A Dissipative String Model For Human Tendons

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

1 Copyright 20xx by ASME

Proceedings of the ASME 2014 International Design Engineering Technical Conferences &
Computers and Information in Engineering Conference
IDETC/CIE 2014
August 17-20, 2014, Buffalo, New York, USA
DETC2014-34304
A DISSIPATIVE STRING MODEL FOR HUMAN TENDONS


Davide Piovesan
Gannon University
Erie, PA, USA
Michael J. Panza
Gannon University
Erie, PA, USA


ABSTRACT
This work presents an analytical modelling of a tendon
driven system. We modelled the tendon as a continuum string
using physiologically plausible value for the intrinsic
characteristics of biological tissues to verify the systems
stability. We used Poincares sections to verify the types of
attractors the system has and proved that the system is non-
chaotic in a large range of physiologically plausible excitation
frequencies. This work is relevant for understanding the control
of force and movement as a function of the muscle contraction.
Wide applications can also be found as a guideline for the
design of artificial muscles and cable driven robots.
INTRODUCTION
Ones ability to exert controlled forces on the environment,
such as when manipulating fragile objects, is very important in
everyday life. This suggests that force regulation is a necessary
component of motor control [1].
Classical force controllers are generally implemented in
systems with small contact impedance in order to limit the force
applied to the end effector by external disturbances [2]. In
biological motor systems, as the force generated by the muscle
fibers increases, so does the muscles stiffness [3]. This
behavior is directly in contrast with the need to keep the contact
impedance low. More refined muscles models must include a
tendon elastic element in series with the muscle, making them
more adherent to the systems physiology and better suited to
the study of force control strategies [4, 5]. This solution seems
appropriate as the Golgi tendon organs (i.e. the organelles
responsible to the biological transduction of force) are located
at the interface between tendon an muscle. Compliant tendons
act as stabilizing impedance buffers between the force-
generating element (muscle) and the environment. This
mechanical configuration is known in robotics as series-elastic-
element control [6], where the force at the end effector is
exerted controlling the position of an actuator in series with a
linear spring. Recently, this solution has been used to control
robotic humanoid locomotion [7].
Notice, that the presence of non-linearities in a tendon
driven system can generate instabilities and limit-cycle
behaviors that needs to be compensated by the controller. Using
an analytical approach, this work investigates the physiological
parameters that influence the exertion of force by controlling
the tension at the interface between muscles and tendons.
Modelling the tendon as a string, we established the relationship
between the time-varying modal shape of the tendon and the
time-varying position of limbs inertia which is directly related
to the contact force.
When the tendon is described as an elastic string rather
than a lumped stiffness, the regulation of force occurs only by
tension regulation via the force generated by the muscle. Hence,
the control scheme is a position control where the only
controlled variable is the extension of the tendon. This is in
agreement with findings previously published that empirically
showed the construction of an internal feed-forward model
trajectory as part of a force-regulation strategy [1].

GENERAL ANALYTICAL MODEL
When studying the motion of a mechanical system, ( ) t x

is
a vector of generalized position coordinates (angles, Cartesian
coordinates, etc.). We can define x D
n
as the set representing
the position coordinates and their derivatives with respect to
time up to the
th
n order so that
|
|

\
|

= x
t
x
t
x
t
x
x D
n
n
n
, , ,...,
1
1
2
2
, in
general n [8].
A mechanical system must comply with the Lagrange
d'Alembert principle so that

( ) ( ) ( ) ( ) ( ) t x D g t x D t x
dt
d
t x M
n n
, , ,
2
2

= +
(1)
2 Copyright 20xx by ASME
where ( ) t x M , is the inertial matrix of the system in the
chosen coordinate frame, ( ) t x D g
n
,

is the external force field,


and ( ) t x D
n
,

is the internal force field generated by the


mechanical network [8].
POSITION CONTROL AT THE TENDON INTERFACE
We derived a set of differential equations describing the contact
force of a tendon-driven inertia. In such equations the only
variable governing the system is the commanded extension of
the tendon
0
z .

