Conference PaperPDF Available

An integrated motor control loop of a human-like robotic arm: Feedforward, feedback and cerebellum-based learning

Authors:

Abstract and Figures

A new complex model of human motor control has been developed, combining brain internal models and neural network mechanisms. Based on nervous system structures and operating principles, a feedforward block, a feedback controller and a cerebellum-like learning module have been integrated and tested with an anthropometric robotic arm. A simulated sequence of 8-like tracking tasks showed the contributions of these main loops over time. Different external dynamics were introduced. The role of feedback corrections, intrinsically imprecise due to sensorimotor delays, decreases, while the output of cerebellum, which has been learning, increases; the movement becomes more accurate. Moreover, an experimental session on a subject performing the task repetitions using a haptic device was carried out, recording upper limb kinematics.
Content may be subject to copyright.
mi
Abstract— A new complex model of human motor control
has been developed, combining brain internal models and
neural network mechanisms. Based on nervous system
structures and operating principles, a feedforward block, a
feedback controller and a cerebellum-like learning module
have been integrated and tested with an anthropometric
robotic arm. A simulated sequence of 8-like tracking tasks
showed the contributions of these main loops over time.
Different external dynamics were introduced. The role of
feedback corrections, intrinsically imprecise due to
sensorimotor delays, decreases, while the output of cerebellum,
which has been learning, increases; the movement becomes
more accurate. Moreover, an experimental session on a subject
performing the task repetitions using a haptic device was
carried out, recording upper limb kinematics.
I. INTRODUCTION
HE biological motor system is a high performance
control engine. Unlike artificial control systems, it
exhibits much higher performance with great flexibility
and versatility in spite of nonlinearities, uncertainties and
large Degrees of Freedom (DoF) of animal bodies.
Sensorimotor function is created from a highly distributed
circuit that includes different neural centers, such as cerebral
cortex, cerebellum, and spinal cord.
The movement kinematic planning to achieve a
particular task is assigned to the premotor and
somatosensory cortical areas; they generate the optimal
trajectory and transform this external-space Cartesian
coordinates into internal-space joint coordinates through
inverse kinematics processing. It was shown that
somatosensory cortex cells encode joint-centered
kinematics; their activity is correlated with position, velocity
and acceleration parameters [1].
Then, the motor commands, in order to achieve such
desired kinematics, are defined. The brain must construct
internal models of the plant, objects and environment only
through learning by experience and memorize them in its
neural networks in a usable format for motor control. The
primary motor cortex (M1) is considered the site where basic
inverse dynamic models are stored, thus behaving as a
nonlinear feedforward controller able to compute torque
values. Since possible joint miscalibrations, context changes,
noise and other uncertainties, this structure is not capable of
guaranteeing an accurate control on its own.
The cerebellum is able to fine-tune motor skills by
processing incomplete or approximate commands issued by
higher levels of motor system [2]. It is in charge of temporal
and spatial movement coordination. Its structure, made up of
microzones acting as functional units, fits well with the
learning mechanisms. Patients suffering cerebellar
dysfunctions (e.g. cerebellar ataxias) are almost unable to
deal with disturbances as they can rely only on the imprecise
and unstable feedback control to enhance the basic inverse
model of motor cortex [3].
The action of a feedback controller in motor control is
well accepted. The role of M1 in this loop is proved,
neurophysiologically, by a dense projection from M1 to the
spinal cord, often directly onto motor neurons, and by a
number of correlations between M1 firing and end-effector
kinematics [4]. Ito [5] showed that the feedback controller
generates a command in motor cortex, which can tune the
viscoelastic properties of musculoskeletal system (tension-
length and tension-velocity relationships). Adaptive
feedback controllers have been proposed [6], which means
that the pre-programmed arm impedance changes in
response to feedback information. For instance, it has been
shown that impedance increases around the task constraints
[7]. Feedback gains, which convert sensory state variables
into motor signals, are optimized based on specific goals of
a particular behavior, by following the ‘minimum
intervention principle’ [8], [9]. Thus, the optimal feedback
control consists of two main steps: state/error estimation and
control laws.
No model of human-like motor control including this
overall complexity has been built. In all biological systems,
in which all the different parts have been evolved together
towards global aims, the whole is more complex than the
sum of its parts. Kawato [10] proposed the adaptive
nonlinear feedforward controller, based on a feedback-error
learning architecture; that is, error signals from linear
feedback controller tune the feedforward inverse model
parameters. Schweighofer and colleagues [11] showed how
the cerebellum may increase the accuracy in target reaching
movements by compensating for the interaction torques, thus
by learning a portion of the inverse dynamics model that
refines a previously stored basic inverse model in the motor
cortex. Other models were proposed and tested on planar
movements of a robotic arm, e.g. [12], [13].
Starting from functional/anatomical schemes and from
these previous important steps, the present study integrates
control models [14], learning models, neural network
dynamics and behavioral observations, by using both
modeling/computational and experimental approaches on
multi-joint 3D movements.
An integrated motor control loop of a human-like robotic arm:
feedforward, feedback and cerebellum
-
based learning
C. Casellato, A. Pedrocchi, J.A. Garrido, N.R. Luque, G. Ferrigno,
E. D’Angelo
, E. Ros
II. M
ODELING
A
PPROACH
A. Control system
Fig.1 comes up with the main neural structures and
mutual connections involved in motor control, highlighting
the ones implemented in our model.
Premotor and sensory cortex blocks compute the
kinematic planning. First, the desired trajectory is generated
following a minimum-jerk criterion in external space, so
facing the kinematic redundancy [15]. Then, closed-form
inverse kinematics is carried out to compute displacement,
velocity and acceleration for each of the three joints [16].
The nonlinear feedforward controller is placed into M1
block; it is made up of an inaccurate inverse dynamic model
of the arm based on recursive Newton-Euler dynamic
equations computing joint torques. These dynamic equations
do not take into account friction torques, inertial interaction
torques (i.e. inertia tensor matrix presents zero terms for the
off-diagonal elements), and internal neural noise. The latter
here consists of both sensorial noise on actual kinematics
and signal-dependent noise on total torques, i.e. proportional
to motor command amplitudes [17], [18]. Moreover, the
inverse model does not include unexpected external force
changes, embedding just the very well-known gravity action.


