Representations of Space and Time in The Maximization of Information Flow in The Perception-Action Loop

Download as pdf or txt
Download as pdf or txt
You are on page 1of 47

LETTER Communicated by Nihat Ay

Representations of Space and Time in the Maximization of


Information Flow in the Perception-Action Loop

Alexander S. Klyubin
[email protected]
Adaptive Systems Research Group, School of Computer Science,
University of Hertfordshire, Hatfield Herts AL10 9AB, U.K.

Daniel Polani
[email protected]
Adaptive Systems and Algorithms Research Groups, School of Computer Science,
University of Hertfordshire, Hatfield Herts AL10 9AB, U.K.

Chrystopher L. Nehaniv
[email protected]
Adaptive Systems and Algorithms Research Groups, School of Computer Science,
University of Hertfordshire, Hatfield Herts AL10 9AB, U.K.

Sensor evolution in nature aims at improving the acquisition of infor-


mation from the environment and is intimately related with selection
pressure toward adaptivity and robustness. Our work in the area indi-
cates that information theory can be applied to the perception-action
loop.
This letter studies the perception-action loop of agents, which is mod-
eled as a causal Bayesian network. Finite state automata are evolved as
agent controllers in a simple virtual world to maximize information flow
through the perception-action loop. The information flow maximization
organizes the agent’s behavior as well as its information processing. To
gain more insight into the results, the evolved implicit representations of
space and time are analyzed in an information-theoretic manner, which
paves the way toward a principled and general understanding of the
mechanisms guiding the evolution of sensors in nature and provides
insights into the design of mechanisms for artificial sensor evolution.

1 Introduction

1.1 Information in the Perception-Action Loop. In nature, sensors and


actuators of living agents are constantly engaged in complicated interaction.
Despite the sensors and the actuators being specific to agents’ niches and
needs, a general and at the same time a minimal model of sensorimotor
interaction is desired.

Neural Computation 19, 2387–2432 (2007) 


C 2007 Massachusetts Institute of Technology
2388 A. Klyubin, D. Polani, and C. Nehaniv

Although a large body of research identifies information as useful for


understanding sensors, less work has been done on actuators. The pressure
to capture information relevant to an agent’s survival may push sensors
to the limits. For example, vision sensors can operate at the limits set by
physics (Bouman, 1959; Bialek, 1987). Capturing and processing sensory
information have their metabolic costs (Laughlin, de Ruyter van Stevenick,
& Anderson, 1998). Neural pathways have limited information transmis-
sion rates. Optimization of information-related quantities in artificial neural
networks results in phenomena similar to ones found in biological sen-
sory pathways (Barlow, 1959; Linsker, 1988; Nadal & Parga, 1995; Bell &
Sejnowski, 1995) and information-theoretic tools provide versatile ways
to analyze biological and artificial neural systems (Avraham, Chechik, &
Ruppin, 2003; Schneidman, Bialek, & Berry, 2003).
Actions have also been identified as important from an informational
perspective; however, to the best of our knowledge, less research has
treated actions information-theoretically on par with sensors. Actions may
facilitate active perception to uncover hidden information or simplify in-
formation processing in sensory pathways (Kirsh & Maglio, 1994; Nolfi
& Marocco, 2002). Actions are also crucial for offloading and later reac-
quisition of information (Hutchins, 1995; Klyubin, Polani, & Nehaniv,
2004c).
In this letter, the use of causal Bayesian networks is introduced to model
the perception-action loop. This formalism allows the treatment of percep-
tion action in a consistent information-theoretic manner: from the capture
of environmental information through the sensors, to the “imprinting” of
information onto the environment through the actuators. The causal struc-
ture of the networks allows the formulation of the concept of information
flow (Ay & Wennekers, 2003; Klyubin, Polani, & Nehaniv, 2004a; Ay &
Krakauer, 2007). Information flows are a signature of the causal structure,
and they allow one to track the flow of pieces of information through the
system analogous to how the flow of radioisotope tracers is tracked through
a living organism.
Quantitative information-theoretic treatment of sensors and actions as
parts of the perception-action loop was implicitly done in the early work
of Ashby (1956) and recently revived by Touchette and Lloyd (2000). This
letter proposes an information-theoretic method with a minimal set of as-
sumptions for studying and constructing perception-action loops, which
may span an arbitrary number of time steps as opposed to just a single
time step as in the models of Ashby, and Touchette and Lloyd. The use of
information theory reduces the bias of the processing architecture to a mini-
mum, creating a “coordinate-free”1 view of what happens with information
processing.

