawn upon in an attempt to rationalize joint properties.
Hopefully, this
work will stimulate further investigations of the neuro-mechanical transducer
properties of muscle.
These properties are of central importance to under-
standing neural control of movement.
Lanman, 46
REFERENCES
Agarwal, G.C. and Gottlieb, G.L. (1977a) Oscillation of the human ankle joint
in response to applied sinusoidal torque on the foot.
J. Physiol.,
268: 151-176.
Agarwal, G.C. and Gottlieb, G.L. (1977b) Compliance of the human ankle joint.
J. Biomech. Engn,, 99: 166-170.
Allum, J.H.J. and Young, L. (1976) The relaxed oscillation technique for the
determination of the moment of inertia of limb segments.
Biomechanics,
J.
9: 21-25.
Bawa, P., Mannard, A. and Stein, R.B. (1976a) Effects of elastic loads on the
contractions of cat muscles.
Biol.
Cybernetics, 22: 129-137.
P., Mannard, A. and Stein, R.B. (1976b) Predictions and experimental
Bawa,
tests of a visco-elastic muscle model using elastic and inertial loads.
Biol. Cybernetics,
E.,
Bizzi,
Dev,
22: 139-145.
P., Morasso,
P. and Polit, A. (1978) Effect of load dis-
turbances during centrally initiated movements.
J. Neurophysiol., 41:
542-555.
Bizzi,
E. and Polit, A. (1978) Processes controlling arm movements in monkeys.
Science, 201: 1235-1237.
Cohen, C. (1975) The protein switch of muscle contraction.
Sci. Am., 233:
36-46.
Edman, K.A.P., Elzinga, G. and Noble, M.I.M. (1978) Enhancement of mechanical
performance by stretching during tetanic contractions of vertebrate
skeletal muscle fibers.
J. Physiol., 281: 139-155.
Lanman, 47
Emonet-Denand, F., Laporte, Y., Matthews, P.B.C. and Petit, J. (1977) On the
subdivision of static and dynamic fusimotor actions on the primary
ending of the cat muscle spindle.
J. Physiol., 268: 827-861.
Feldman, A.G. (1966a) Functional tuning of the nervous system with control of
movement or maintenance of a steady posture.
meters of the muscles.
II. Controllable para-
Biofizika, 11: 498-508.
Feldman, A.G. (1966b) Functional tuning of the nervous system during control
of movement or maintenance of a steady posture.
III.
Mechanographic
analysis of the execution by man of the simplest motor tasks.
Biofizika,
11: 667-675.
Ford, L.E., Huxley, A.F. and Simmons, R.M. (1977) Tension responses to sudden
J.
length change in stimulated frog muscle fibers near slack length.
Physiol., 269: 441-515.
Goodwin, G.M., Hoffman, D. and Luschei, E.S. (1978) The strength of the
reflex response to sinusoidal stretch of monkey jaw closing muscles
during voluntary contraction.
J. Physiol., 279: 81-111.
Goodwin, G.M., Hulliger, M. and Matthews, P.B.C. (1975) The effects of fusimotor stimulation during small amplitude stretching on the frequencyresponse of the primary ending of the mammalian muscle spindle.
J.
Physiol., 253: 175-206.
Gordon, A.M., Huxley, A.F. and Julian, F.J. (1966) The variation in isometric
tension with sarcomere length in vertebrate muscle fibers.
J. Physiol.,
184: 170-192.
Gottlieb, G.L. and Agarwal, G.C. (1978) Dependence of human ankle compliance
on joint angle.
J. Biomechanics, 11: 177-181.
Lanman,
48
Gottlieb, G.L., Agarwal, G.C. and Penn, R. (1978) Sinusoidal oscillation of
the ankle as a means of evaluating the spastic patient.
J. Neurol.,
Neurosurg. & Psychiatry, 41: 32-39.
Grillner, S. (1972) The role of muscle stiffness in meeting the changing
postural and locomotor requirements for force development by ankle
extensors.
Acta Physiol. Scand., 86: 92-108,
Hasan, Z. and Houk, J.
(1975a) Analysis of response properties of deafferent-
ed mammalian spindle receptors based on frequency response.
J.
Neurophysiol., 38: 663-672.
Hasan, Z. and Houk, J. (1975b) Transition in sensitivity of spindle receptors
that occur when muscle is stretched more than a fraction of a millimeter.
