You are on page 1of 23

Biorobotic Experiments for the Discovery of Biological Mechanisms*

Author(s): Edoardo Datteri and Guglielmo Tamburrini


Source: Philosophy of Science , Vol. 74, No. 3 (July 2007), pp. 409-430
Published by: The University of Chicago Press on behalf of the Philosophy of Science
Association
Stable URL: https://www.jstor.org/stable/10.1086/522095

JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide
range of content in a trusted digital archive. We use information technology and tools to increase productivity and
facilitate new forms of scholarship. For more information about JSTOR, please contact support@jstor.org.

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at
https://about.jstor.org/terms

The University of Chicago Press and Philosophy of Science Association are collaborating with
JSTOR to digitize, preserve and extend access to Philosophy of Science

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
Biorobotic Experiments for the
Discovery of Biological Mechanisms*
Edoardo Datteri and Guglielmo Tamburrini†‡

Robots are being extensively used for the purpose of discovering and testing empirical
hypotheses about biological sensorimotor mechanisms. We examine here methodolog-
ical problems that have to be addressed in order to design and perform “good” ex-
periments with these machine models. These problems notably concern the mapping
of biological mechanism descriptions into robotic mechanism descriptions; the dis-
tinction between theoretically unconstrained “implementation details” and robotic fea-
tures that carry a modeling weight; the role of preliminary calibration experiments;
the monitoring of experimental environments for disturbing factors that affect both
modeling features and theoretically unconstrained implementation details of robots.
Various assumptions that are gradually introduced in the process of setting up and
performing these robotic experiments become integral parts of the background hy-
potheses that are needed to bring experimental observations to bear on biological
mechanism descriptions.

1. Introduction. Robots are being extensively used for the purpose of


discovering and testing hypotheses about biological sensorimotor behav-
iors. This use of robots is the distinguishing feature of what is now being
called “biorobotics” (Webb 2001, 2006; Webb and Consi 2001). Significant
examples of biorobotic inquiries include the investigation of phonotaxis
in crickets (Webb 2002), chemotaxis in lobsters (Grasso et al. 2000), ant
homing capacities (Lambrinos et al. 2000), and hippocampal navigation
in rats (Burgess et al. 1997). Biorobotic approaches are adopted to study

*Received December 2005; revised January 2007.


†To contact the authors, please write to: Edoardo Datteri, Dipartimento di Scienze
Umane, Università degli Studi Milano-Bicocca, Piazza dell’Ateneo Nuovo 1, Milano
20126, Italy; e-mail: edoardo.datteri@unimib.it, or to Guglielmo Tamburrini, Dipar-
timento di Scienze fisiche, Università di Napoli Federico II, Compl. Univ. Monte S.
Angelo, Via Cintia, Napoli 80126, Italy; e-mail: tamburrini@na.infn.it.
‡ Both authors wish to thank Barbara Webb and an anonymous referee for stimulating
comments and criticisms of an earlier version of this paper. We are grateful to Franco
Giorgi, Hykel Hosni, and Massimo Mugnai for engaging discussions on the meth-
odological problems addressed here.
Philosophy of Science, 74 (July 2007): 409–430. 0031-8248/2007/7403-0004$10.00
Copyright 2007 by the Philosophy of Science Association. All rights reserved.

409

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
410 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

human sensorimotor systems too: arm posture maintenance is investigated


by implementing the hypothesized neural control mechanism onto an
artificial arm equipped with a robotic muscle spindle (Chou and Han-
naford 1997; Jaax and Hannaford 2002); primate optomotor systems are
investigated by means of VLSI implementations (Horiuchi and Koch
1996); the role of basal ganglia in action selection (Prescott et al. 2002),
characteristic patterns of human arm trajectories (Schaal and Sternad
2001), cerebellar control (Eskiizmirliler et al. 2002), and visuo-motor co-
ordination (Dario et al. 2002) are being explored on similar grounds. All
of these inquiries involve experimental comparisons between biological
system and robotic system behaviors.
Biorobotics shares with artificial intelligence and so-called “biologically-
inspired” robotics the goal of achieving adaptive behaviors in machines.
Biologically inspired robotics adopts solutions to robotic control problems
suggested by models of adaptive biological behaviors only insofar as these
solutions enable one to improve the adaptivity of service robots (Beer et
al. 1997). Biorobotics has to meet more stringent methodological de-
mands, insofar as robots serve there the special purpose of testing or
discovering theoretical models of biological sensorimotor behaviors. The
potential contribution of this research strategy to the scientific under-
standing of biological systems was emphasized as early as the beginning
of last century, when American biologist Jacques Loeb claimed that newly
constructed phototropic machines provided empirical support for his
model of phototropic behavior in moths (Loeb 1918; Cordeschi 2002).
More widely acknowledged ancestors of biorobotics include pioneers of
Cybernetics Arturo Rosenblueth, Norbert Wiener, and John Bigelow. In
articles published in this journal (Rosenblueth, Wiener, and Bigelow 1943;
Rosenblueth and Wiener 1945), these scientists conjectured that a variety
of goal-directed behaviors observed in biological systems are governed by
negative feedback control mechanisms, and suggested on these grounds
the opportunity of using feedback controlled machines as material models
of biological systems: these material models would enable one to generate
empirical predictions that are difficult to work out on the basis of the-
oretical models only, and to make interventions that are difficult or uneth-
ical to perform on biological systems.1

1. The pioneers of Cybernetics envisaged experiments with machine models in con-


nection with their broad research programme of identifying unified mechanisms for
goal-directed behaviors. These were taken to span an amoeba’s reaction to chemical
gradients, a mouse chasing cat, and Mary taking a glass of water to her mouth.
Methodological reflections on machine experiments concerning this vast family of
behaviors are bound to run into the vexing problems of intentionality in psychology.
See, in this connection, the early critical comments to Rosenblueth, Wiener, and Bigelow
(1943) published in this journal (Taylor 1950). Contemporary biorobotic inquiries into

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 411

Distinctive methodological problems arise in the process of designing


and performing experiments with machine models. The methodology of
machine experiments is examined here in connection with contemporary
biorobotic investigations of biological sensorimotor mechanisms. In these
investigations one has to address the problem of constructing or licensing
the use of robots for experimental purposes. More specifically, one has
to map biological mechanism descriptions into robotic mechanism de-
scriptions, through the intermediary of some “mechanism schema” (Ma-
chamer, Darden, and Craver 2000) specifying abstract role functions that
are filled in by means of component parts and processes of biological and
robotic systems, respectively. This mapping enables one to discriminate
features of a biorobot that carry a modeling weight from theoretically
unconstrained “implementation details.” In addition to this, calibration
experiments are often needed to tune robot behavior so as to meet mech-
anism schema requirements or else to achieve working implementations
from underspecified mechanism schemata. And experimental environ-
ments have to be carefully monitored for disturbing factors that affect
both modeling features and implementation details of biorobots.
The assumptions that are gradually introduced in the process of setting
up and performing biorobotic experiments become integral parts of the
background hypotheses that are needed to bring experimental observa-
tions to bear on biological mechanism descriptions. Distinctive back-
ground assumptions in biorobotics concern implementation accuracy with
respect to mechanism schemata and descriptions. Genuine examples of
biorobotic research are selectively introduced here to illustrate the meth-
odological complexities arising in the preparation of “good” biorobotic
experiments or the evaluation of their impact on empirical hypotheses
about biological sensorimotor mechanisms.