1 We are indebted to Peter Dauscher for suggesting this analogy.


Information Flow in the Perception-Action Loop 2389

1.2 Contributions of the Letter. The main contributions of this letter


are:
1. Formulating an information-theoretic picture of the perception-action
loop with a minimal set of assumptions. The interactions of an agent’s
sensors, actuators, memory, and the environment are modeled using
causal Bayesian networks.
2. Maximizing information flow through the full perception-action loop
of an agent, replacing structural assumptions such as those of infomax
of Linsker (1988; e.g., receptive fields, layers) or the compositional-
ity of temporal infomax of Ay and Wennekers (Ay & Wennekers,
2003; Wennekers & Ay, 2005; i.e., components of a finely grained
computational architecture, which are identifiable through time) by
agent-environment interaction over many time steps and minimal
assumptions about the agent’s information processing.
3. Using information-theoretic tools to analyze representations2 of space
and time resulting from the perception-action loop infomax. Since
some of the representations are hard to interpret due to the lack of
architectural assumptions, the judicious use of information-theoretic
tools is essential.

1.3 Outline of the Letter. This letter is structured as follows. First,


related work is discussed in relation to this letter in section 2. In section
3, a formalism for treating the perception-action loop using information-
theoretic methods is laid out.
In section 4, the formalism is used to maximize an information flow
through a simulated agent and its environment through time. Evolutionary
search is employed in the experiment to find agent controllers that capture
the maximum amount of information about the state of the world.
The evolved controllers are analyzed in the second part of the letter
by looking at their mechanisms and their behavior through information-
theoretic tools. The following questions are addressed. What information
about the initial position do the evolved automata capture? How do the
controllers capture the information? How is the information represented in
the state of the controller?
In section 5 the representation of the agent’s initial position in the state
of its controller at the end of the run is analyzed, extending the results of
Klyubin et al. (2004a) by studying how the representations change with
small variations in the agent’s “embodiment” (what the sensors capture
from the environment and what the actuators do in the environment).

2Representation is a loaded term (cf. Millikan, 2002; Usher, 2001). In this letter, the
“representation” of one variable by another means the probabilistic mapping or corre-
spondence between the two variables.
2390 A. Klyubin, D. Polani, and C. Nehaniv

In section 6 the representation of time inside the agent and by the agent-
environment system as a whole is studied. A method to quantify the time
dependence of representations is proposed in section 7. Since some of the
representations appear to be factorizable (decomposable into independent
components), in section 8 the representations of space (initial position)
and time are factorized. Finally, overall discussion and conclusions are
presented in section 9.
In appendix A, the information-theoretic quantities used in the letter
are briefly defined. Appendix B remarks on modeling an agent’s controller
using causal Bayesian networks. Appendices C and D explain how infor-
mation flows are measured and how the controllers are evolved in this
letter. Appendix E proves the assertion for the grid-world model used in
this letter that when an agent follows the gradient and reaches the source,
most of the information about the agent’s initial position relative to the
source passes through its sensors. Appendix F continues the discussion
of time dependence of representation by considering two examples. Ap-
pendix G contains the details of the factorization method used in section 8
and discusses its relation to synergy and the parallel bottleneck case of the
multivariate information bottleneck method.

2 Background

2.1
. Related Work

2.1.1 Perception-Action Loop and Information Theory. Gibson (1979) pro-


posed that animals and humans do not normally view the world in terms
of a geometrical space, independent arrow of time, and Newtonian me-
chanics and that the natural description is rather in terms of what an agent
can perceive and do. The interplay between sensors and actuators, the
perception-action loop, is also central to the notion of homeostasis, intro-
duced by Cannon (1939) and treated in more detail by Ashby (1952, 1956).
The main assumption is that agents act so as to keep certain internal vari-
ables (essential variables) within desirable ranges. Considering that the
essential variables are the result of perception of external or internal states,
homeostasis creates a purely internal goal, where action is for the purpose
of controlling perception (Powers, 1974). Homeokinesis (Der, Steinmetz, &
Pasemann, 1999) is a dynamic version of homeostasis where the agent acts
so as to minimize the prediction error of its past and future sensoric input.
If the sensors can be seen as taking information in from the environment,
the actuators can then also be thought of as having the capacity to modify the
environment in terms of information. Interestingly, Gibson (1979) argued
against applying information theory to the agent’s perceptions, his main
argument being that the environment does not intentionally “speak” to the
observer and that “the assumption that information can be transmitted and
Information Flow in the Perception-Action Loop 2391

the assumption that it can be stored” are not appropriate for the theory of
perception (p. 242).
However, work in perceptual neural networks (see section 2.1.2), recent
work in control (Touchette & Lloyd, 2004), and implicitly also Ashby’s
work (1956) show that information theory can be productively applied to
perception-action without the need to assume any intentionality. Touchette
and Lloyd (2000, 2004) provided bounds on the efficiency of any sensor for
one-step control (entropy reduction). Their result for closed-loop control
can be interpreted as a generalized statement of Ashby’s law of requisite
variety, which states that only variety (entropy) in actions can reduce the
variety (entropy) in the environment.