Figure 1 Schematic of the tendon driven system
Figure 1 represents a schematic of a tendon-driven system
where L is the length of the tendon, M is the mass of bones
and teguments, h is the curvilinear co-ordinate of the tendon,
is the distributed damping effect of fluid around the tendon,
0
z is the initial displacement of the tendon,
0
T its initial
tension,
2
X the position of the point of contact in the Cartesian
space.
Let us consider a modified equation derived from the standard
string theory that relates the transverse displacement of the
tendon
1
x at position z to the tension force ( ) t T of the tendon
interface, generated by the muscle fibers:
2 2
1 1 1
2 2 T T
x x x
A d L T
t t z


+ =

(2)

Equation (2) takes into consideration the total damping of the
fluid applied on the tendon longitudinal surface, where
T
d is
the average diameter of the tendon.
We can safely assume that human have a limited control
bandwidth [9]. Hence, we can restrict our analysis to the first
mode shape of the tendon transversal oscillation that is the
form:
( ) ( ) L z t X x / cos
1 1
= (3)

At the interface between muscle and tendon where 0 = z
(hence, the tendon is neither stretched nor slaked), substituting
(3) in (2) gives, after an easy algebraic manipulation, the
following differential equation:

2
1 1 1 2
0
T T
T
X X X
A A L


| | | |
+ + =
| |
\ \

(4)

We can define the tendon stiffness as a function of its shape and
material
L
E A
K
T S
= , where
T
A is the section area of the
tendon, and E is the Youngs modulus of the material [4]. The
total tension of the tendon is given by [10]

1) the extension of the tendon
0
z generated by the
displacement induced by the muscle fibers;
2) the tension generated by the displacement
2
X of the
arm mass M and
3) the extension due to the tendon elastic deformation.

Namely:

( ) + + =
2 0
X z
L
E A
T
T
(5)

The extension due to the tendon elastic deformation of an
elastic string can be calculated as follows

( )


|
|
|

\
|

|
|

\
|
|

\
|

+ = =
2
2
2
1
1
L
L
dz dz
z
x
dz dh (6)
where dh is the variation of the curvilinear coordinate along
the tendon. Assuming 1
1
<<

z
x
due to the extensive length of
the tendon with respect to the lateral amplitude of oscillation,
dh can be approximated with the first two terms of its Taylors
series, that is [11]:

2
1
2
1
2
1
1 1 |

\
|

+
|
|

\
|
|

\
|

+
z
x
z
x
(7)

Substituting (7) and (3) in (6) we obtain

L
X
dz
z
x
dz dz
z
x
L L
L
4 2
1
1
2
1
2
2
0
2
1
2
2
2
1

= |

\
|

=
|
|

\
|

|
|

\
|
|

\
|

(8)


Therefore, the total tension of the tendon is

3 Copyright 20xx by ASME
|
|

\
|
+ + = + =
L
X
X
L
E A
z
L
E A
T T T
T T
d
4
2
1
2
2 0 0

(9)

Substituting (9) in (4) we obtain


2 2
2 2 1
1 1 1 1
0 0
1 0
4
T
T
d L X X
X X X
A z Lz

| | | |
+ + + + =
| |
\ \

(10)

Which is the equation determining the wave function of the
tendon, with transversal natural frequency:

2 2
2 0 0
1 0 2 3
T
T Ez
z
A L L


| | | |
= = =
| |
\ \
(11)

We can readily see that the temporal frequency of the first mode
shape is linearly dependent with the extension of the tendon
0
z .
The longitudinal oscillation of the force at the point of contact
with the environment is:

0
2
= + X M T
d

(12)

Hence, considering the substitutions

L
W
4
2

= ;
L M
E A
M
K
T
S
n

= =
2
;
T
T
d L
A

= (13)

We can find a system of two non-linear equations that describe
the behavior of the system in free space.