2. Theoretical Models and Biorobotic Experiments. We take here mech-


anisms to be “entities and activities organized such that they are pro-
ductive of regular changes from start or set-up to finish or terminating

the adaptive and intelligent behaviors of relatively simple biological systems offer the
valuable opportunity of analyzing the methodology of machine experiments in isolation
from problems of intentionality in psychology (Tamburrini and Datteri 2005). In this
respect, contemporary biorobotics is on a par with Marr’s computational investigations
of human vision (Marr 1982) or the more recent modeling of human sensorimotor
coordination described in Berthoz (1997). To the extent that problems of intentionality
do not arise in biorobotics, one need not be concerned here with arguments questioning
the possibility of providing mechanistic accounts of intentional behavior (Von Eckardt
and Poland 2004).

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
412 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

conditions” (Machamer, Darden, and Craver 2000, 3).2 Accordingly,


mechanism descriptions include both temporal and spatial constraints.
Temporal constraints concern order, rates, duration, and frequency of
activities; spatial constraints concern structure and orientation of entities,
in addition to compartmentalization, localization, and interconnection of
activities (Craver and Darden 2001). Mechanism descriptions can be or-
ganized into multi-level hierarchies, so that some mechanisms figure as
modular components of higher-level mechanisms. Mechanism schemata
are more sparing than mechanism descriptions, insofar as no reference to
specific material systems is made there. Role functions mentioned in mech-
anism schemata can be “filled with descriptions of known component
parts and activities.” In particular, “mechanism schemata can be instan-
tiated in biological wet-ware . . . or represented in the hardware of a
machine” (Machamer, Darden, and Craver 2000, 17).3
The discovery of biological sensorimotor mechanisms starts with the
characterization of the phenomenon of interest (Craver and Darden 2001,
121). And the initial stages in the discovery of biological mechanisms
often involve the formulation of mechanism sketches, which allow for
functional gaps (Machamer, Darden, and Craver 2000; Craver and Dar-
den 2001) or imprecise specification of functional roles. Mechanistic
sketches are progressively worked out and revised until a mechanism
description is obtained. Underspecification of mechanism sketches is the
source of arbitrary choices in biorobot design. In order to set up mean-
ingful biorobotic experiments, one has to understand the contribution of
these arbitrary choices on the shaping of overall biorobot behavior and
to preserve implementation accuracy with respect to mechanism sketches
or schemata.
Mechanism schemata may admit a variety of machine instantiations
and implementations. Computer simulation of mechanism schemata, for
example, is pursued to support both prediction and explanation (Amit
1998) from mechanism descriptions insofar as, by running the simulation,

2. Related attempts to characterize the notion of mechanism and the nature of mech-
anistic explanation were pursued by Bechtel and Richardson (1993), Bechtel and Abra-
hamsen (2005), Glennan (1996, 2002, 2005), Woodward 2002), and Tabery (2004).
Craver (2001) attempts to reconcile mechanistic and functional modeling, as the latter
is explicated by Cummins (1975). Analysis of the nature of mechanistic explanation
in biology (Craver and Darden 2005) is beyond the scope of this article, which is chiefly
concerned with methodological problems arising in biorobotic approaches to the dis-
covery and testing of empirical hypotheses about biological sensorimotor mechanisms.
3. The Cybernetic proposal of identifying control functions shared by adaptive bio-
logical systems and machines was guided by the assumption that “it is possible to
obtain the same output for the same input with different physical structures” (Rosen-
blueth and Wiener 1945, 318).

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 413

one achieves a seamless account of how “termination conditions are pro-


duced by the set-up conditions and intermediate stages” (Machamer, Dar-
den, and Craver 2000, 3), one identifies some component contribution to
the regular functioning of the overall mechanism, and predicts what be-
havioral changes manipulative interventions give rise to in accordance
with the regularities regimenting interactions between mechanism com-
ponents (Woodward 2002).
Biorobots are built starting from mechanism schemata or sketches to
support scientists much in the way that computer simulations do. Robotic
implementations, however, enable one to sidestep major modeling prob-
lems that arise in computer simulations of mechanism schemata. In a
computer simulation, one has to represent environmental factors that
release, sustain or modulate activities of mechanism components. In con-
trast with this, the need for environmental simulation does not arise in
biorobotic investigations, insofar as robotic and biological systems are
behaviorally compared to each other in real world environments (typically
in the ecological niche of the biological system under investigation). Thus,
biorobotic experimental practice does not give rise to the methodological
problem of controlling whether behavioral (dis-)similarities between target
biological systems and computer-simulated agents take their origin in ad
hoc or inaccurate simulations of the environment. One should be careful
to note, however, that there is a methodological trade-off between com-
puter simulation and robotic implementation of mechanism schemata.
Computer simulations are unaffected by methodological problems arising
in biorobotics on account of the fact that biorobots are immersed in
natural environments. Indeed, biorobot malfunctioning may be induced
by environmental features (temperature, humidity, illumination, and so
on) that are perfectly normal for the biological system under investigation.
Accordingly, biorobotic experimental settings have to be monitored for
two different sets of boundary conditions, with the aim of preserving
“normal” functioning conditions for both biological and robotic systems
there.
These preliminary observations are sufficient to bring out some dis-
tinctive features of biorobotic approaches to the investigation of biological
sensorimotor mechanisms. The implementation of a biorobot presupposes
that the role functions mentioned in some mechanism sketch or schema
M admit both a biological instantiation MS and a robotic instantiation
MR. The robotic instance MR, which makes reference to robotic parts and
activities, is used to guide the implementation of some robotic system R.
Thus, M is the theoretical basis licensing similarity judgments between
MR, a biological instance of M, and the implementations of both. Ex-
periments involve behavioral comparisons between R and a biological
system S. The latter is hypothesized to implement some biological instance

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
414 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

of M. The background assumptions that are needed to bring these ex-


periments to bear on this hypothesis include specific biorobotic assump-
tions about R’s design and implementation process.
There are various ways of bringing biorobotic experiments to bear on
mechanism descriptions and sketches. In addition to providing data for
model corroboration or falsification, these experiments may enable one
to identify previously undetected functional components in the target bi-
ological system, to identify causally relevant features of the environment,
and to suggest specific expansions of mechanism sketches. Biorobotic
experiments, which invariably involve running a biorobot and observing
its behavior, may differ from each other along more specific dimensions.
One can observe overt motor behavior (such as the robot trajectory), the
behavior of internal components (such as the firing rate of artificial neu-
rons embedded in the robot control system), or environmental properties
(such as the spatial distribution of chemical gradients to which the robotic
system reacts or physical modifications of the environment induced by
robot actions). In preparing an experiment, one can manipulate both
environmental conditions (by occluding, say, sensory cues or modifying
illumination conditions) and internal variables (by modifying, say, pa-
rameters in system modules). Thus, biorobotic experiments afford the
opportunity of comparing robotic and biological system behaviors in vary-
ing external or internal conditions. Distinctive methodological problems
this experimental practice gives rise to are analyzed in the following sec-
tions, chiefly by reference to biorobotic inquiries on adaptive rat navi-
gation and lobster chemotaxis.