2.1.2 Neural Networks and Information Theory. Attneave and Barlow in


the mid-1950s recognized the importance for the brain not just to capture
information but also to represent it in a convenient way. Attneave (1954) ob-
served that sensoric input of living organisms is very much redundant. He
hypothesized that sensory pathways take advantage of the redundancy and
also abstract away unnecessary or overwhelming details to represent the
remaining sensoric information in an economical way. Barlow (1959, 1963)
proposed the principle of redundancy reduction according to which sensory
information in the brain is incrementally recoded to reduce the redundancy
of its representation. Recently (Barlow, 2001) relaxed the claim by noting
that (1) further up the processing chain, redundancy of representations can
actually increase rather than decrease, and (2) highly compressed represen-
tations may be harder to use than more redundant ones. The emphasis was
thus shifted from redundancy reduction to redundancy exploitation.
In a layered or hierarchical system it is unclear what information each of
the layers ought to preserve without designing the knowledge of ultimate
goals of the system into each layer. This is addressed by Linsker’s princi-
ple of maximum information preservation, or infomax (Linsker, 1988). The
principle requires each layer to maximize information transmission from
its inputs to its outputs without bias as to what information to preserve
and without the need for a layer to know anything about other layers. Info-
max usually assumes spatial structure. For example, applied to a simulated
retina-like layered feedforward neural network with localized receptive
fields, the principle made the network self-organize and develop various
feature detectors found also in the retinas of living beings.
In his later work Barlow extended what he meant by redundancy to in-
clude not only the unused capacity of a channel but also what he called “hid-
den redundancy”—the statistical interdependencies between components
of output (Barlow, 1963, 2001). He proposed minimal entropy coding, which
reduces “hidden” redundancy by decorrelating the components and leads
to factorial codes where probabilities of individual components are inde-
pendent of each other (Barlow, 1989). Nadal and Parga (1995) demonstrate
that this particular type of Barlow’s redundancy reduction and Linsker’s
2392 A. Klyubin, D. Polani, and C. Nehaniv

infomax are related and sometimes equivalent. A similar result is obtained


in the context of source separation by Bell and Sejnowski (1995). A related
method for arriving at factorial codes and feature detectors is the prin-
ciple of predictability minimization by Schmidhuber (1992; Schmidhuber,
Eldracher, & Foltin, 1996). The parallel information bottleneck principle
proposed by Friedman, Mosenzon, Slonim, and Tishby (2001) as a vari-
ety of multivariate information bottleneck can also create factorial codes,
the advantage over other methods above being the capability for lossy
relevance-based compression.
Ay (2002) noted that although infomax may explain the functioning
of early sensory stages, at later stages information from various path-
ways is integrated. He proposed the hypothesis of invariant maximization
of interaction according to which “learning . . . is driven by mechanisms
that maximize the stochastic interaction” between components of systems.
The maximization of global stochastic interaction results in local infomax
consistent with Linsker’s idea. By incorporating the dynamical aspect Ay
and Wennekers (Ay & Wennekers, 2003; Wennekers & Ay, 2005) proposed
the principle of temporal infomax according to which the spatiotemporal
stochastic interaction between various pathways or components is max-
imized. The approach is compositional as it assumes that an informa-
tion processing system consists of components (e.g., neurons) that retain
their identity through time. The complexity of its information processing
is quantified information-theoretically by comparing components at two
adjacent time steps. Temporal infomax consists of the maximization of this
complexity.