J. Neurophysiol., 38: 673-689.
Hayes, K.C. and Hatze, H. (1977) Passive visco-elastic properties of the
structures spanning the human elbow joint.
Europ. J. Appl. Physiol.,
37: 265-274.
Houk, J. and Henneman, E. (1974) Peripheral mechanisms involved in the con
control of muscle.
pp. 608-616.
In:
V.B. Mountcastle (Ed.), Medical Physiology, 1:
C.V. Mosby: St. Louis, Mo.
Hulliger, M., Matthews, P.B.C. and North, J. (1977a) Static and dynamic fusimotor action on the response of IA fibers to low frequency sinusoidal
stretching of widely ranging amplitude.
J. Physiol., 267: 811-838.
Hulliger, M., Matthews, P.B.C. and North, J. (1977b) The effects of combining
static and dynamic fusimotor stimulation on the response of the muscle
spindle primary ending to sinusoidal stretching.
839-856.
J. Physiol., 267:
Lanman, 49
Huxley, A.F. (1957) Muscle structure and theories of contraction.
Prog.
Biophys. & Biophys. Chem., 7: 255-318.
Huxley, A.F. (1974) Review lecture on muscular contraction.
J. Physiol.,
243: 1-43.
Huxley, A.F. and Simmons, R.M. (1971) Mechanical properties of the crossbridges of frog striated muscle.
Huxley, A.F. and Simmons,
muscular force.
Huxley, H.E.
R.M.
J. Physiol., 218: 59-60P.
(1972) Mechanical transients and the origin of
Cold Spring Harbor Symp. on Quant. Biol., pp. 669-680.
(1965) The mechanism of muscular contraction. Sci. Am., 212:
18-27.
Joyce, G.C. and Rack, P.M.H. (1969) Isotonic lengthening and shortening
movement of the cat soleus muscle.
J. Physiol., 204: 475-491.
Joyce, G.C., Rack, P.M.H. and Ross, H. (1974) The forces generated at the
human elbow joint in response to imposed sinusoidal movements of the
forearm.
J. Physiol., 240: 351-374.
Joyce, G.C., Rack, P.M.H. and Westbury, D. (1969) The mechanical properties
of cat soleus muscle during controlled lengthening and shortening
movements.
J. Physiol., 204: 351-374.
Julian, F.J. and Sollins, M.R. (1975) Variations of muscle stiffness with
force at increasing speeds of shortening.
Kawai, M. (1978) Head rotation or dissociation?
J. Gen. Physiol., 66: 287-302.
Biophys. J., 22: 97-103.
Lazarides, E. and Revel, J.P. (1979) The molecular basis of cell movement.
Sci. Am., 240: 100-113.
Lestienne, F. and Pertuzon, E. (1973) Determination, in situ, de la viscoelasticite du muscle humain inactive.
1-12.
Europ. J. Appl. Physiol., 32:
Lanman, 50
Mason, P. (1978) Dynamic stiffness and crossbridge action in muscle.
Biophys.
Struct. & Mechanism, 4: 15-25.
Merton, P.A. (1953) Speculations on the servo control of movement.
Malcolm and J.A.B. Gray (Eds.), The Spinal Cord: pp. 247-260.
In: J.L.
Little,
Brown & Co.: Boston, Ma.
Morgan, D.L. (1977) Separation of active and passive components of shortrange stiffness of muscle.
Am. J. Physiol., 232: C45-C49.
Moss, R.L., Sollins, M.R. and Julian, F.J. (1976) Calcium activation produces
a characteristic response to stretch in both skeletal and cardiac muscle.
Nature, 260: 619-621.
Murray, J.M. and Webber, A. (1974) The co-operative action of muscle proteins.
Sci. Am., 230: 59-71.
Nichols, T.R. and Houk, J.C. (1976) Improvement in linearity and regulation
of stiffness that results from actions of stretch reflex.
J.
Neurophysiolo, 39: 119-142.
Polit, A. (1977) Functional organization of the motor system during a simple
act.
Ph.D. Thesis, MI.T., Dept. Psychol.
Pringle, J.W.S. (1967) The contractile mechanism of insect fibrillar muscle.
Prog. Biophys. & Mol.
Biol., 17: 1-60.
Rack, P.M.H. and Ross, R.F. (1974) Electromyography of human biceps during
imposed sinusoidal movement of the elbow joint.