3. Initial Corroborations and New Biological Mechanism Predictions. The


capability of rats to learn the spatial structure of mazes, and to reach
previously encountered goal locations, such as food sites, has been ex-
tensively investigated in the past decades. In the wake of Tolman (1948),
O’Keefe and Nadel (1978) proposed to explain these capabilities by ref-
erence to some sort of internal map of the environment which is acquired
by learning. And empirical evidence emerged for the hypothesis that the
hippocampus—a bilateral structure placed between the cerebral cortex
and the thalamus—plays a key functional role in spatial tasks. Notably,
rats with hippocampal lesions are unable to perform a variety of spatial
tasks (Barnes 1988; Jarrad 1993); and neurons in areas CA3 and CA1 of
the hippocampus of freely moving rats—called “place cells”—have been
found to fire maximally when rats reach specific locations of the testing
platform (O’Keefe and Dostrovsky 1971; Fox and Ranck 1975). On ac-

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 415

count of this finding, place cells were hypothesized to support rat self-
localization and heading capabilities.4
Empirical law-like correlations between place cells firing and rat po-
sition enable one to predict rat position on the basis of place cells firing
(Wilson and McNaughton 1993). But in order to provide a mechanistic
explanation of rat spatial navigation which takes into account place cells
behavior, one has to regard place cells as elements of a complex of in-
teracting components which comprise perceptual systems and effectors.
How is it that place cells firing drive rat movements? How is it that sensory
stimuli determine place cells firing? And how are hippocampal maps ac-
quired by learning? These questions point to crucial gaps in our under-
standing of the mechanisms producing rat navigation behaviors and the
role of place cells therein. In order to fill in some of these gaps (Burgess
et al. 1997), Burgess et al. (2000) proposed a more comprehensive mech-
anistic model (let us call it MS) which we turn to summarize now.
MS postulates components formed by various neural maps with Heb-
bian connections, that are modifiable by unsupervised learning processes.
According to MS, a ‘sensory’ map is organized into several ‘sensory cells
groups’, each one of them encoding distances from one of the walls sur-
rounding the environment. An ‘entorhinal’ map encodes the distance be-
tween the rat and two adjacent walls (each entorhinal cell being connected
with two sensory cells of distinct groups). Entorhinal cells are connected
with place cells. The receptive field of the former is sharper than the
receptive field of the latter.5 The place cell layer is connected with a ‘goal
cell’ layer. Each goal location is represented there by a set of goal cells;
and each element of this set is tuned so as to respond maximally when
the system is displaced from the goal into a specific direction. Assuming
that four goal cells only encode for a goal location, GCN fires when the
rat is north of the goal, GCS fires when the rat is south of the goal, and
so on. The firing rate of each goal cell is proportional to the displacement
of the system from the goal in the corresponding direction. The displace-

4. The behavior of place cells has been extensively investigated. Place cells seem to be
responsive to visual stimuli at or beyond the limits of the rat’s reachable environment,
but their behavior is unaffected by objects placed centrally in the environment; addi-
tional experiments provide evidence for some sort of “internal compass’ that anchors
place cell activity to spatial locations in the absence of visual cues (Muller and Kubie
1987; Cressant et al. 1997; Jeffery et al. 1997); and the limits of validity of law-like
correlations between place cells firing and rat position have been explored too (O’Keefe
and Dostrovsky 1971). The firing of place cells seem to depend on the phase of the
EEG theta rhythm too (O’Keefe and Recce 1993; Cacucci et al. 2004). Theta modu-
lation was replicated in the robotic system described in this section. See Burgess et al.
(2000, Section 2.1) for details.
5. See Burgess et al. (1994) for discussion.

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
416 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

ment and the heading of the system is given by the vector sum of the
directions represented by each goal cell, weighted by their firing rate.
Learning proceeds in a feedforward fashion: while the rat is wandering,
an unsupervised competitive learning algorithm modifies the connections
between entorhinal and place cells, thus creating a place cell map of the
environment; one-shot strengthening of synaptic connections between goal
cells, ‘head direction cells’ (signaling system heading), and active place
cells obtains whenever a reward signal (indicating the presence of the
goal) is received. After learning, the firing of goal cells will depend on rat
position (signaled by active place cells) and heading. Movements are issued
in the directions coded by goal cell vector representations.
This mechanistic model fills in gaps of mechanistic sketches which make
reference to place cells and their functional roles. In particular, this mech-
anism description specifies subsystems which generate heading (on the
basis of place cells outputs), determine place cell firing (on the basis of
sensory cell representation of distances from walls), and provide an ac-
count of hippocampal map learning. Burgess and O’Keefe tested this
mechanism for accurate learning of the environment and successful nav-
igation towards previously learnt goals by robotic simulation experiments.
These were carried out on a robotic system R consisting of a small wheeled
platform, with an on-board camera, infrared proximity detectors, and a
neural network which includes 60 sensory cells, 900 entorhinal cells, 900
place cells, and four goal cells representing a single goal position. The
experimental environment is a rectangular arena formed by white walls
and a dark floor, the north wall being signaled by a distinctive mark. At
each processing cycle, the robot rotates in the estimated north, south, east,
and west directions, and captures images of the four walls; these images
are processed by a black and white edge detection algorithm, which reveals
corners between walls and floor; corners are used to evaluate the distance
of the system from the walls; by separately controlling the velocities of
the two motors, the robot turns in the direction computed by the neural
network and moves 3 cm forward.
In experimental trials, the robotic rat learnt successfully an accurate
map of the environment, and was able to reach previously learnt goal
locations. Initially, the robot was set free to explore (by wandering) a
50#50 cm square environment, according to the mechanism described
above. Robot performance after exploration was evaluated in the same
environment, and in a “stretched” 50#75 cm environment. In additional
trials, the activity of artificial neurons was recorded during navigation in
the stretched environment. While entorhinal cells maintained their peaks
at fixed distances from the walls, the receptive field of some place cells
became stretched and occasionally two distinct peaks were recorded. No-
tably, this behavior of artificial place cells in stretched environments was