( )
2
1 1 0 2 1 1
2 2 2
2 2 1
0
0
n n
X X z X WX X
X X WX

+ + + + =

+ + =

(14)

Or in a more concise form

3
1 1 2 1 1 0
2 2 2
2 2 1
0
n n
X X X X WX z
X X WX


+ + + =

+ + =

(15)

Notice, that (15) is a system of Duffing equations with
nonlinear terms (quadratic and cubic) and nonlinear coupling
(i.e.
2 1
X X ). We can notice that the term
0
z is the only
control variable for the previous equation. The commanded
extension of the tendon has become a forcing term, while all the
other parameters are constants (i.e. intrinsic characteristics of
the biomechanical system).
NUMERICAL EXAMPLES

We assumed the contraction of the muscle (i.e. the forcing
function of the system) in the following form:

( )
0
cos 2 z Z f t = (16)

where Z =0.01[m]. Equation (15) was solved for a finite set of
input forcing frequencies f =0.253[Hz] with 0.25[Hz]
increments.
We transformed the system into its ordinary differential
equations (ODE) form and solved numerically using ODE23tb
in MATLAB.

( )
3 1
4 2
3
3 2 1 1 1
2 2 2
4 2 1
cos 2
n n
X X
X X
X X X X WX Z f t
X X WX


=

(17)

With initial conditions

0 1 0 2 0 3 0 4 0
0; ( ) ( ) ( ) ( ) 0 t X t X t X t X t = = = = =

The dimensions of the tendon were adjusted to fit the oscillation
of a human forearm as in [12]. The intrinsic features of the
material (i.e. , , E ) have the same order of magnitude to the
frog animal model used in [13] .
Given the numerical values of the parameters used in the
simulation (See Table 1), a forcing function with frequency
lower than 0.25[Hz] would lead to stable but unlikely results
(see Figure 2 for f =0.125[Hz]).


Table 1 Mechanical parameter of the system

T
d 0.003[m]
L 0.15[m]
M 2[kg]

1000 [kg/m]
T
A 7.0686e-06[m
2
]
E 20 [MPa]
20e07 [Ns/m
2
]
4 Copyright 20xx by ASME

Figure2 solution of (15) using f =0.125[Hz]
In Figure 2, the displacement of the mass is one order of
magnitude larger than the imposed forcing-function amplitude
and approaches the same order of magnitude of the tendons
length, making this condition physiologically impossible. This
ill outcome is likely to depend upon the fact that the condition
1
1
<<

z
x
is not satisfied.

Figure3 Lateral deviation of the tendon as function of time
and frequency


Figure4 Movement of the mass as function of time and
frequency
The results in the aforementioned forcing signal frequency
range are presented in Figure 3 and 4. Immediately, we can
observe that the non-linearity of the system create a frequency
shift between the forcing term
0
z , and the movement of the
mass
2
X .
Figure 4 shows a beating for
2
X in the 1.52[Hz] range. The
phenomenon can be observed in more details in Figure 5 where
the interval 1.52[Hz] has been analyzed with a frequency
resolution of 0.125[Hz]. We can observe a very low frequency
modulation of
2
X envelop.
The aforementioned phenomenon is a self-oscillation effect
caused by a parametric resonance.


Figure 5 Beating of Variable X
2
caused by the phase shift
between
0
z and
1
X

Recalling the first equation of (15) in the form

( )
2 3
1 1 1 1
cos 2 X X X WX Z f t

+ + + =

(18)

the natural frequency of the system is a function of the mass
oscillation, namely:

2
2
( ) ( ) t X t

= (19)

If the mass oscillation is sinusoidal and the non-linear term is
negligible due to the small amplitude of oscillation

( ) ( ) ( )
2
2
1 cos X t a t

;
3
1
0 WX (20)

Equation 18 becomes a forced Mathieu equation in the form

( ) ( ) ( )
2
1 1 1
1 cos cos 2 X X a t X Z f t

+ + =



with 1 a << , small oscillations are seen to grow exponentially in
time if the angular frequency is close to 2 / ; n n