 



The linear feedback controller, receiving somatosensory
information from the periphery, is sited within M1. By
means of exploiting muscular viscoelasticity, an additive
torque value is produced depending on the ongoing error, as
an online correction. Its performance is limited due to the
system nonlinearities and the inevitable feedback
sensorimotor delays. Because of muscle spindles do not
carry a significant amount of acceleration information, only
position and velocity are present in the feedback controller.
Position and velocity errors (e
p
, e
v
) are weighted by gains
(K
p
and K
v
: elasticity and viscosity features, respectively);
this arm impedance is selected depending on the task
requirements and keeping in mind that high feedback gains
enhance robustness to external perturbations but, at the same
time, increase noise (signal-dependent noise) and metabolic
cost. It would imply non-compliant and non-stable
movements [19], [20].

The plasticity mechanisms are implemented by the
cerebellum and inferior olive blocks. The cerebellum learns
to provide corrective torques towards reducing the kinematic
errors in incoming trials; thus, it acts as a predictor.
This biological adaptation takes place on the parallel fiber
to Purkinje cell synapses, driven by the activity from the
Inferior Olive that here encodes a teaching signal (dependent
on the accuracy of the movement execution compared to the
desired movement trajectory).This system implements a
look-up table which associates each parallel fiber state [27]
with a Purkinje cell output. This association is iteratively
modified during the learning process.
The adaptation mechanism is based on LTD/LTP (Long-
Term Depression and Potentiation) processes validated in
previous approaches [21] with a linear firing rate cerebellar
model. This cerebellum-like model delivers add-on output
corrective torque terms based on the received feedback error
Figure 1. Model
The scheme includes the main neural structures and functional interconnections involved in motor control. In red the blocks and the
connections implemented in our control model.
along previously executed trials. The cerebellar module
torque action is defined following the equation:
  
Where Gainsout allow a rescaling based on the torque ranges,
MFgain represents the activity coming through mossy fibers
(this input activity has been fixed to 1 in order to normalize
the output activity in between 0 and 1) and PCi(t) is the
Purkinje cell firing rate associated to the currently active
parallel fiber. This activity is iteratively modified following
the equation:


 

Where LTPMax and LTDMax are parameters which regulate
the learning plasticity mechanism speed (both have been
fixed to 0.2), e represents the error signal (a linear
combination of joint position and velocity error, normalized
between 0 and 1) and α regulates the LTP/LTD interaction
(it has been set up to 1000 to reduce LTP action in presence
of a significant error). Finally, PCi(t) is configured to be
always working in the range [0, 1].
The expected behavior of the whole system should be
that the feedback controller progressively is driven out, since
the cerebellum adjusts progressively internal models. Thus,
the desired motions will be mainly predicted and only small
correction forces will be required, so increasing the system’s
control compliance.
B. Plant
A robotic arm is built with 3 rotational DoFs: q1
represents shoulder abduction/adduction, q2 shoulder
lowering/elevation, and q3 elbow extension/flexion, as
reported in Fig.2-a. The kinematic parameters are defined
according to Denavit-Hartenberg convention (Fig.2-c),
setting the link lengths depending on subject’s
anthropometric measurements. The inertial parameters for
each link, such as mass, CoM position, and inertia tensor are
set depending on subject’s anthropometric measures as well
[22], [23], [24], [25].
C. Simulations
The control loop and the robot plant have been built up
in Simulink (Mathworks®), using a Robotic Toolbox [26].
A simulation of an 8-like trajectory tracking task in 3D
was carried out, where one trial lasted 4 s and 20 repetitions
were performed. The unexpected external force
perpendicular to end-effector was a step: from 0.6 N to 2.3 N
at half of each trial duration. The time resolution was 2 ms.
The signal-dependent noise was a white noise with
amplitude equals to 2% of the torque amplitude.
The feedback controller was characterized by Kp and Kv
proportional to the external force modulus: Kp = 3·|Fext|
[Nm/rad]; Kv = 1·|Fext| [Nm/(rad/s)]. The delay was 50 ms.
The cerebellum module was implemented by a linear firing
rate model of the cerebellum which includes some of the
traditional working hypothesis of the cerebellum, such as
the generation of non-recurrent states at the granular layer
[27], synaptic plasticity at the parallel fibers driven by the
climbing fiber activity and synaptic integration at the
Purkinje layer [28], [29].
For each task repetition, multiple variables were
recorded. The different contributions on the total torque (τtot)
were computed, as ratio of Root Mean Square (RMS)
values:
τcerebellum/τtot = RMS(τcerebellum) / [ RMS(τcerebellum) +
RMS(τfeedback) + RMS(τfeedforward) ]
τfeedback/τtot = RMS(τfeedback) / [ RMS(τcerebellum) +
RMS(τfeedback) + RMS(τfeedforward) ]
The Cartesian error of the end-effector was evaluated by
using two main parameters: RMS-Error and the correlation
between the desired and the actual 3D trajectories.
D. Results
The main simulation results are reported in Fig.3. It is
evident, from the joint angles (3-a), that the distance
between the desired and the actual movements decreases
over time. It is also supported by the Cartesian trajectories
(3-c) and the performance indexes (3-d and 3-e).
Panel 3-b shows how the different controllers contribute
to the whole motor commands over time; the feedback
component is predominant in the very first trials, while the
cerebellum is still learning.
Figure 2. Human-like robotic arm
(a) Robotic arm built for our model, with 3 DoF. Green (q1):
shoulder abduction/adduction. Red (q2): shoulder
lowering/elevation. Blue (q3): elbow extension/flexion.
(b) Experimental set-up: the 3-
marker tools on the involved joints,
the haptic device and the graphical interface with the required task.
(c) The conventional Denavit-Hartenberg parameters which define
the kinematic features of the anthropometric robotic arm.
Latest repetitions show a cerebellum correction activity
bigger than the feedback one for all joints, with quite stable
trends. It is worthy to note that the q3 curves (green) are
higher than the other joints even if the error is smaller (as it
is shown by the green curves in panel 3-a) since this joint
feedforward torque does not include the gravity component;
q3 indeed moves on the horizontal plane.