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 417

found to match the behavior of biological place cells in similar conditions


(O’Keefe and Burgess 1996).
The behavioral match between actual rats S and robotic rats R in similar
environments was evaluated chiefly in terms of goal reaching.6 The ex-
perimental outcomes were taken to provide initial corroboration for MS
and the mechanistic explanation of goal reaching by rats which flows from
it (Burgess et al. 2000). Thus, data about robotic behavior were used as
evidence for a theoretical hypothesis about a biological mechanism.
Clearly, this corroboration claim, which is made on the basis of an ob-
served behavioral match between R and S, presupposes that robotic rat
R is an accurate implementation of mechanism description MR and that
MR is an instance of some abstract mechanistic schema M which admits
MS as biological instance.
Predictions which MS makes on the receptive fields of sensory, entor-
hinal, place, and head cells have been corroborated by purely biological
experiments. It is worth noting, in this respect, that biorobotic inquiries
can motivate biological experiments for identifying biological entities and
activities postulated by some mechanistic model. This additional role for
biorobotic inquiry is exemplified by reference to the ‘goal cell’ layer of
MS. The robotic analog of this layer accounts for the capability of robotic
rats to navigate towards goal sites that are not signaled by sensory cues.
In additional biorobotic measurements performed in the stretched envi-
ronment, the goal cells representation was found to be stretched as well,
and to affect navigation on this account: “the principal effect of expanding
an environment is a separation of the peaks of each goal cell’s firing-rate
map (the locus of search remaining between them). By contrast, con-
tracting an environment by a large enough factor can cause the locations
of peak firing of opposing goal cells to cross over, and produces a more
dramatic effect: the robot searches only at the edges of the environment”
(Burgess et al. 1997, 1541). These predictions on the behavior of goal cells
may guide the search for cells exhibiting a similar behavior in the rat
nervous system; and the ‘goal cells’ hypothesis may be tested by observing
search patterns of rats in stretched environments, and comparing them
to predictions obtained by means of biorobotic systems.
In concluding this section, let us note that new components and role
functions of rat navigation mechanisms have been recently hypothesized
on account of the discovery of grid cells in the dorsocaudal medial en-
torhinal cortex of rats. Grid cells might support some spatial computations
that were previously attributed to the hippocampus (Hafting et al. 2005).

6. It is worth noting that this conclusion hinges on the criteria adopted for behavioral
comparison, for R requires far longer navigation times than S, and only a coarse match
between their trajectories obtains.

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
418 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

This recent discovery does not affect the broad methodological picture
outlined in this section on the basis of biorobotic inquiries on place cells,
insofar as biorobotic experiments may prove useful to fill in the gaps of
mechanism sketches which include grid cells, to test the behavioral pre-
dictions of the resulting mechanism schemata, and to drive the search for
biological components and activities interacting with grid and place cells.

4. Feedback-Based Hypotheses on Lobster Chemotaxis. Biorobotic ex-


periments were performed by Grasso et al. (2000) to investigate the ca-
pability of lobsters to track turbulent chemical plumes to their sources
(chemotaxis). The mechanism investigated there is based on feedback
analysis: lobster antennae can detect local gradients in the chemical plume;
and local gradients provide lobsters with directional information to lo-
calize the plume source. Computer simulations appear to be unsuitable
to test mechanism schemata that are consistent with this broad picture
of lobster chemotaxis, insofar as accurate replication of plume turbulence
in water is unfeasible (especially in view of the difficulty of developing
accurate fluid-dynamical models at the temporal and spatial scales that
are needed to simulate lobster sensorimotor loops in chemo-orientation
tasks). In contrast with this, robots placed in real chemical plumes allow
for direct comparison between robotic and biological behaviors.
Two feedback-based mechanisms have been implemented in a small
mobile wheeled robotic platform, called RoboLobster, which is provided
with two antennae for chemical gradient detection. By operating the first
mechanism (M1), when gradient exceeds a threshold the system turns
towards the side of the sensor detecting higher concentration or moves
forward when same concentration information is detected by both sensors.
The second mechanism (M2) is obtained from M1 by adding a rule, which
causes the system to back up when the concentration level detected by
both sensors falls below a threshold and until the threshold is exceeded
again. In experimental trials, RoboLobster is immersed in a stream of
water to which a turbulent plume was added. The source of the plume is
immersed too, and the plume assumes a typical oblong shape (replicating
the hydrodynamic output from a filter-feeding bivalve, which is a natural
food item for lobsters). The plume shape is easily replicated in various
trials. Steering is achieved by independently controlling the two wheels.
Several features of RoboLobster motor behavior are monitored while
the robot is searching for the plume source. These include closest approach
(the shortest observed distance between robot and source) and path tor-
tuosity (the angular variance of turns measured over the entire path). In
trials running M1 RoboLobster consistently failed to hit the source (the
closest approach distance was high). Better results were obtained by M2,
except when the robot started 100 cm away from the source. The authors

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 419

called ‘distal patch field’ (DPF) this region of the plume, in which both
algorithms performed poorly. Before claiming falsification of both mech-
anism schemata, the authors of this biorobotic study were careful to test
whether the selected distance between antennae could affect RoboLobster
performance, due to the impossibility of detecting the gradient between
close points in the DPF region. It turned out, however, that changing the
distance between antennae does not improve RoboLobster performance
in DPF. Additional trials were performed by systematically changing ini-
tial orientations of the robot, to control whether successful hits of the
source from the DPF were due to particular starting positions, but no
significant effect of starting orientation on robot performances was found.
Having excluded distance between antennae and starting orientation as
alternative explanations for observed behavioral discrepancies, these bio-
robotic experiments were taken to show that in the DPF, no matter how
initial conditions change, neither mechanism accounts for target system
behavior. In particular, these results were directed against a specific as-
sumption embedded in M1 and M2, which concerns the structure of the
plume in the DPF. The fact that feedback analysis alone is inadequate to
account for chemotaxis in the DPF is taken to suggest that “there is no
concentration gradient information available to either algorithm in the
DPF” (Grasso et al. 2000, 126). A local gradient is instead detectable in
regions that are nearer to the plume source, thus allowing for effective
robotic chemotaxis.
To sum up: The RoboLobster experiments show that biorobotics may
enable one to falsify mechanism schemata and sketches, and to isolate
significant, but previously overlooked features of the environment. (“In
one sense the plume, rather than the algorithms or the American lobster,
is the entity under the closest scrutiny in these experiments” [Grasso et
al. 2000, 127].) In addition to this, the RoboLobster inquiry illustrates
how biorobotic experiments performed after selective manipulations of
environmental or internal features contribute to rule out specific disturb-
ing factors or the influence of theoretically unconstrained options in ex-
perimental settings. These experimental strategies in biorobotic inquiry
supplement those illustrated by reference to rat navigation mechanisms,
whereby biorobotic experiments may enable one to isolate mechanism
schemata accounting for biological behaviors, and drive the search for
biological entities and activities producing target system behaviors.