.
5 Copyright 20xx by ASME

Figure 6 Quadrature between
0
z and
1
X at f =1.75[Hz]

It is interesting to notice that this effect produces a difference in
phase between the forcing function
0
z and the amplitude of the
tendons transversal oscillation
1
X . Figure 6 illustrates the
quadrature between
0
z and
1
X . In a linear system, such 90
phase difference would correspond to a resonance. However,
given the physical configuration of the system, the instability is
induced in the longitudinal vibration of the mass
2
X and not
on the transversal oscillation
1
X .
From Figure 4 we can also infer that
2
X is usually negative.
When equation (19) is negative, the system does not have an
oscillating stable solution if the damping term is not included in
the system. The effect of the term
1
X

is to stabilize the
lateral oscillation of the tendon by allowing it to store enough
elastic energy to bounce back, making the string acting similarly
to a long flexible beam [11].
FLOW ANALYSIS
Equation (17c) and (17d) are in the form ( ) ,
dX
F X t
dt
= and
therefore are, by definition, 3-Dimentional flows in the phase
space. To test the stability of the system, we plotted the phase
diagrams of both the tendons transversal oscillation (
1
X vs
3 1
X X =

) and the longitudinal oscillation of the mass (
2
X vs
4 2
X X =

). The former are represented in Figure 7, the latter in
Figure 8.
For each excitation frequency the flow of the transversal
amplitude
1
X follows a single closed orbit and has an attractor
point in (0,0) (Figure 7). This behavior is also confirmed by the
Poincares sections in Figure 9.


Figure 7 Phase diagram of tendons transversal oscillation
as function of the forcing frequency


Figure 8 Phase diagrams of the mass longitudinal
oscillation as function of the forcing frequency. Origin is
highlighted in red.
6 Copyright 20xx by ASME


Figure 9 Poincares sections of tendons transversal
oscillation as function of the forcing frequency


Figure 10 Poincares sections of the mass longitudinal
oscillation as function of the forcing frequency.

Poincares sections are used to check the isochronicity of the
orbits. A point is represented for each period of the forcing
function. If there exist only one point in the map the system is a
simple oscillator with one frequency. If multiple points are
present, there exist multiple frequencies of the system that are
phased locked. We can speculate that even though multiple
points appear in Figure 9 for many excitation frequencies, the
variation along
1
X is sub-millimiter with respect to a total
amplitude that is one order of magnitude larger. These multiple
points are likely be produced by numerical integration errors.
Instead, the analysis of the flow generated by
2
X is not so
immediate. The phase diagram indicate that the flow can switch
from a relative close orbit (Figure 8 0.25[Hz]) to a toroidal flow
or possibly a strange attractor (Figure 8 1.5[Hz] and 2[Hz]). By
the simple observation of the phase diagram it is not possible to
assess if the system is not chaotic. To determine the non-chaotic
behavior of the system we again used Poincares sections.
Figure 10 shows clearly that the points in the sections are
distributed around a close line; thus the flow is toroidal and
non-chaotic. As the excitation frequency increases we can
observe an increase of points along the curve indicating a shift
in the frequency ratio between
2
X and
0
z . If a finite number
of points exist
2
0
X
z
f
f
is a rational number. If the curve on the
section becomes continuous, the ratio
2
0
X
z
f
f
is an irrational
number.

DISCUSSION
In this paper we developed a series of analytical tools to
investigate the force control of human subject when the muscle-
tendon system is describe using standard string theory. We
developed an analytical set of equations where we demonstrate
that when represented the tendon as an elastic string, the control
of the external force is done indirectly by controlling the
position of the muscle-tendon interface.
This analytical demonstration based on a realistic
biomechanical model suggest that in order to control the force,
subjects need to create an internal model of their mechanics, so
that the time-varying contraction of the muscle must have a
different frequency than the load-tendon complex. These results
shed some light on the mechanical properties of muscle-tendon
system, confirming that using a continuous string model does
not produce chaotic behaviors.
NOMENCLATURE
t time
s complex number
7 Copyright 20xx by ASME
( ) t x