III. EXPERIMENTAL APPROACH
A. Set-up
Preliminary experimental sessions on one healthy subject
were carried out in order to qualitatively compare the
simulator kinematic movement and a realistic one. The
subject’s upper limb segments were acquired by a motion
capture system (VICRA, PolarisTM), thus placing a 3-
passive-markers tool on each joint (shoulder, elbow and
wrist). A haptic device (PHANToM OMNI, SensAbleTM)
was used to perform the task, developing in Visual C++ a
task-specific visual interface and a control algorithm proving
the subject with the external force changes. By displaying a
countdown, the subject was aware of the required trial
duration. Start and end points were marked through
touchable spheres within the task environment. The set-up
picture is reported in Fig.2-b.
In order to constrain the movement to the selected 3 DoFs,
the subject worn a wrist plaster cast, so as the haptic device
pen was like a forearm extension. The subject was instructed
to avoid as much as possible the use of finger DoFs, the
shoulder rotation and any translation.
After few familiarization trials, the subject was asked to
perform 5 trials with a low constant external vertical force
field (0.6 N) and 5 trials with a force field change from the
half of each trial duration (from 0.6 N to 2.3 N).
Figure 3. Simulation with external perturbation at half of each task repetition
(a) The 3 DoF angles (q1, q2, and q3, as in Fig.2-
a), with time (80 s, i.e. 20 repetitions). Solid curves: the actual joint
angle; dashed curves: the desired joint angle.
(b) The % contributions, with respect to the total generated torque, of cerebellum torque (star c
urves) and of feedback
torque (circle curves). These values are reported for each joint and for each task repetition.
(c) The 3D Cartesian end-
effector trajectories. In violet: the desired one. In black: the first repetition. In thick grey: the
last repetition. In thin grey: the intermediate repetitions.
(d) The Root Mean Square Error between the 3D desired trajectory and the actual one, for each repetition.
(e) The correlation coefficient between the 3D desired trajectory and the actual one, for each repetition.
B. Results
Fig.4 depicts the main representative results, concerning
the 5 trials with external disturbance. In Fig.4-a, the
experimental angles are laid on the ones which come from
simulator planning (desired joint angles). It is evident that,
whereas the shoulder DoFs (q1 and q2) fit quite well with
the desired ones, the elbow flexion (q3) is significantly
smaller in the experimental data than in the simulation
approach. This could be because the subject used also other
DoFs, such as fingers, to achieve the task.
Fig.4-b draws the cartesian end-effector trajectories, and in
Fig.4-c and 4-d, performance indexes, analogous to the ones
computed for simulation, are reported along the 5
repetitions. Both parameters show values that are similar to
the ones achieved in simulation after the first trials. It could
be explained by the fact that the subject performed some
trials for familiarization before recordings, even if not
enough training time to get a stable behavior. Next
experiments will foresee more repetitions, so as to achieve a
convergent trend of performance indexes.
IV. DISCUSSION
The model presented here for motor control revealed
itself neurophysiologically plausible and comparable with
experimentally-based modeling. It successfully puts together
different flexible controllers and predictors, including both
control-based and neural-network blocks in a whole complex
system.
The model behavior can be explored for other tasks (e.g.
first tests on reaching task have been carried out) and for any
dynamic environment. Multiple factors can be set; for
instance, the inverse dynamic model inaccuracies, the task-
dependent optimal feedback law and the time constant of
cerebellar learning rules.
A lot of enhancing steps will be implemented within the
model. The analogical model of the cerebellum can be
replaced with a spiking network version similarly as
presented in [13] and [21] (EDLUT: Event-Driven Look-Up
Tables), which can naturally include more realistic plasticity
mechanisms [30], starting from the most recent dualism
between neurophysiology evidences and neural network
modeling, e.g. [27]. Furthermore, the cerebro-cerebellar loop
could be exploited within the model. In this direction, a first
attempt was carried out through a recurrent architecture
model, where the cerebellum output modified the motor
cortex input, i.e. the kinematic planning, so solving the
motor error problem [31].
Finally, the neurophysiology demonstrated that, after
learning, the inferior olive response decreases significantly
[12], thus suggesting that when the cerebellum learning has
been completed, the learning consolidation occurs
transferring this information directly to the motor cortex, i.e.
making directly the feedforward generated motor commands
more accurate.
In conclusion, this model is using a control scheme
consistent with the motor-learning theory, in which the
motor error is pre-computed and sent to the parallel fiber -
Purkinje cell connection of the cerebellum, in order to
generate LTD and LTP through a supervised learning rule.
Thus, the model now provides the basis for testing more
biologically plausible architectures and computational
solutions, including vector coding in the motor cortex,
implicit learning in the cerebellar granular layer, and various
signal transformations in the different nuclei involved. In
particular, the expansion of the cerebellum into a detailed
neuronal network using the EDLUT simulator will allow to
test the impact of biological circuit and cellular properties on
the control capabilities of the cerebro-cerebellar loop.
Figure 4. Experimental data
(a) The 3 DoF angles (q1, q2, and q3, as in Fig.2-a), with time (20
s, i.e. 5 repetitions). The vertical lines bound each repetition. Solid
curves: experimental data; dashed curves: the desired joint angles
from simulator planning.
(b) The 3D Cartesian end-effector trajectories. In violet: the desired
one. In black: the first repetition. In thick grey: the last repetition.
In thin grey: the intermediate repetitions.
(c) The Root Mean Square Error between the 3D desired trajectory
and the actual one, for each repetition.
(d) The correlation coefficient between the 3D desired trajectory
and the actual one, for each repetition.
ACKNOWLEDGMENT
This work has been supported by the EU grant REALNET
(FP7-ICT-270434).
REFERENCES
[1]
J. F. Kalaska, D. A. Cohen, M. Prud'homme, and M. L. Hyde, "Parietal
area 5 neuronal activity encodes movement kinematics, not movement
dynamics.," Exp Brain Res, vol. 80, pp. 351-364, 1990.
[2] M. Ito, "Cerebellar circuitry as a neuronal machine.," Prog Neurobiol
,
vol. 78, pp. 272-303, 2006.
[3] A.J Ba stian, "Cerebellar limb ataxia: abnormal control of self-
generated and external forces.," Ann N Y Acad Sci, vol. 978, pp. 16-
27, 2002.