5. Biorobotics and the Upgrading of Implementation Details. The discov-


ery of mechanisms often proceeds by piecemeal elaboration and revision
of mechanistic sketches (Craver and Darden 2001, 133–134). Biorobotics
can assist the iterative specification of mechanism descriptions from mech-
anism sketches. One starts with a sketch representing an initial ‘guess’ on

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
420 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

target system mechanisms, and progressively identifies role functions and


fillers for gaps or underspecified components of the original sketch. Sen-
sory processing and motor control of rat navigation are underspecified
components in the mechanistic hypothesis proposed by Burgess and
O’Keefe. Their model assumes that the distance between the system and
the walls is encoded in sensory cells, and that the goal cell layer outputs
a vector pointing to the goal. However, the model specifies neither the
mechanistic connections between sensory cells and sensory receptors, nor
the connections between goal layer and motor actuators. In particular,
the model does not specify how it is that visual, auditory, olfactory, and
other sensory stimuli are processed so as to determine sensory cell firing.
Similarly, the model does not specify how rat muscles are activated on
the basis of the vector computed by the goal cell layer. To continue, this
model of rat navigation does not provide an estimate for the population
of sensory cells that represent distances (this parameter is crucial to de-
termine how coarse the encodings of distance are). And unfixed param-
eters are involved in the modeling of sensory neurons too.
The filling in of these mechanistic gaps is achieved in the robotic rat
implementation by adding mechanisms for sensory processing and motor
control that are unconstrained by the initial hypothesis. In particular, the
robot is provided with an artificial camera and with infrared proximity
detectors as sensors, and with two independently driven wheels to move
in the environment. And the mechanism used to estimate distances from
the walls is based on the extraction of the dark-white corner between the
wall and the floor, as well as on geometric calculations outputting the
distance and the orientation of the wall with respect to the system.
Success in replicating biological behaviors may suggest that some role
functions filled in by theoretically unconstrained implementation details
be included in the biological hypothesis. For example, experiments on the
robotic rat may lead one to conjecture that the rat nervous system exploits
light-dark contrast to evaluate distances from the walls. This hypothesis
may be tested by observing the behavior of rats in poorly contrasted
arenas and seeing whether their performance decreases there; and one
may additionally run biorobotic experiments with a modified robotic rat
using, say, an alternative visual processing algorithm and checking
whether the behavioral match between biological system and robot im-
proves on account of this modification. Indeed, alternative goal-seeking
mechanisms were implemented, tested, and the selection between them
was performed by Burgess et al. (2000, 305) on the basis of some behav-
ioral match criteria.
Similar considerations arise in connection with theoretically uncon-
strained implementation features that are added for the sole purpose of
preserving normal working of the robotic system. Suppose that mecha-

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 421

nism description M postulates an open-loop control of actuators which


is consistent with neural recordings of motoneurons; and suppose that
feedback control is required in robotic implementation of M insofar as,
say, rough surfaces in the experimental ground hinder robot movements
and give rise to significant discrepancies between desired and actual move-
ments. Corrective motor commands can be issued by endowing the system
with proprioceptive sensors (e.g., a shaft encoder which detects the actual
steering of the system) and with a feedback-based mechanism. If the
robotic system including this feedback mechanism—which is initially not
postulated in M—successfully replicates biological behaviors, then one
may look for a similar mechanism in the biological system, which includes
proprioceptive sensors and feedback-based inhibiting neurons, and coun-
teracts the effects of rough surfaces on biological system movement. This
experimental possibility is an additional example of a broad biorobotic
strategy for the iterative specification of mechanism descriptions: the role
functions performed by arbitrary implementation details of biorobots are
selectively introduced in mechanism sketches and schemata on account
of the successful replication of biological behaviors. The upgraded model,
in its turn, prompts the search for independent evidence about biological
entities and activities supporting these role functions.

6. Background Assumptions in Biorobotic Experiments. The need for dis-


tinctive background assumptions has been diffusely emphasized in the
above discussion of biorobotic experiments. In addition to multiple in-
stantiation hypotheses for mechanism schemata and normalcy conditions
for robot behavior, these assumptions notably concern implementation
accuracy: if the entities and activities of robotic system R do not fulfill
the constraints set by mechanism description MR , then one can hardly
extract from an analysis of R’s behaviors significant information on the
relationship between target biological system S and biological mechanism
description MS (which is supposed to preserve the role functions of MR).7
The significance that implementation accuracy assumes in biorobotics
becomes particularly evident by contrasting simulation requirements in
biorobotics and biologically-inspired robotics. Analyses of biological so-
lutions to sensorimotor control problems play heuristic roles in biologi-
cally-inspired robotics, but accurate modeling of biological mechanisms
is not pursued when more efficient solutions to engineering problems are
at hand. Consider, for example, biological response delays. Mechanism
descriptions of sensorimotor coordination in biological systems take into

7. Independent testing of background assumptions is deemed to be part of model


corroboration by Lloyd (1987). For discussion of implementation accuracy assumptions
and their testing, see Tamburrini and Datteri (2005, 346–347).

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
422 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

account neural delays occurring in both sensory data and motor command
transmission and processing. Current robotic technology often allows for
faster responses. Thus, a close match between biological and robotic mech-
anism descriptions may require appropriate robotic simulation of neural
delays which, in fact, decrease overall robot performance.8
Accuracy may be compromised in the process of biorobotic implemen-
tation by adaptations of theoretically constrained items or additions of
theoretically unconstrained implementation details. Let’s see.
Adapting theoretically constrained items. Perfect functional mapping of
biological mechanism descriptions onto robotic mechanism descriptions
is a regulative ideal for biorobotic inquiry, which is hardly ever attained
in practice, in view of practical and theoretical impediments. Practical
impediments chiefly originate in limitations of current robotic technology,
insofar as hardware components meeting every constraint imposed by
mechanism descriptions are not available. Consider, for example, sensors
for detecting chemical plumes that are needed for chemotaxis simulation
purposes. As pointed out in Ishida et al. (2001, 222), “sensors for a robot
to detect those signals with capabilities similar to those of animals are
not yet available.” Striking limitations of available sensors concern tem-
poral features of response onset and recovery of initial conditions after
stimulus removal (which may amount to 30s for some sensor technology).
Thus, opting for the use of some particular sensor technology becomes
a theoretically significant decision, to the extent that constraints on sen-
sitivity, response or recovery times of chemical sensors are specified in
mechanism descriptions.
Adaptations are occasionally pursued even when adequate modeling
hardware is available, in order to facilitate implementation and reduce
control problems. Consider, in this connection, the detailed model of hex-
apod locomotion formulated by Quinn and Ritzmann (2001), which takes
into account insect leg kinematics. Since accurate replication of the insect
kinematic chain in robot hardware gives rise to difficult implementation
and control problems, some degrees of freedom in robotic legs were pro-
visionally disregarded. Simulation data obtained on the basis of the sim-
plified implementation were encouraging, insofar as they suggested that
the eliminated degrees of freedom were immaterial to a satisfactory op-
eration of the middle and rear legs. In general, however, the effects of
these adaptations may become unpredictably significant in real-world ex-
perimental settings and under uncontrolled boundary conditions. There-
fore, the background assumptions arising from such simplifying adap-

8. Cases in point are the biorobotic investigation of primate oculo-motor systems


described in Horiuchi and Koch (1996, 261) and the neural simulator described in
Reeve et al. (2005, 44).