J. Physiol., 244: 44-45P.
Rack, P.MoH. and Westbury, D.R. (1969) The effects of length and stimulus
rate on tension in isometric cat soleus muscle.
443-460.
J. Physiol., 204:
Lanman, 51
Rack, P.MoH. and Westbury, D.R. (1974) The short range stiffness of active
mammalian muscle and its effect on mechanical properties,
J. Physiol.
240: 331-350.
Robinson, D.A. (1975) Tectal oculomotor connections in sensorimotor function
of the midbrain tectum.
Neurosci. Res. Prog.
Bull., 13: 238-244.
Schoenberg, M., Wells, J.B. and Podolsky, R.J. (1974) Muscle compliance and
longitudinal transmission of mechanical impulses.
J. Gen, Physiol.,
64: 623-642o
Walker, S.M. and Schrodt, G.R. (1974) I segment lengths and thin filament
periods skeletal muscle fibers of the rhesus monkey and the human.
Anat. Rec.,
178: 63-82.
Lanman, 52
APPENDIX I
Figure Al -- Block Diagram of Apparatus
The torque motor (1) is driven by a sinusoidal current.
The resulting
torque (t ) and acceleration (e)are monitored by transducers (2) and
amplified (3).
Ts
(4).
The joint torque (tj) is computed by subtracting Ie from
The acceleration and joint torque signals are then filtered at the
frequency of torque motor drive (5).
The amplitude (6) and phase relation-
ship (7) of these filtered signals are determined and recorded for
subsequent computer analysis.
Figure A2, a, b -- Schematic Diagram of Electronics
Figure A3 -- Narrowband Filter Characteristics
Figure A3, a -- Gain/Frequency Plot with a Center Frequency of 30 Hz
Figure A3, b -- Phase/Frequency Plot
Although the filter has considerable phase shift, this does not
interfere with phase measurement since both the torque (r.) and acceleration
(e)signals are filtered and thus phase shifted by the same amount.
The
small difference in phase shift between the two filters is plotted on the
bottom of the figure.
Figure A4 -- Equivalent Circuit for Strain Gauge Preamplifier
The preamplifier circuit used with the strain gauge is slightly
unusual in that the intrinsic resistance of the gauge is an integral part
of the circuit.
This configuration gives good low noise performance with a
minimum of components.
Figure A5 -- Equivalent Circuit for Accelerometer Preamplifier
The preamplifier used with the accelerometer is a charge amplifier.
To clarify its operation, the accelerometer is redrawn in the equivalent
Lanman, 53
circuit as a signal source (es ) in series with a capacitor, Ce . If
Cfb = Ce , then the voltage gain of the circuit is -1. The resistor network
provides bias current without significantly degrading the low frequency
performance of the circuit.
Fig. Al
ARM
VIBRATION
preamplifiers
Lanman, 54
APFARATUS - ELECTRONICS
subtractor
narrowband
filters
rectifier &
lo pass
2T
Tj
I
strain guage
!
to phase
I
I
c
I
Inertim adetector
detector
trim
I
I
9
I
I
iT
5
6
II
DETECT OR
PHASE
7
no
lo pass filter
0
Sphase
"I
I
I
i- 1
I
I
torque
motor
power
amplifier .
~~AP
2-2. 1K
-r
1
"T t 2
"W"
10--
ZE &O
390k
30P~
-
LP4
KJ
j
LocAkL
ViS-kv%
Sk-500k
5
Oi
50k
zH
06
a
q9
IIIII~IUU
017E
Ol
t+i~~-i+~tttt+ttt~~+tttt+ttt+~tr--
11
-41 L11414
-1
f 41
Nil
- W44 I- 4
" 1 1
4i4-
i ii4 4
-
I,,z
141
41 1-1
1 1I
1
SL
I
---
rr.
-I -
-+
-1
-- -
Hiiii-f
+
I t 1 1 11
+-t-H-I-t-H--iH
1 1 1 1 1 1 1 1 1
1
i-ff-t'
"
1 1
1
1
-
1
"- F
1
1
1
--I- !
1-4-+ i i
1 1 t I I
I
I
I
-HH-i
I
I
I
I
-t-!hf-th
I
I
I
I
i
I
a:
-tH
i
tH
i± ±±
ttl±If
i-
f±H-1i
t-t±1±t h! ±
i
I
t
LE
N1, D
+ 1
--
t -1 Ll
-flT
oo
-45
4tt
-F _F
T-1IT
it -
T
0-
_UAA.