[4] E. Todorov, "Direct cortical control of muscle activation in volun
tary
arm movements: a model.," Nat Neurosci, vol. 3, pp. 391 -398, 2000.
[5]
M. Ito, "Control of mental activities by internal models in the
cerebellum.," Nat Rev Neurosci, vol. 9, pp. 304-313, 2008.
[6] J. Nakanishi and S. Schaal, "Feedback error learnin
g and nonlinear
adaptive control.," Neural Netw, vol. 17, pp. 1453-1465, 2004.
[7]
R. Osu, K. Morishige, H. Miyamoto, and M. Kawato, "Feedforward
impedance control efficiently reduce motor variability.," Neurosci Res
,
vol. 65, pp. 6-10, 2009.
[8]
S.H Scott, "Optimal feedback control and the neural basis of volitional
motor control.," Nat Rev Neur osci, vol. 5, pp. 532-546, 2004.
[9]
D. Mitrovic, S. Klanke, R. Osu, M. Kawato and S. Vijayakumar, "A
computational model of limb impedance control based
on principles of
internal model uncertainty. " PLoS One, vol. 5, e13601, 2010.
[10]
M. Kawato, K. Furukawa, and R. Suzuki, "A hierarchical neural-
network model for control and learning of voluntary movement.,"
Biol
Cybern, vol. 57, pp. 169-185, 1987.
[11]
N. Schweighofer, M. A. Arbib, and M. Kawato, "Role of the
cerebellum in r eaching movements in humans. I. D istributed inverse
dynamics control.," Eur J N eurosci, vol. 10, pp. 86-94, 1998.
[12]
N. Schweighofer, J. Spoelstra, M. A. Arbib, and M. Kawato, "Role of
the cerebellum in reaching movements in humans. II. A neural model
of the intermediate cerebellum.," Eur J Neurosci, vol. 10, pp. 95-
105,
1998.
[13]
R.R Carrillo, E. Ros, C. Boucheny, and O.J. Coenen, "A real-
time
spiking cerebellum model for learning robot control.," Bi osystems
, vol.
94, pp. 18-27, 2008.
[14]
J.J Craig, Introduction to robotics: mechanics and control., 2005.
[15]
E. T odorov, "Optimality principles in sensorimotor control.,"
Nat
Neurosci, vol. 7, pp. 907-915, 2004.
[16]
D. Kostic, M. Bra m de, a nd R. Hensen, "Modeling and identification
for high-performance robot control: an RRR-
robotic arm case study,"
IEEE Transactions on Control Systems Technology, vol. 12, pp. 904-
919, 2004.
[17]
C. M. Harris and D. M. Wolpert, "Signal-
dependent noise determines
motor planning.," Nature, vol. 394, pp. 780-784, 1998.
[18]
K.E Jones, A.F Hamilton, and D.M Wolpert, "Sources of signal-
dependent noise during isometric force production.," J Neurophysiol
,
vol. 88, pp. 1533-1544, 2002.
[19]
E. Burdet, R. Osu, D. W. Franklin, T. E. Milner, and M. Kawato, "The
central nervous system stabilizes unstable dynamics by learning
optimal impedance.," Nature, vol. 414, pp. 446-449, 2001.
[20]
J. Porrill and P.Dean, "Recurrent cerebellar loops simplify adaptive
control of redundant and nonlinear motor systems.," N eural Comput
,
vol. 19, pp. 170-193, 2007.
[21]
N.R. Luque, J.A. Garrido, R.R. Carrillo, O.J. Coenen, and E. Ros,
"Cerebellarlike corr
ective model inference engine for manipulation
tasks.," IEEE Trans Syst Man Cybern B Cybern, vol. 41, pp. 1299-
1312, 2011.
[22]
D.A. Winter, Biomechanics and Motor Control of Hu man Movement
.
John Wiley \& Sons, 1990.
[23]
P. de Leva, "Adjustments to Zatsiorsky-
Seluyanov's segment inertia
parameters.," J Biomech, vol. 29, pp. 1223-1230, 1996.
[24]
H.A. Abdullah, C. Tarry, R. Datta, G. S. Mittal, and M. Abderrahim,
"Dynamic biomechanical model for assessing and monitoring robot-
assisted upper-limb therapy.," J Rehabil Res Dev, vol. 44, pp. 43-
62,
2007.
[25]
A.H. Vette, T. Yoshida, T.A. Thrasher, K. Masani, and M.R. Popovic,
"A complete, non-
lumped, and verifiable set of upper body segment
parameters for three-dimensional dynamic modeling.," Med Eng Phys
,
vol. 33, pp. 70-79, 2011.
[26]
P. Corke, "Robotics Toolbox for Matlab," 2008.
[27]
T.Yamazaki and S. T anaka, "The cerebellum as a liquid state
machine.," Neural Netw, vol. 20, pp. 290-297, 2007.
[28]
D. Marr, "A theory of cerebellar cortex.," J Physiol, vol. 202, pp. 437-
470, 1969.
[29]
J.S. Albus, "A theory of cerebellar function," Math Biosci
, vol. 10, pp.
25-61, 1971.
[30]
E. D'Angelo, "Neural circuits of the cerebellum: hypothesis for
function.," J Integr Neurosci, vol. 10, pp. 317-352, 2011.
[31]
N. R. Luque, J. A. Garrido, R. R. Carrillo, S. Tolu, and E. Ros,
"Adaptive cerebellar spiking model embedded in the control loop:
context switching and robustness against noise.," Int J Neural Syst,
vol.
21, pp. 385-401, 2011.
... The ultimate challenge appears then to run the whole-cerebellum network model in a simulated brain operating in closed-loop. While a radical approach is out of reach at the moment (it would require, in addition to fully developed cerebellum models, also realistic models of large brain sections outside the cerebellum), a first attempt has been done by reducing the complexity of cerebellar models and using simplified versions to run closedloop robotic simulations (Casellato et al., 2012Garrido et al., 2013;Luque et al., 2014Luque et al., , 2016. ...