, ( ) s X

generalized co-ordinate and respective


Laplace transform
n degrees of the derivative with respect to time
rational numbers
x D
n
set of all the derivatives of ( ) t x

with respect
to t up to the
th
n degree
M , generalized mass of the arm and respective
Laplace transform
S
K series stiffness of a PT system representing
the stiffness of tendons
j imaginary unit 1
2
n
square of the systems natural frequency
z coordinate along the un-deformed tendon
) (
0
t z initial extension of the tendon
h curvilinear coordinate along the deformed
tendon
) , (
1
t z x modal shape of the tendon as function of time
and position along it
) (
1
t X time-varying modal shape of the tendon
) (
2
t X time-varying position of arm inertia
L tendon length
linear density of the tendon
T
A Tendons cross section area
T Tendon total tension
0
T Tendon tension fron the extension
0
z
d
T Tendon tension due to the movement of the
arm and intrinsic deformation due to the
tendons elasticity
W curvature module of the tendon
2
1
natural frequency of the wave propagation in
the tendon
REFERENCES

[1] Kolesnikov, M., Piovesan, D., Lynch, K. M., and
Mussa-Ivaldi, F. A., 2011, "On Force Regulation Strategies
in Predictable Environments," eds., pp. 4076-4081.
[2] Mason, M. T., 1981, "Compliance and Force Control
for Computer Controlled Manipulators," Systems, Man
and Cybernetics, IEEE Transactions on, 11(6), pp. 418-
432.
[3] Piovesan, D., Pierobon, A., Dizio, P., and Lackner, J.
R., 2013, "Experimental Measure of Arm Stiffness During
Single Reaching Movements with a Time-Frequency
Analysis," Journal of Neurophysiology, pp.
[4] Piovesan, D., Pierobon, A., and Mussa Ivaldi, F. A.,
2013, "Critical Damping Conditions for Third Order Muscle
Models: Implications for Force Control," Journal of
Biomechanical Engineering, 135(10), pp. 101010-101010.
[5] Piovesan D., Pierobon A., and Mussa-Ivaldi F.A., 2012,
"Third-Order Muscle Models: The Role of Oscillatory
Behavior in Force Control. ," eds., pp.
[6] Pratt, G. A., and Williamson, M. M., 1995, "Series
Elastic Actuators," eds., 1, pp. 399-406 vol.1.
[7] Bortoletto, R., Sartori, M., He, F., and Pagello, E.,
2012, Simulation, Modeling, and Programming for
Autonomous Robots, Springer Berlin Heidelberg,
Modeling and Simulating Compliant Movements in a
Musculoskeletal Bipedal Robot.
[8] Padovan, J., and Guo, Y., 1988, "General Response of
Viscoelastic Systems Modelled by Fractional Operators,"
Journal of the Franklin Institute, 325(2), pp. 247-275.
[9] Piovesan, D., Pierobon, A., Dizio, P., and Lackner, J.
R., 2012, "Measuring Multi-Joint Stiffness During Single
Movements: Numerical Validation of a Novel Time-
Frequency Approach," PloS one, 7(3), pp. e33086.
[10] Sun, B. N., Wang, Z. G., Ko, J. M., and Ni, Y.-Q.,
2001, "Cable Parametric Oscillation and Its Control for
Cable-Stayed Bridges," eds., 4330, pp. 366-376.
[11] Panza, M. J., 1999, "Mathematical Model for Large
Deflection Dynamics of a Compliant Beam Device,"
Journal of Dynamic Systems, Measurement, and Control,
123(2), pp. 283-288.
[12] D., P., and Huang, F. C., 2013, "Control Mechanisms
in Oscillatory Motor Behavior," eds., pp.
[13] Cecchi, G., Griffiths, P. J., and Taylor, S., 1986,
"Stiffness and Force in Activated Frog Skeletal Muscle
Fibers," Biophysical Journal, 49(2), pp. 437-451.

You might also like