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 423

tations have to be isolated and possibly supported by independent


experimental results, in order to make proper use of adapted robots in
biorobotic experimental inquiries.
Selecting theoretically unconstrained items. The rat navigation mecha-
nism proposed by Burgess and O’Keefe does not constrain the choice of
motor actuators in the process of robotic implementation. Viable options,
which do not violate mechanism description constraints, include various
kinds of independently controlled wheels or leg arrays (endowed with
different degrees of freedom). Similar considerations apply to the selection
of values for underspecified parameters in mechanism sketches, possibly
achieved by calibration, as long as this selection does not lead one to
violate mechanism sketch constraints.
Even though no violation of theoretical constraints is introduced in
these cases, one should be careful to note that the overall system behavior
may depend on options concerning theoretically unconstrained items.
Consider, for example, different visual mechanisms for detecting the rat’s
distance from the walls. Visual processing time is not constrained by the
mechanistic hypothesis under investigation, but different processing
speeds may affect system behavior in significant respects, including system
reactivity to visual stimuli and the achievement of an adequate neural
representation of the environment. Behavioral match judgments concern-
ing artificial and biological behaviors have to take into account (and
possibly factor out) the effects of these arbitrary implementation choices.
Similarly, different values for theoretically unconstrained parameters may
produce similar behaviors with respect to a limited range of set up con-
ditions, and significantly differing behaviors outside that range. Thus,
biorobotics gives rise to special cases of the arbitrary value adjustment
problem affecting mechanism-based modeling in biology at large, informa-
tively discussed by Hopkins and Leipold (1996).
Additional background assumptions that distinctively arise in bioro-
botics concern boundary conditions that are special to biorobot behavior.
Clearly, alterations of environmental or internal system conditions may
reveal limits in the scope of regularities governing the behavior of the
target biological system and its components. Similarly, the boundary con-
ditions ensuring “normal” behavior of a biorobot in experimental settings
may not coincide with the boundary conditions ensuring normal behavior
in the target biological system. For example, a plausible constraint on the
proper behavior of many robotic systems is that environmental temper-
ature does not exceed 40⬚C, insofar as the performance of microchips
rapidly deteriorates at higher temperatures. This condition concerns a
theoretically unconstrained feature of the implemented system; it can be
dropped if one replaces microchips with more robust chips or software
code (a similar issue is dealt with in Reeve et al. [2005]). This example

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
424 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

illustrates from another perspective the dependence of biorobot behaviors


on a wide variety of theoretically unconstrained details. Failure of the
robotic rat to perform properly in goal-seeking tasks may be attributed
to disturbing environmental conditions of this sort. Accordingly, bioro-
boticists have to monitor experimental environments for boundary con-
ditions concerning robot material components, and particularly items that
are theoretically unconstrained in the implementation, in order to make
sensible attributions of behavioral mismatches to incorrect predictions
flowing from mechanism descriptions and schemata.9

7. Summary. Biorobotic experiments involve running robotic implemen-


tations of mechanism schemata and descriptions, and evaluating their
behavior with respect to the behaviors produced by the target biological
system in similar experimental environments. Behavioral match evalua-
tions of robotic and biological behaviors may enable one to corroborate
or falsify mechanism schemata and descriptions, to identify previously
overlooked interactions with the environment, to suggest the existence of
entities and activities implementing some role function in the target bi-
ological system, and more generally to support the iterative specification
of mechanism descriptions from mechanism sketches.
These roles for robotic material models in experimental inquiries about
adaptive sensorimotor mechanisms have been chiefly explored here by
reference to biorobotic experiments concerning rat navigation and lobster
chemotaxis behaviors. In the former case, biorobotic experiments were
performed which suggest the existence of particular entities and activities
in the biological mechanism under investigation, in the light of the fact
that a biorobot successfully duplicating target biological system behaviors
was obtained by filling in gaps or underspecified components of mecha-
nism sketches and implementing the resulting mechanism description. In
the latter case, biorobotic experiments were performed which call and
heuristically guide the search for a new mechanistic model, insofar as
these experiments show that the mechanism description instantiated and
implemented in a biorobot fails to produce specific behaviors of the target
biological system.
Biorobotic experiments enable one to address a variety of additional
questions about adaptive sensorimotor mechanisms. If the set of variables
included in some mechanism description is reasonably small, one may use
biorobotic experiments to look for new correlations in a systematic fashion
(see, e.g., Chou and Hannaford 1997; Jaax and Hannaford 2002). And

9. In addition to this, bioroboticists have to adopt normal standards for experimental


inquiry, by checking whether these results are due to the improper working of mea-
surement tools (see related discussion in Franklin [2005]).

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 425

minor modifications of a mechanism description accounting for, say, rat


navigation behavior may enable one to replicate in a biorobotic experi-
ment adaptive navigation by another biological species, thus suggesting
that similar mechanisms produce the navigation behavior of different
biological species. Careful examination of these additional experimental
roles for biorobots is called for in a thorough methodological analysis of
biorobotics.
Mechanistic theorizing based on biorobotic experiments rests on a va-
riety of distinctive background assumptions, chiefly concerning imple-
mentation accuracy and multiple instantiation of mechanism schemata.
In particular, we have emphasized the role of implementation accuracy
assumptions in the inference process which leads one to claim corrobo-
ration (falsification) for some mechanistic hypothesis about some biolog-
ical system S, when starting from the experimental observation of match-
ing (mismatching) behaviors of biological system S and a biorobot R.
Implementation accuracy with respect to some mechanistic sketch or
description is compromised by implementation choices violating theoret-
ical constraints (either imposed by lack of adequate robotic hardware or
introduced to reduce implementation difficulties) and by adaptations of
theoretically unconstrained implementation details (achieved, say, by add-
ing modules or calibrating robot parameters). Additional methodological
problems that are special to biorobotics, and have to be properly addressed
to bring robotic behaviors to bear on biological mechanism schemata and
sketches include the selection of adequate matching criteria for comparing
behaviors produced by structurally different (biological and robotic) sys-
tems and the identification of boundary conditions that affect robot be-
havior but do not affect biological system behavior in the same experi-
mental setting.
On the whole, biorobotic inquiries pose distinctive methodological
problems that do not arise in computer simulation studies or in experi-
ments involving biological systems only. Notably, the methodological
problem of identifying undetected environmental features causing robot
malfunctioning does not arise in experiments involving computer simu-
lations, insofar as the environment is fully controlled there. And the need
for controlling biorobot implementation accuracy does not arise in ex-
periments involving biological systems only. An estimate of the theoretical
and experimental efforts that are needed to address these distinctive meth-
odological problems may prove crucial to deciding whether a biorobotic
experimental strategy is worth pursuing in the investigation of biological
sensorimotor mechanisms.10

10. The nontrivial character of this decision was already emphasized in Rosenblueth
and Wiener (1945, 318).

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
426 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