... Clearly, in this case a top-down approach is adopted and the relationship of the simplified model with the real system is a matter of speculation. This approach has been used to generate cerebellar spiking networks (SNN) allowing to reproduce a single basic cerebellar module running with high efficiency in a robotic controller yet maintaining some fundamental features of neurons and connections (Casellato et al., 2012Garrido et al., 2013;Luque et al., 2014Luque et al., , 2016. For example, in these models, neurons were represented by integrate-and-fire single-compartment elements, the local inhibitory interneuron networks were not included and the GCL was not fully implemented resorting to the concept of a non-recurrent states in a liquid-state machine (Yamazaki and Tanaka, 2007). ...
... Despite the simplicity of the cerebellar SNN (Figure 6), the robots that incorporated it revealed remarkable emerging properties (Casellato et al., 2012. The SNN robots correctly performed multiple associative learning and correction tasks, which ranged from eye-blink conditioning to vestibulo-ocular reflex (VOR) and force-field correction. ...
Article
Full-text available
The cerebellar microcircuit has been the workbench for theoretical and computational modelling since the beginning of neuroscientific research. The regular neural architecture of the cerebellum inspired different solutions to the long-standing issue of how its circuitry could control motor learning and coordination. Originally, the cerebellar network was modelled using a statistical-topological approach that was later extended by considering the geometrical organization of local microcircuits. However, with the advancement in anatomical and physiological investigations, new discoveries have revealed an unexpected richness of connections, neuronal dynamics and plasticity, calling for a change in modelling strategies, so as to include the multitude of elementary aspects of the network into an integrated and easily updatable computational framework. Recently, biophysically accurate "realistic" models using a bottom-up strategy accounted for both detailed connectivity and neuronal non-linear membrane dynamics. In this perspective review, we will consider the state of the art and discuss how these initial efforts could be further improved. Moreover, we will consider how embodied neurorobotic models including spiking cerebellar networks could help explaining the role and interplay of distributed forms of plasticity. We envisage that realistic modelling, combined with closed-loop simulations, will help to capture the essence of cerebellar computations and could eventually be applied to neurological diseases and neurorobotic control systems.
... Spiking neural networks (SNN) exploit a biologically observed phenomenological element in ML, allowing optimization and parallelizability to algorithms (Naveros et al., 2017) which may be event-driven and time-driven and may incorporate spatiotemporal information processing capabilities of biological neural circuits. Algorithms that are based on different brain circuits, such as the visual cortex (Fu et al., 2012;Yamins and DiCarlo, 2016), basal ganglia (Doya, 2000;Baladron and Hamker, 2015;Girard et al., 2020), and cerebellum (Casellato et al., 2012;Garrido et al., 2013;Antonietti et al., 2015;D'Angelo et al., 2016b;Luque et al., 2016;Yamaura et al., 2020;Kuriyama et al., 2021) with spiking neural models help understand the circuitry and in turn, help reconstruct and train systems. EDLUT (Ros et al., 2006), SpiNNaker (Khan et al., 2008), MuSpiNN (Ghosh-Dastidar and Adeli, 2009), and biCNN (Pinzon-Morales and Hirata, 2013) are some of the existing brain-inspired models which are used in the field of control systems for robotic articulation control. ...
... The current model was reconstructed based on known microcircuitry of the cerebellum (Eccles, 1967), electrophysiological behavior (D'Angelo et al., 2001), and significant plasticity rules (Mapelli et al., 2015). Even though there are several models available that look at the different aspects of the cerebellum (Casellato et al., 2012;Garrido et al., 2013;Antonietti et al., 2015;D'Angelo et al., 2016b;Luque et al., 2016;Yamaura et al., 2020;Kuriyama et al., 2021), the present model covers certain aspects of the cerebellum while some loops are skipped. Initial models looked at only single-layered neurons (Marr, 1969;Albus, 1971) which were extended with many other layers and plasticity rules. ...
Article
Full-text available
Spiking neural networks were introduced to understand spatiotemporal information processing in neurons and have found their application in pattern encoding, data discrimination, and classification. Bioinspired network architectures are considered for event-driven tasks, and scientists have looked at different theories based on the architecture and functioning. Motor tasks, for example, have networks inspired by cerebellar architecture where the granular layer recodes sparse representations of the mossy fiber (MF) inputs and has more roles in motor learning. Using abstractions from cerebellar connections and learning rules of deep learning network (DLN), patterns were discriminated within datasets, and the same algorithm was used for trajectory optimization. In the current work, a cerebellum-inspired spiking neural network with dynamics of cerebellar neurons and learning mechanisms attributed to the granular layer, Purkinje cell (PC) layer, and cerebellar nuclei interconnected by excitatory and inhibitory synapses was implemented. The model’s pattern discrimination capability was tested for two tasks on standard machine learning (ML) datasets and on following a trajectory of a low-cost sensor-free robotic articulator. Tuned for supervised learning, the pattern classification capability of the cerebellum-inspired network algorithm has produced more generalized models than data-specific precision models on smaller training datasets. The model showed an accuracy of 72%, which was comparable to standard ML algorithms, such as MLP (78%), Dl4jMlpClassifier (64%), RBFNetwork (71.4%), and libSVM-linear (85.7%). The cerebellar model increased the network’s capability and decreased storage, augmenting faster computations. Additionally, the network model could also implicitly reconstruct the trajectory of a 6-degree of freedom (DOF) robotic arm with a low error rate by reconstructing the kinematic parameters. The variability between the actual and predicted trajectory points was noted to be ± 3 cm (while moving to a position in a cuboid space of 25 × 30 × 40 cm). Although a few known learning rules were implemented among known types of plasticity in the cerebellum, the network model showed a generalized processing capability for a range of signals, modulating the data through the interconnected neural populations. In addition to potential use on sensor-free or feed-forward based controllers for robotic arms and as a generalized pattern classification algorithm, this model adds implications to motor learning theory.
... The NRP not only includes a variety of robot and environments, but also a detailed physics simulator. The architecture follows the guidelines from different studies (Caligiore et al., 2017;Houk and Wise, 1995;Casellato et al., 2012;Tomita and Yano, 37 2007;Santos and Matos, 2011) and includes CNS regions such as the brain stem (action regulation), the cerebellum (motor adaptation), the spinal cord (motor pattern generation), the basal ganglia (action selection), and the motor cortex(Initiation, planning, procedure of motion ). The CNS areas will be modeled combining classical control and robotics methods together with bio-inspired AI techniques. ...