50
t Iti-I
11
M
Hz
-e
__-15
20-
tfflt
.25
9
.30
31
40--
1
____.60
-------
DIFFERENTIAL
PHASE S FIT
____.. 1-fl-
ec
-4Q
1IL
I I I II I IIt
II
Fig. A4
STRAIN GUAGE PREAMPLIFIER
+5V
strain guage
bridge
EQUIVALENT CIRCUIT
Lanman, 59
Fig. A5
Lanman, 60
ACCELEROMETER PREAMPLIFIER
.01 mfd polycarbonate
piezoelectric
220K
accel erometer
2.2K
EQUIVALENT CIRCUIT
Cfb
Lanman, 61
APPENDIX II
The muscle model, including a viscoelastic coupling element is as follows:
f
x
Kt
Bt
Km
Km
B
Bm
coupling element
x
muscle
The basic equations describing this system are:
(1) f = Kt(x-x ) + Bt(x-x')
(2) f = Kmx' + Bmx'
Rewriting these equations using the Laplace variable gives:
(3) f = Kt(
x-
x') + Bt(x-x')s
(4) f = K x' + B x's
m
m
Solving these equations for the real and imaginary (quadrature) components
of resistance to sinusoidal displacement (x) gives:
equivalent stiffness = freal
x
2
2
Kt Km + Km Kt + K B
22 + KtB 22
(Kt+Km)2 + (Btw+BW)
equivalent viscosity =
f.
imaginary
2
=
x
2
2
(Kt B + K Bt)
2
2
+ (Bm Bt +B
(Kt+Km) 2 + (BtW+Bm )2
) 2
Lanman, 62
These equations were evaluated using the following parameter values:
(1) For flexor loads:
Kt = 240 Nt-m/rad
Bt = 1.6 Nt-m-sec/rad
K =kT
m
w
B =bT
m
T = bias load
w
k and b are functions of frequency:
w (Hz) k
(sec)
b
15
14
.15
20
20
.128
25
25
.064
30
30
.032
(2) For extensor loads all parameter values were multiplied by
the constant .66.
This constant may reflect the ratio of the
moment arms of the flexor to the extensor muscles.
Lanman, 63
FIGURE LEGENDS
Fig. 1
(A) This figure shows the arrangement of actin and myosin filaments in
skeletal muscle (from Grey's Anatomy,
3 5th
British Edition, p. 481; 1973).
(B) This figure shows how the structure of muscle determines the
isometric tetanic length/tension characteristic.
It is likely that the
physiologic range of muscle lengths corresponds to the rising portion of
this length/tension curve (from Gordon et al., 1966).
Fig. 2
This figure, from Rack and Westbury (1969), is a comparison of the
length/tension properties of muscle as measured under static and dynamic
conditions.
The dashed lines show static length/tension characteristics
measured at different rates of muscle stimulation.
is the static stiffness of muscle.
The slope of these curves
Superimposed on these static characteris-
tics is a series of solid lines representing the response of muscle to
dynamic (ramp) stretch.
In each case muscle tension increases rapidly for
the first millimeter or so of stretch but then yields.
The initial slope of
these dynamic characteristics is a measure of the short range stiffness of
muscle.
This slope, or SRS, becomes greater with increasing muscle tension.
Fig. 3
Schematic diagram of apparatus.
Fig. 4
This figure shows the effect of small errors on the computed values of
equivalent stiffness and viscosity.
An error in amplitude measurement, or in
estimating the inertia of the arm cast, effects computation of stiffness.
The effect of a 1% error is plotted in the top graph.
An error in phase
Lanman, 64
measurement principally effects computation of viscosity, and the effect of
a 0.5 degree error is plotted in the bottom graph.
Errors of the magnitude
shown are likely, however, such constant errors would not interfere with
measurements of changes in equivalent stiffness or viscosity.
Fig. 5
This figure shows that the computation of K and B is uninfluenced by
non-oscillatory movements of the arm/cast.
This control data was recorded by
filling the arm/cast with water, vibrating the system at 15 Hz and computing
K and B .
Since the arm/cast had no viscous or elastic connections, the
true value of both K and B is zero.
The computed value of K and B stay
near zero despite the imposed movements of the arm/cast.