8. Concluding Remarks. Relatively recent developments of robotic tech-


nology have paved the way to the wide variety of biorobotic experiments
motivating the present methodological analysis. Recent advances in bi-
onics are now paving the way to novel experimental uses of robotic systems
and components in biological inquiry. In bionic systems, mutually inter-
connected robotic and biological parts are often interfaced to the central
or peripheral nervous system of some biological entity. This was achieved,
for example, in a bionic system including a lamprey brain stem (Reger et
al. 2000), engineered to perform sensorimotor transformations controlling
motion of a wheeled robotic platform toward light sources. Prosthetics
and rehabilitation arguably provide the primary motivation and range of
applications for bionic research (Dario et al. 2005). Bionic systems, how-
ever, naturally suggest a wide range of experiments related to purely the-
oretical inquiries into biological sensorimotor mechanisms. Indeed, bionic
systems implemented on the basis of abstract mechanism schemata which
admit both biological instantiation and “hybrid” bionic instantiation
might contribute to discovering and testing the functional roles of bio-
logical components in sensorimotor tasks. Consider a purely notional
bionic assembly, obtained by substituting nervous tissue extracted from
some biological system for the feedback-analyzer module of a light-seek-
ing robot. Experiments involving this bionic mechanism may enable one
to isolate the functional contribution of that biological component to the
light-seeking behavior of the biological system it was extracted from.
Additional experiments involving bionic assemblies may concern the stim-
ulation of learning or adaptation processes in biological nervous systems.
For example, the observation of behavioral plasticity in human subjects
whose peripheral nervous system is connected to a robotic arm prosthesis
might shed light on mechanisms for updating internal models of parts of
the body in the cerebellum. Thus, one is naturally led to extend the above
methodological analysis towards methodological issues distinctively aris-
ing in bionic experimentation about sensorimotor mechanisms.
Stimulation and inhibition experiments, conducted by means of bionic
or purely robotic systems, can be deployed to discover or test hypotheses
about mechanism decomposition hierarchies, in which lower level mech-
anism descriptions are obtained from higher level descriptions by opening
black boxes and identifying new regularities involving system variables.
This decomposition hierarchy, however, has to be carefully distinguished
from another kind of hierarchical organization of mechanism schemata
and descriptions, which is obtained by operations of abstraction or in-
stantiation. Mechanism schemata admit different instantiations, and each
one of these may come in different degrees, insofar as one can provide a
partial instantiation of an abstract mechanism schema (by indicating, e.g.,
appropriate fillers for only a few of its functional roles). Clearly, a mech-

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 427

anism schema M can be instantiated in mechanism description MS without


“opening” any functional box in the transition from M to MS (by mere
identification of fillers for the boxes that are already specified in M). And
M can be decomposed into sub-modules without providing any instan-
tiation for them. The interplay between functional decomposition and
instantiation hierarchies is crucial for the methodology of biorobotic ex-
perimentation at large. To illustrate, consider the above mentioned meth-
odological problems concerning implementation details. Working robotic
systems are obtained by appropriately instantiating abstract schemata and
implementing the resulting mechanism descriptions. To achieve this result,
one is forced to identify artificial entities and activities “all the way down,”
so as to generate proper system behavior. But instantiation and imple-
mentation of some mechanism schema M does not coincide with the
“opening of boxes” in M, insofar as some entities and activities in the
material robotic system are bound to be theoretically unconstrained im-
plementation details. An instantiated mechanism description does not
necessarily specify all relevant causal factors; and instantiation of mech-
anism schemata is not obtained by further modular decomposition. On
the whole, biorobotics appears to provide a suitable framework to pursue
an analysis of the interplay between abstraction and decomposition hi-
erarchies, of their respective role in the identification of different levels
of analysis for mechanism entities and activities, of the unification prob-
lems that a multilevel mechanistic analysis gives rise to.
The methodological reflections developed here are confined to mech-
anistic modeling and experimenting in biorobotics. These reflections con-
tribute to outlining a regulative methodological framework for contem-
porary biorobotic investigations of sensorimotor mechanisms, without
questioning the broad mechanistic assumption that adaptive sensorimotor
behavior is produced by the interaction of an array of structured and
hierarchically organized subcomponents. But an extension of this reflective
work, possibly relinquishing this mechanistic assumption, is naturally sug-
gested by the pervasive role of neural networks and dynamical systems
in biorobotics: neural network may come with a nonmodular organiza-
tion; dynamical systems are often used to describe relations between motor
variables, such as legs angles and insect motion (Beer 1997), without
identifying system components and their causal roles in the generation of
behavior. Biorobotics appears to be a particularly suited domain for an
in-depth exploration of both potentials and limitations of mechanistic
modeling and explanation.

REFERENCES
Amit, D. J. (1998), “Simulation in Neurobiology: Theory or Experiment?”, Trends in Neu-
rosciences 21: 231–237.

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
428 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

Barnes, C. A. (1988), “Spatial-Learning and Memory Processes—the Search for Their Neu-
robiological Mechanisms in the Rat”, Trends in Neurosciences 11: 163–169.
Bechtel, W., and A. Abrahamsen (2005), “Explanation: A Mechanist Alternative”, Studies
in History and Philosophy of Biological and Biomedical Sciences 36: 421–444.
Bechtel, W., and R. C. Richardson (1993), Discovering Complexity: Decomposition and Lo-
calization as Strategies in Scientific Research. Princeton, NJ: Princeton University Press.
Beer, R. D. (1997), “The Dynamics of Adaptive Behavior: A Research Program”, Robotics
and Autonomous Systems 20: 257–289.
Beer, R. D., et al. (1997), “Biologically Inspired Approaches to Robotics”, Communications
of the ACM 40: 30–38.
Berthoz, A. (1997), Le sens du movement. Paris: Odile Jacob.
Burgess, N., et al. (1994), “A Model of Hippocampal Function”, Neural Networks 7: 1065–
1081.
——— (1997), “Robotic and Neuronal Simulation of the Hippocampus and Rat Naviga-
tion”, Philosophical Transactions of the Royal Society; B: Biological Sciences 352: 1535–
1543.
——— (2000), “Predictions Derived from Modeling the Hippocampal Role in Navigation”,
Biological Cybernetics 83: 301–312.
Cacucci, F., et al. (2004), “Theta-Modulated Place-by-Direction Cells in the Hippocampal
Formation in the Rat”, Journal of Neuroscience 24: 8265–8277.
Chou, C. P., and B. Hannaford (1997), “Study of Human Forearm Posture Maintenance
with a Physiologically Based Robotic Arm and Spinal Level Neural Controller”, Bi-
ological Cybernetics 76: 285–298.
Cordeschi, R. (2002), The Discovery of the Artificial: Behavior, Mind and Machines Before
and Beyond Cybernetics. Dordrecht: Kluwer Academic Publishers.
Craver, C. F. (2001), “Role Functions, Mechanisms and Hierarchy”, Philosophy of Science
68: 31–55.
Craver, C. F., and L. Darden (2001), “Discovering Mechanisms in Neurobiology: The Case
of Spatial Memory”, in P. K. Machamer, R. Grush, and P. McLaughlin (eds.), Theory
and Method in Neuroscience. Pittsburgh: University of Pittsburgh Press, 112–137.
——— (2005), “Introduction”, Studies in History and Philosophy of Biological and Biomedical
Sciences 36: 233–244.
Cressant, A., et al. (1997), “Failure of Centrally Placed Objects to Control the Firing Fields
of Hippocampal Place Cells”, Journal of Neuroscience 17: 2531–2542.
Cummins, R. (1975), “Functional Analysis”, Journal of Philosophy 72: 741–765.
Dario, P., et al. (2002), “Design and Development of a Neurorobotic Human-Like ‘Guinea
Pig’”, Proceedings of Engineering in Medicine and Biology, 2002: 24th Annual Conference
and the Annual Fall Meeting of the Biomedical Engineering Society 3: 2345–2346.
——— (2005), “Robotics as a Future and Emerging Technology”, IEEE Robotics & Au-
tomation Magazine 12: 29–45.
Eskiizmirliler, S., et al. (2002), “A Model of the Cerebellar Pathways Applied to the Control
of a Single-Joint Robot Arm Actuated by McKibben Artificial Muscles”, Biological
Cybernetics 86: 379–394.
Fox, S. E., and J. B. Ranck (1975), “Localization and Anatomical Identification of Theta
and Complex Spike Cells in Dorsal Hippocampus in Rats”, Experimental Neurology
49: 299–313.
Franklin, A. (2005), No Easy Answers: Science and the Pursuit of Knowledge. Pittsburgh:
University of Pittsburgh Press.
Glennan, S. (1996), “Mechanisms and the Nature of Causation”, Erkenntnis 44: 49–71.
——— (2002), “Rethinking Mechanistic Explanation”, Philosophy of Science 69 (Proceed-
ings): S342–S353.
——— (2005), “Modeling Mechanisms”, Studies in History and Philosophy of Biological
and Biomedical Sciences 36: 443–464.
Grasso, F. W., et al. (2000), “Biomimetic Robot Lobster Performs Chemo-Orientation in
Turbulence Using a Pair of Spatially Separated Sensors: Progress and Challenges”,
Robotics and Autonomous Systems 30: 115–131.