Conference Paper
Full-text available
Nanotechnology is a frontier science with tremendous theranostic potential for applications in neurological disease management. Gold nanoparticles (GNP’s) have remarkable surface plasmonic optical properties that enable use-cases in diagnostic medical imaging and treatment of movement dis-orders such as Parkinson’s Disease where it has recently been shown to be neuroprotective. Turkevich’s method of GNP synthesis was used to study monodispersity and nucleation by varying parameters such as reactant stir-rate, concentration etc. Experiments involved optical and spectroscopic data-capture as time-series measurements of the formation process. Deep neural network (DNN) and ensemble models were built and applied with MATLAB and Python.
... The NRP not only includes a variety of robot and environments, but also a detailed physics simulator. The architecture follows the guidelines from different studies (Caligiore et al., 2017;Houk and Wise, 1995;Casellato et al., 2012;Tomita and Yano, 37 2007;Santos and Matos, 2011) and includes CNS regions such as the brain stem (action regulation), the cerebellum (motor adaptation), the spinal cord (motor pattern generation), the basal ganglia (action selection), and the motor cortex(Initiation, planning, procedure of motion ). The CNS areas will be modeled combining classical control and robotics methods together with bio-inspired AI techniques. ...
Article
Full-text available
Recent studies have demonstrated that autonomous robots can outperform the task they are programmed for, but are limited in their ability to adapt to unexpected situations (Ingrand and Ghallab, 2017). This limitation is due to the lack of generalization, i.e., the robot can not transfer knowledge across multiple situations. Even the application of modern artificial intelligence (AI) techniques does not support a robust generalization when the range of probable inputs is infinite (Yang et al., 2018; Mnih et al., 2015; Cai et al., 2017; Kober et al., 2013). As a matter of fact, AI methods can interpolate knowledge but not extrapolate it, i.e., they can adapt on new, unseen data that are within the bounds of their experience, but not on data that are outside the bounds. So far, robots have been mostly treated as stand-alone systems in a vacuum, while the real world is more complex and includes continuous interaction with external entities. Accordingly, the design of a generalized robotic controller is not trivial, in particular when the dynamical condition are unknown. From the observation of nature, it is possible to deduct the level of competence that animals have when interacting with the environment. The study and understanding of the central nervous system (CNS), which is the main responsible of the body complex movements during the interaction with the environment (Wolpert and Ghahramani, 2000; D’Angelo and Wheeler- Kingshott, 2017), may give new insights about the artificial replication of the animals’ interactive and adaptive behaviour. As a matter of fact, the CNS is constituted by different regions which role, relation and distribution are important for the optimal execution of complex tasks (see (Caligiore et al., 2017) for a review). This investigation has its foundation in the Human Brain Project (Markram et al., 2011), which is trying to achieve a more clear understanding of the brain’s capabilities. Here, we propose the initial design of a distributed and modular bio-inspired control architecture that aims to artificially replicate the CNS areas involved in planning and executing voluntary movements (figure 1). The distribution of the architecture is based on the “divide and conquer” concept, where the whole system is decomposed into separated and specialized components. The modularity refers to the independence of each component and its interdependence to the other structures of the architecture. The malfunctioning of each module only affects its contribution to the system and not to the operating state of the other modules. The design of the architecture is specific for gross motor skills involved in the coordination of a robotic arm during the interaction with an external system, such as dynamical target reaching and object manipulation. The control system will be tested mainly on virtual robots in the physical simulation environment offered by the Neurorobotics Platform (NRP) (Falotico et al., 2017). The NRP not only includes a variety of robot and environments, but also a detailed physics simulator. The architecture follows the guidelines from different studies (Caligiore et al., 2017; Houk and Wise, 1995; Casellato et al., 2012; Tomita and Yano, 37 2007; Ryczko et al., 2016; Santos and Matos, 2011) and includes CNS regions such as the brain stem (action regulation), the cerebellum (motor adaptation), the spinal cord (motor pattern generation), the basal ganglia (action selection), and the motor cortex(Initiation, planning, procedure of motion ). The CNS areas will be modeled combining classical control and robotics methods together with bio-inspired AI techniques. This study does not only aim to artificially mimic the connectivity and functionality of the CNS (as seen in previous studies (Floreano et al., 2014; Prescott et al., 2016; Mitchinson and Prescott, 2016)), but to also analyze, with practical evidence, how different brain regions map context-sensitive motor skills as proposed by Wolpert and Kawato (Wolpert and Kawato, 1998). This is because we believe that the modularity of each brain region is fundamental for the extrapolation of valuable information from heterogeneous dynamical stimuli. This extrapolation could facilitate the motor prediction and adaptation in changing or unknown conditions. Among these CNS regions, it is well known the pivotal role of the cerebellum in motor learning and adaptation (Ito, 2008; Dean et al., 2010; Verduzco- Flores and O’Reilly, 2015; D’Angelo, 2014). Several robots have been already endowed with cerebellar-like control architectures with promising results (Garrido Alcazar et al., 2013; Tolu et al., 2012, 2013; Vannucci et al., 2016; Casellato et al., 2015). However, these studies mostly focused on the functionality of a specific CNS region, keeping the contribution and dependency with other brain structures neglectable. Moreover, the experiments were run in simplified conditions with marginal dynamics, absent interaction with the environment, and relative goal, i.e., goal not related to an external reference or exteroceptive sensors. Our investigation will firstly focus on the cerebellum. The way the cerebellum maps and processes the sensory information in relation to the execution of complex dynamical tasks is not totally clear. We assume that an answer could be found in the regular and modular structure of the cerebellum, where distinct functional units can be observed (Ruigrok, 2011). In 2006, Ito claimed that in each unit a forward or an inverse internal model is captured for representing the relation between action and outcome (Ito, 2006). In addition to Ito’s internal models theory, there is also evidence that the human cerebellum can be modeled by a combination of both inverse and forward internal models (Wolpert and Kawato, 1998). Nonetheless, this mixed model has not widely been used in robotic control in particular when the characteristics of the robot and/or the environment change. The secondary aspect to be investigated is the reciprocal interaction between the cerebellum and other CNS areas (Houk and Wise, 1995). The cerebellum will be integrated in the distributed architecture shown in figure 1. Starting from the theory that the cerebellum is decoupled into sub-units, we are going to analyze how the specialization of each unit and their cooperation influences the mapping of heterogeneous dynamical information onto motor skills. From this analysis, we expect to comprehend how the malfunctioning of a specific unit can influence the final corrective action of the cerebellum. At the same time, this could help to understand which feature is not mapped correctly inside the internal model and consequentially correct this lack. Thereafter, from a macro-level perspective, we will investigate how the learned experience is exchanged and utilized across different CNS regions for planning and executing context-related motor commands. This study could give new guidelines for modeling a more robust and distributed robotic control architecture. As matter of fact, the CNS demonstrated that the malfunctioning of one system component does not preclude the operating state of the whole architecture, which is a beneficial aspect for modern autonomous robot. On the other hand, the application of neuro-scientific assumptions on practical experiments could give a feedback and open new lines of research. To conclude, the outcome of the present investigation will provide the state-of the-art for more complex bio-inspired control architectures for neuro-robots that learn from experiences under varying dynamical conditions.