Fig. 6
This figure shows the transient response of the system.
Again, the arm/
cast is filled with water, so the true values of K and B are zero.
A small
inertial load (0.0018 kg m2 or 1% of the total inertia of the arm/cast) is
held to the cast by an electro-magnet.
off, allowing the inertial load to drop.
time constant of about 100 msec.
transient.
At arrow #1 the magnet is switched
K responds exponentially with a
B shows a small increase during this
Since no viscous load was present at this time, this increase is
probably an artifact; however, it is too small to be significant.
arrows #2 and #3 the load was reattached to the magnet by hand.
Between
Contact of
the hand with the cast produces a large apparent viscous load during this
interval.
After arrow #3, both traces return to their initial values.
Fig. 7
This figure was recorded by asking the subject to generate increasing
torque with his forearm.
The arm is held approximately isometric by a nylon
Lanman, 65
cord attached to the counterweight.
The resistance of the forearm joint to
a superimposed 25 Hz perturbation is continuously plotted.
Both equivalent
stiffness (K ) and equivalent viscosity (B ) increase as the steady arm
torque increases.
Fig. 8
This plot shows the increase in equivalent stiffness (AK ) as a function
of increasing bias load.
The data was recorded by asking the subject to hold
a constant arm position while various weights were suspended from the arm.
Stiffness was computed from the resistance of the forearm to a superimposed
15 Hz oscillation.
The solid lines show the stiffness/torque relationship
predicted from isolated muscle studies (see text).
Fig. 9
This figure shows the effect of passive arm movement on K and B .
The
servo was switched on for about one second causing the arm to move at about
50/sec.
There is no change in EMG activity during this movement.
decreases dramatically during movement, and then recovers.
K
B also shows a
moderate change.
Fig. 10
This plot shows the decrease in equivalent stiffness (AK ) as a
function of velocity of arm movement.
Fig. 11
This figure is a comparison of arm movements made with different levels
of voluntary muscular activation.
Condition (A) shows a passive arm movement,
i.eo, a servo driven movement unaccompanied by any change in EMG activity,
(B) shows a voluntary arm movement, a movement generated by a voluntary burst
of muscle activity and (C) shows a voluntary movement against a viscous load.
Lanman, 66
Note the increased EMG activity required to overcome the load.
stiffness and viscosity were measured at 25 Hz.
Equivalent
In the passive movement
case, K decreased, much as in Fig. 9. During voluntary movement, K drops
again, but not so far as during passive movement.
During the loaded
movement K initially drops but then recovers during the period that the
muscles are actively generating force to overcome the viscous loado
Fig. 12
This figure shows the repeatability of measurements of joint properties
during movement.
Figures A, B and C show passive, active and loaded
movements, respectively.
In each case four individual records are
superimposed.
Fig. 13
This schematic figure shows how the pattern of stiffness changes seen
during a viscous loaded movement (C) can be explained as the sum of the
stiffness changes associated with (A) isometric muscle activation and (B)
with passive movement.
See text for further explanation.
Fig. 14
This figure shows the relative timing of a viscous loaded movement.
sequence of events is:
(1) just before one second the triceps
EMG decreases slightly
(2) just after one second stiffness begins to decrease and the arm
starts moving at about 2 degrees/second
(3) at about 1.4 seconds the biceps muscle is strongly activated,
stiffness increases markedly and the arm starts moving rapidly.
The
Lanman, 67
Fig. 15
This figure shows the decrease in joint stiffness with increasing joint
velocity.
Data points are plotted together with predictions of the model
(continuous curves).
shown.
Data for measurement frequencies of 15 to 30 Hz are
For the model, all flexor and extensor muscles are lumped together
(ri = r, i = z and a. = a) and the parameters are adjusted to satisfy two
equations.
(AK)
w max
=
61/2 =
Kb
r
2
(a + s)
Na
r/r
For most plots, two curves are shown for slightly different levels of resting
muscle tone (a).
These two curves are fit to the highest and lowest values
of AK wmax
max"
Fig. 16
This figure shows the increase in equivalent stiffness (AK ) and
equivalent viscosity (AB ) as a function of bias load.
shown as negative.)
Data points are plotted together with predictions of the
model (continuous curve).
are shown.
(Triceps loads are
Data for measurement frequencies of 15 to 30 Hz
Note that the inclusion of a visco-elastic coupling element can
account for the leveling off of these curves at higher torque levels.