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
BIOROBOTIC EXPERIMENTS 429

Hafting, T., et al. (2005), “Microstructure of a Spatial Map in the Entorhinal Cortex”,
Nature 436: 801–806.
Hopkins, J. C., and R. J. Leipold (1996), “On the Dangers of Adjusting the Parameters
Values of Mechanism-Based Mathematical Models”, Journal of Theoretical Biology 183:
417–427.
Horiuchi, T., and C. Koch (1996), “Analog VLSI Circuits for Visual Motion-Based Ad-
aptation of Post-Saccadic Drift”, Proceedings of the 5th International Conference on
Microelectronics for Neural Networks and Fuzzy Systems, 60–66.
Ishida, H., et al. (2001), “Plume-Tracking Robots: A New Application of Chemical Sensors”,
Biological Bulletin 200: 222–226.
Jaax, K. N., and B. Hannaford (2002), “A Biorobotic Structural Model of the Mammalian
Muscle Spindle Primary Afferent Response”, Annals of Biomedical Engineering 30: 84–
96.
Jarrad, L. E. (1993), “On the Role of the Hippocampus in Learning and Memory in the
Rat”, Behavioral and Neural Biology 60: 9–26.
Jeffery, K. J., et al. (1997), “Directional Control of Hippocampal Place Fields”, Experimental
Brain Research 117: 131–142.
Lambrinos, D., et al. (2000), “A Mobile Robot Employing Insect Strategies for Navigation”,
Robotics and Autonomous Systems 30: 39–64.
Lloyd, E. A. (1987), “Confirmation of Ecological and Evolutionary Models”, Biology and
Philosophy 2: 277–293.
Loeb, J. (1918), Forced Movements, Tropism, and Animal Conduct. Philadelphia: Lippincott.
Machamer, P., L. Darden, and C. F. Craver (2000), “Thinking about Mechanisms”, Phi-
losophy of Science 67: 1–25.
Marr, D. (1982), Vision: A Computational Investigation into the Human Representation and
Processing of Visual Information. San Francisco: Freeman.
Muller, R. U., and J. L. Kubie (1987), “The Effects of Changes in the Environment on the
Spatial Firing of Hippocampal Complex-Spike Cells”, Journal of Neuroscience 7: 1951–
1968.
O’Keefe, J., and N. Burgess (1996), “Geometric Determinants of the Place Fields of Hip-
pocampal Neurons”, Nature 381: 425–428.
O’Keefe, J., and J. Dostrovsky (1971), “The Hippocampus as a Spatial Map: Preliminary
Evidence from Unit Activity in the Freely Moving Rat”, Brain Research 34: 171–175.
O’Keefe, J., and L. Nadel (1978), The Hippocampus as a Cognitive Map. Oxford: Clarendon
Press.
O’Keefe, J., and M. Recce (1993), “Phase Relationship between Hippocampal Place Units
and the EEG Theta Rhythm”, Hippocampus 3: 317–330.
Prescott, T. J., et al. (2002), “The Robot Basal Ganglia: Action Selection by an Embedded
Model of the Basal Ganglia”, in L. F. B. Nicholson and R. Faulls (eds.), Basal Ganglia
VII. New York: Plenum Press, 349–356.
Quinn, R. D., and R. E. Ritzmann (2001), “Construction of a Hexapod Robot with Cock-
roach Kinematics Benefits Both Robotics and Biology”, in B. Webb and T. R. Consi
(eds.), Biorobotics: Methods and Applications. Cambridge, MA: AAAI Press, 87–105.
Reeve, R., et al. (2005), “New Technologies for Testing a Model of Cricket Phonotaxis on
an Outdoor Robot”, Robotics and Autonomous Systems 51: 41–54.
Reger, B. D., et al. (2000), “Connecting Brains to Robots: An Artificial Body for Studying
the Computational Properties of Neural Tissues”, Artificial Life 6: 307–324.
Rosenblueth, A., and N. Wiener (1945), “The Role of Models in Science”, Philosophy of
Science 12: 316–321.
Rosenblueth, A., N. Wiener, and J. Bigelow (1943), “Behavior, Purpose and Teleology”,
Philosophy of Science 10: 18–24.
Schaal, S., and D. Sternad (2001), “Origins and Violations of the 2/3 Power Law in Rhythmic
Three-Dimensional Arm Movements”, Experimental Brain Research 136: 60–72.
Tabery, J. (2004), “Synthesizing Activities and Interactions in the Concept of a Mechanism”,
Philosophy of Science 71: 1–15.
Tamburrini, G., and E. Datteri (2005), “Machine Experiments and Theoretical Modeling:
from Cybernetic Methodology to Neuro-Robotics”, Minds and Machines 15: 335–358.

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms
430 EDOARDO DATTERI AND GUGLIELMO TAMBURRINI

Taylor, R. (1950), “Comments on a Mechanistic Conception of Purposefulness”, Philosophy


of Science 17: 310–317.
Tolman, E. C. (1948), “Cognitive Maps in Rats and Men”, Psychological Review 55: 189–
208.
Von Eckardt, B., and J. S. Poland (2004), “Mechanism and Explanation in Cognitive Neu-
roscience”, Philosophy of Science 71: 972–984.
Webb, B. (2001), “Can Robots Make Good Models of Biological Behaviour?”, Behavioral
and Brain Sciences 24: 1033–1050.
——— (2002), “Robots in Invertebrate Neuroscience”, Nature 417: 359–363.
——— (2006), “Validating Biorobotic Models”, Journal of Neural Engineering 3: R25–R35.
Webb, B., and T. R. Consi (2001), Biorobotics: Methods and Applications. Cambridge, MA:
MIT Press.
Wilson, M. A., and B. L McNaughton (1993), “Dynamics of the Hippocampal Ensemble
Code for Space”, Science 261: 1055–1058.
Woodward, J. (2002), “What Is a Mechanism? A Counterfactual Account”, Philosophy of
Science 69 (Proceedings): S366–S377.

This content downloaded from


138.36.64.82 on Wed, 23 Jan 2019 20:14:56 UTC
All use subject to https://about.jstor.org/terms

You might also like