... Often, these neurorobotic systems (Krichmar, 2018) do not recreate all the internal mechanisms of a biological system, but instead try to mimic them at a functional level. Such models are explored for tasks such as control of robotic arms (Bakkum et al., 2007;Casellato et al., 2012), navigation and planning (Cuperlier et al., 2007). Platforms which provide an environment to test the behavior of neuromotor models have also been developed (Goodman et al., 2007;Cofer et al., 2010;Falotico et al., 2017). ...
Article
Full-text available
Decades of research on neuromotor circuits and systems has provided valuable information on neuronal control of movement. Computational models of several elements of the neuromotor system have been developed at various scales, from sub-cellular to system. While several small models abound, their structured integration is the key to building larger and more biologically realistic models which can predict the behavior of the system in different scenarios. This effort calls for integration of elements across neuroscience and musculoskeletal biomechanics. There is also a need for development of methods and tools for structured integration that yield larger in silico models demonstrating a set of desired system responses. We take a small step in this direction with the NEUROmotor integration and Design (NEUROiD) platform. NEUROiD helps integrate results from motor systems anatomy, physiology, and biomechanics into an integrated neuromotor system model. Simulation and visualization of the model across multiple scales is supported. Standard electrophysiological operations such as slicing, current injection, recording of membrane potential, and local field potential are part of NEUROiD. The platform allows traceability of model parameters to primary literature. We illustrate the power and utility of NEUROiD by building a simple ankle model and its controlling neural circuitry by curating a set of published components. NEUROiD allows researchers to utilize remote high-performance computers for simulation, while controlling the model using a web browser.
... We simulated the perturbed reaching task using a simplified model of the upper limb. The model consisted of an anthropometric arm moving on a horizontal plane, with 2 degrees of freedom representing elbow and shoulder [13]. Starting from the desired cartesian trajectory of the end effector, we offline computed the desired joint kinematics and feedforward torques. ...
Conference Paper
The fundamental role of the cerebellum in motor learning explains the deficits of cerebellar patients in adaptation to a changing environment. For example, lesions to the cerebellar cortex compromise performance during tasks like reaching a target under a force field perturbation. However, the exact relationship between neural damages and misbehaviors still needs to be clarified. To this aim, it could become a turning point to exploit a bio-inspired cerebellar model able to simulate multiple tasks in closed-loop, under physiological and different pathological conditions. In the present study, we used a well-established Spiking Neural Network representing a cerebellar microcomplex to reproduce alterations in a perturbed reaching task, after lesions to the neural population in the cerebellar cortex. Following a multiscale approach, we explored different amounts of damage and analyzed the modified behavior, matching the results of a literature reference study. Then, we could make predictions about the underlying altered neural activity, showing the neural causes of high-level impairments. The results demonstrate the generalization capabilities of the model, extending previous studies on different lesions and tasks. We showed the strong potentialities of computational neuroscience in investigating cerebellar diseases through a non-invasive approach, allowing to isolate damages, test multiple configurations, and suggest treatments thanks to a deeper understanding of pathologies.
... Adaptive filter model [12] developed by Fujita was extended the existing work by introducing a 'phase lead-lag compensator with learning capability' which performs a filtering action. Another advancement was an 'Error mapping controller' [13] which have used concepts of Artificial Neural 978-1-5090-6367-3/17/$31.00 ©2017 IEEE Network (ANN) for neuroprosthesis and looked into simple tasks like knee flexion which was further extended into kinematic prediction for robotic arm with cerebellar learning [14]. Later in 1987, Kawato proposed a computation model for controlling voluntary movement [15] which looked at the modularity to the cerebellum to bring out a concept of "multiple paired forward inverse model". ...
... One brain structure extensively studied and modeled is the cerebellum [3], which is involved in cognition [4], [5], [6] and in motor learning and coordination [7], [8], [9], [3], [10]. Computational models of the cerebellum have been successfully employed in motor control of inverted pendulums [2], [11], robotic arms [12], [13], wheeled robots [14], humanoid robots [1], [15], [16], among other engineering applications [17]. ...
Conference Paper
Full-text available
Abstract—Development of computational models of the brain is relevant not only for deepening our understanding of the biological system but also for potential applications to various engineering problems. In this paper the implementation of a bi-hemispherical neuronal network model of the cerebellum (biCNN) in a stand-alone, portable real time (RT) device is presented. The biCNN is tested during a control engineering application, namely, control of a highly unstable two-wheel balancing robot. The RT device considered is the National In- struments myRIO-1900, which provides flexibility and portability to the biCNN. Execution times obtained with the RT device are compared with a personal computer implementation as reference. The results demonstrate the suitability of the RT implementation of the biCNN for robot control, and provide a successful bridge between the cerebellar research and engineering.