Appendix II for the parameter values used.
See
Lanman, 68
Fig. 1
SONE
SARCOMERE
I 8S-1 90), b-c
I ..------- II4.65 2S(20
I
-
105
1.
40
10
Acem
LENGTH
1-$
2
LENGTH
1.
20
5
30
.1s
4
25
30
3-
40
--
Fig
Lanman, 69
2
35 impulses/sec
10 impulsessec
2
5 impulsesisec
C
0
,,7
-"
/-.
,_
C
4,
,/
d
.,,:r.-
, Passive
"- Muscle length (cm)
1
1500
1
1200
I
900
Angle of ankle
I
1
600
300
Fig. 3. Tei'iisii (ldring e'Xtelis(iiI thr
ligh differenIt 1ariisf t he lhllngt
h rniii'e.
frion a liumber (of (differelit exteilsionis simlihlr to) titse
I'tetisi)l re e(iri
to show ho1( the teltSili
shown in Fig. 2 have beenl traced (conitnulOs lille)
varios pai'ts of the
ditring lengthening differs froin the isometri tentision ini
leitgth lirange. Positioins, of the ankle ji int are sho'wn1 Ielo\\ c irirsl igli
pairts of the abseissa.
In nell case theo muscle was lengthetned tllroulgh 6 ilmm at 7-2
The intterrupted lioies
are isonmetric length-tension plots.
tnul/sec.
ACCE LE ROME TER
COUNT ERWE IGHT
CRETE
BI AS
LOAD
OT i NBG
SINE
ORiVE
Lanman, 71
Fig.
ERROR
K
Nt-m
rad
30 Hz
15
+1% amplitude error
a
v,
I
sS
.
I
I
I
a
I
I
quadrature component
in-pha se
error
ERROR
0.6 4
B
Nt-m-sec
rad
0.3
I
I
r
. -1
J quadrature error
error
f
.. ..
.50
--
'
|
in-phase component
Fig.
2.5
t
----r~t~
_sccSrc~t~LI~-~
B
-
ru
I
f
I0025
Lanman, 72
5
Nt-m-sec
rad
POSITION
1 SECOND
~i~
t
I
I
I
10 Nt-m
rad
I
I
K
I
J-n
B(I-
O.05 Nt-m-sec
rad
500 msec
I
I
2
I
3I
I
Fig.
N-m
100rad
rad
N-m-s
r.6
rad
18 ftlb
Lanman, 74
STIPPFFNESS
VISCOSITY
TORQUE
,,
biceps
I
,
,
ii,
,
i
t riceps
1 SECi
llIII
1
Iil i i i li;,iuli l~i titi i~i il i ilil iiiiiiI :II
iiSEC1.
I
I
!
lllliIl
llil
itlll"
""' "" """ll
111iilllll nllIldIIIHlil
Lanman,
Fig. 8
75
EQUIVALENT
STIFFNESS
AKw
-r
( w= 15 Hz)
I
prediction from isolated
muscle
I
Nt-m
150
rad
e
100
o
50
20
10
20
ft-lb
30
nt-m
BIAS LOAD
0 - JL
10/23/79
E
- JL
10/7/79
o
- TZ
11/14/79
O-
NH
10/8/79
30
40
Fig. 9
Lanman,
76
N-m
STI FFPN ESS
so rad
I
1
I\
I\
II
1
I
I
I
I
I
I
\
\
\
\
Z
--
1
II
I
I
I
I
I
I
I
II
,
I
I
I
I
I
I
I
I
I
I
I
t
I
I
I
I
I
I
I
I
I
(
III
1
!
N~m~
I
I
V1SOSIi
.3i
1
1
I
I
t
~I
i
I
--
(
ARM
10
~I
i
,z
POSITION
EMG
biceps
illlll1llllllll llll
l
llllllllilll1
111111111
1l1111IIIIIIIIIIIIII1111
llll lnlllll I111111111lll
triceps
1 SEC
lll
tii
I lllllllllllllllllllll11l
llllllll ll
lllllll Illilllllllllllllllllllllllllllllll llll ull
Fig, 10
Lanman, 77
AK
0
CHANGE IN
EQUIVALENT
STIFFNESS10
(w = 15 Hz)
20
30
40
50
1.5
0
3
6
ARM VELOCITY
TZ
11/14/79
1 - JL
10/23/79
O
- NH
10/8/79
0
- JL
9/17/79
S-
DEGREES/SEC
10
1
1
Nm
50 rad
I
STIFFNESS
I
1
I
I
SN-ms
i
VISCOSITY
rad
-- A I
I
I
1I
I
II(
II /
-rl
- - --------------
20
POSITION
ft-lb
EMG
ILI
biceps
,. .a ,,.11.,,,.1(1,.1)
t l ll'tlllt ll ,tl, l ilil illlli"ltl
III l llll
lllt l l il tillttlli
illillll li l l tlillIllut
l lilll
llllltlllllll
liii
Fil
i iil'jijgii
iiiiiiiii
bilillil
piei
lilll
triceps
1
SEC
A
PASSIVE MOVEMENT
B
ACTIVE
MOVEMENT
Cf
LOADED MOVEMENT
0
---
--
r-o \
-1.
PASSIVE MOVEMENT
(A)
VOLUNTARY MOVEMENT
(B)
LOADED MOVEMENT
(C)
ISOMETRIC CONTRACTION
A
(1) STIFFNESS
C ACTIVE MOVEMENT
LOAD TORQUE
STIFFNESS
ARM POSITION
(2)
-----
(1)
-.-
I-..
LOAD TORQUE
B PASSIVE MOVEMENT
I
(2) STIFFNESS
I
LOAD TORQUE
ARM POSITION
I
zz
ARM POSITION
I
..-
,
1I
Fig.
II
1
I
50
Nm
rad
STIFFNESS
II
I
I
It
I
t
I
II
I
I
I
I
I
II
3
Nms
rad
I
I
I
I
I
r
I
I
VISCOSITY
I
0!
s ±sec
POSITION
VELOCITY
EMG
I
IllIIOCT
II
triceps
,
I
II
,I
I
I
I
I
I
I
I
I
I
,
I
I
I
1
I
I
I
I
I
I
~~-------i
,
II
20o
1I
II
I
I
I
I
I
I
I
r
I
I
I
I
1
I
I
Lanman,
14
II
I
lilil Sbiceps
II
1 SEC
81
~I_
_
___
Nt-m
rad
0
i
Nt-m
rad
= 30 Hz
JL 10/23/79
AS= 20 Hz
e- JL
AK
9/17/79
O- NH 10/ 8/79
50
o- TZ 11/14/79
100
3
6 d/
6 deg/sec >10
deg/sec
>10
S= 25 Hz
0
AK
S= 15 Hz
40
80
3
6
>10
deg/sec
0
3
6
>10
deg/sec
Fig.
16
Lanman, 83
Nt-m
rad
400
Nt-m -sec
rad
+
e
w-30
-20
0
-20
30
0
30
2
400
AK
0
0
AB
00
200
0
ea
1
0o
0
ee,
w-25
I
0
I•
J,
I
',
I
-20
I
,
20
-
I
I
I
0
30
0
30
2
AB
w=20
-20
0
30
-
20
2
AB
W-15
i
-20
30
0
BIAS
LOAD
()- JL
E-
JL
(ft-lb)
|
,
i1
-20
BIAS
LOAD
(ft-lb)
10/23/79
O- NH
10/ 8/79
9/17/79
O - TZ
11/14/79
,
9
Lanman, 84
TABLE I
FREQUENCY (Hz)
AMPLITUDE (degrees p-p)
0.1 - 0.025
0.14 - 0.035
0.2 - 0.05
0.3 - 0.075
Lanman, 85
TABLE II
MEASUREMENT
REFERENCE
Schoneberg et al.
(1974)
INTERVAL
FREQUENCY
0.1 msec
(10 kc)
0.3-1%
0.1 msec
(10 kc)
0,5%
(2 msec)
500 Hz
0.5%
(3 msec)
300 Hz
(10 msec)
100 Hz
(1 sec)
1 Hz
Table I
Ford et al. (1977)
Fig. 26
Julian & Sollins (1975)
Fig.
Mason (1978)
pp. 20
Kawai (1978)
see Nyquist plot
Rack & Westbury (1974)
pp. 347
This study
10 mm/sec or about:
(0.1 sec)
(10 Hz)
(30 msec)
30 Hz
(60 msec)
15 Hz
1.5%
3-5%
1 5%
Download