Article
Biologically inspired neural mechanisms, coupling internal models and adaptive modules, can be an effective way of constructing a control system that exhibits a human-like behaviour. A brain-inspired controller has been developed, embedding a cerebellum-like adaptive module based on neurophysiological plasticity mechanisms. It has been tested as controller of an ad-hoc developed neurorobot, integrating a 3 degrees of freedom serial robotic arm with a motion tracking system. The learning skills have been tried out, designing a vestibular-ocular reflex (VOR) protocol. One robot joint was used to get the desired head turn, while another joint displacement corresponded to the eye motion, which was controlled by the cerebellar model output, used as joint torque. Along task repetitions, the cerebellum was able to produce an anticipatory eye displacement, which accurately compensated the head turn in order to keep on fixing the environmental object. Multiple tests have been implemented, pairing different head turn with object motion. The gaze error and the cerebellum output were quantified. The VOR was accurately tuned thanks to the cerebellum plasticity. The next steps will include the activation of multiple plasticity sites evaluating the real platform behaviour in different sensorimotor tasks.
Article
Full-text available
The cerebellum is involved in learning and memory of sensory motor skills. However, the way this process takes place in local microcircuits is still unclear. The initial proposal, casted into the Motor Learning Theory, suggested that learning had to occur at the parallel fiber-Purkinje cell synapse under supervision of climbing fibers. However, the uniqueness of this mechanism has been questioned, and multiple forms of long-term plasticity have been revealed at various locations in the cerebellar circuit, including synapses and neurons in the granular layer, molecular layer and deep-cerebellar nuclei. At present, more than 15 forms of plasticity have been reported. There has been a long debate on which plasticity is more relevant to specific aspects of learning, but this question turned out to be hard to answer using physiological analysis alone. Recent experiments and models making use of closed-loop robotic simulations are revealing a radically new view: one single form of plasticity is insufficient, while altogether, the different forms of plasticity can explain the multiplicity of properties characterizing cerebellar learning. These include multi-rate acquisition and extinction, reversibility, self-scalability, and generalization. Moreover, when the circuit embeds multiple forms of plasticity, it can easily cope with multiple behaviors endowing therefore the cerebellum with the properties needed to operate as an effective generalized forward controller.
Article
Full-text available
Cited By (since 1996):209, Export Date: 27 November 2013, Source: Scopus
Article
Full-text available
This study focuses on the role of the motor cortex, the spinal cord and the cerebellum in the dynamics stage of the control of arm movement. Currently, two classes of models have been proposed for the neural control of movements, namely the virtual trajectory control hypothesis and the acquisition of internal models of the motor apparatus hypothesis. In the present study, we expand the virtual trajectory model to whole arm reaching movements. This expanded model accurately reproduced slow movements, but faster reaching movements deviated significantly from the planned trajectories, indicating that for fast movements, this model was not sufficient. These results led us to propose a new distributed functional model consistent with behavioural, anatomical and neurophysiological data, which takes into account arm muscles, spinal cord, motor cortex and cerebellum and is consistent with the view that the central nervous system acquires a distributed inverse dynamics model of the arm. Previous studies indicated that the cerebellum compensates for the interaction forces that arise during reaching movements. We show here how the cerebellum may increase the accuracy of reaching movements by compensating for the interaction torques by learning a portion of an inverse dynamics model that refines a basic inverse model in the motor cortex and spinal cord.
Article
Figure A.1 Walking Trial—Marker Locations and Mass and Frame Rate Information Table A.1 Raw Coordinate Data (cm) Table A.2(a) Filtered Marker Kinematics—Rib Cage and Greater Trochanter (Hip) Table A.2(b) Filtered Marker Kinematics—Femoral Lateral Epicondyle (Knee) and Head of Fibula Table A.2(c) Filtered Marker Kinematics—Lateral Malleolus (Ankle) and Heel Table A.2(d) Filtered Marker Kinematics—Fifth Metatarsal and Toe Table A.3(a) Linear and Angular Kinematics—Foot Table A.3(b) Linear and Angular Kinematics—Leg Table A.3(c) Linear and Angular Kinematics—Thigh Table A.3(d) Linear and Angular Kinematics—½ HAT Table A.4 Relative Joint Angular Kinematics—Ankle, Knee, and Hip Table A.5(a) Reaction Forces and Moments of Force—Ankle and Knee Table A.5(b) Reaction Forces and Moments of Force—Hip Table A.6 Segment Potential, Kinetic, and Total Energies—Foot, Leg, Thigh, and ½ HAT Table A.7 Power Generation/Absorption and Transfer—Ankle, Knee, and Hip
Article
The cerebellum is essential for the control of multijoint movements; when the cerebellum is lesioned, the performance error is more than the summed errors produced by single joints. In the companion paper ( Schweighofer et al. 1998 ), a functional anatomical model for visually guided arm movement was proposed. The model comprised a basic feedforward/feedback controller with realistic transmission delays and was connected to a two-link, six-muscle, planar arm. In the present study, we examined the role of the cerebellum in reaching movements by embedding a novel, detailed cerebellar neural network in this functional control model. We could derive realistic cerebellar inputs and the role of the cerebellum in learning to control the arm was assessed. This cerebellar network learned the part of the inverse dynamics of the arm not provided by the basic feedforward/feedback controller. Despite realistically low inferior olive firing rates and noisy mossy fibre inputs, the model could reduce the error between intended and planned movements. The responses of the different cell groups were comparable to those of biological cell groups. In particular, the modelled Purkinje cells exhibited directional tuning after learning and the parallel fibres, due to their length, provide Purkinje cells with the input required for this coordination task. The inferior olive responses contained two different components; the earlier response, locked to movement onset, was always present and the later response disappeared after learning. These results support the theory that the cerebellum is involved in motor learning.
Article
A comprehensive theory of cerebellar function is presented, which ties together the known anatomy and physiology of the cerebellum into a pattern-recognition data processing system. The cerebellum is postulated to be functionally and structurally equivalent to a modification of the classical Perceptron pattern-classification device. It is suggested that the mossy fiber → granule cell → Golgi cell input network performs an expansion recoding that enhances the pattern-discrimination capacity and learning speed of the cerebellar Purkinje response cells.Parallel fiber synapses of the dendritic spines of Purkinje cells, basket cells, and stellate cells are all postulated to be specifically variable in response to climbing fiber activity. It is argued that this variability is the mechanism of pattern storage. It is demonstrated that, in order for the learning process to be stable, pattern storage must be accomplished principally by weakening synaptic weights rather than by strengthening them.