2.2 Relation of Previous Work to This Letter. The prior work already
noted is related to this work as follows:
1. The minimalistic model of an agent employed in this letter (see section
3.1) is defined in terms of what the agent’s controller can access:
sensors, actuators, and memory, conforming to Gibson’s philosophy.
2. In this letter, the treatment of the perception-action loop by Ashby
and by Touchette and Lloyd is generalized by modeling an arbitrary
number of time steps.
3. The conceptual difference between this work and Linsker’s infomax
is that internal spatial structure in the processing architecture (i.e.,
receptive fields and layers of neurons) is replaced by natural con-
straints coming only from the agent-environment interaction (“em-
bodiment”) extended over many time steps.
4. Although information flow is maximized through time in this letter,
the perception-action loop infomax is conceptually different from the
temporal infomax of Ay and Wennekers. First of all, compositionality,
the existence of separate finely grained components, is central to their
Information Flow in the Perception-Action Loop 2393

quantification method, whereas the split into sensors, actuators, and


memory in this letter is not crucial for determining information flow
quantities. Secondly, the temporal aspect of temporal infomax spans
only two adjacent time steps, whereas this letter deals with informa-
tion flows spanning an arbitrary numbers of time steps. The formal-
ism presented here can be used to quantify and maximize information
flows occurring between different parts of a system unconstrained by
any temporal restrictions (e.g., information flows occurring across
components but not through time), whereas the temporal aspect is
central to the temporal infomax.
5. The factorization of the representations resulting from perception-
action infomax (see section 8) is intimately related to but different
from the creation of factorial codes. Here not only the output but
also the mapping between the input and the output is required to be
factorized.

3 Bayesian Model of the Perception-Action Loop

3.1 The Model. Central to the methodology is to model the perception-


action loop of an agent as a causal Bayesian network. This can be a general
and minimal model of an agent with a controller, possibly with memory.
The following assumptions are made in the model:
r The agent is part of a larger agent-environment system.
r The system has discrete states.
r The system is discrete in time.
r Consecutive states of the system form a Markov chain. The momentary
state of the system makes the past of the system statistically indepen-
dent from its future.
r The agent’s controller selects actions and modifies its own memory
having access only to the momentary state of the sensors and own
memory.
There are many ways to partition such a system into subsystems. Here
the partitioning is done from the perspective of the agent’s controller. The
controller has direct access only to the agent’s sensor,3 the actuator, and its
own memory. Everything else in the system that is not encompassed by
these three components is the rest of the system.

3 Although the agent can have more than one sensor and actuator, from now on
they will be referred to in the singular. Multiple sensors can of course be treated as one
composite sensor. Similar reasoning applies to actuators.
2394 A. Klyubin, D. Polani, and C. Nehaniv

All the constituents of the perception-action loop are modeled as random


variables:
r S—the state of the sensor
r A—the action performed by the actuator
r M—the state of the controller’s memory
r R—the state of the rest of the system
R formally accounts for the effects of actuation, the agent’s environment,
and morphology on the sensors.
Sensors and actuators enable information flows between the agent and
its environment. Time is introduced to model the temporal aspect of the
flows. The states of the sensor, the actuator, the controller’s memory, and
the rest of the system at discrete time t are denoted by random variables St ,
At , Mt , and Rt , respectively.
Modeling relations between the variables at different time steps unrolls
the perception-action loop in time. The approach is not new (Touchette &
Lloyd, 2000). However, as opposed to Touchette and Lloyd (2000), where
only one time step is modeled, an arbitrary number of time steps is modeled
here.
The relationships between the variables are modeled as a causal4
Bayesian network (Pearl, 2001), which is a directed acyclic graph where
any node, given its parents, is conditionally independent from any other
node that is not its parent or successor (any node directly or indirectly
reachable from the node).
The pattern of relations between variables at consecutive time steps is
shown in Figure 1. Assume that the pattern of relations is time invariant
and thus holds for any t. Thus, the graph in Figure 1 is just a section of the
network.
The agent’s controller is modeled with minimum assumptions about its
architecture: the controller is discrete state, discrete time, and obeys the laws
of classical probabilities. In the most general case, the controller performs
a probabilistic mapping (St , Mt ) → (At , Mt+1 ).5 See appendix B for a more
detailed discussion of the controller’s model as a causal Bayesian network.

4 A Bayesian network captures statistical relationships between observed variables.


More than one Bayesian network may fit observed data. In a causal Bayesian network, the
relationships between variables are not just statistical but also causal, corresponding to
the underlying mechanisms that generate the data. This allows one to calculate effects of
interventions—changes in some of the mechanisms (consult Pearl, 2001, for an in-depth
treatment). Examples of interventions include fixing a variable to a particular value, or
“injecting” information into the system (see below). The latter is used in this letter. In
general, interventions can be defined only for causal Bayesian networks.
5 Throughout the letter the probabilistic mapping from a random variable X to a

random variable Y is denoted using the shorthand notation X → Y, which stands for the
Markov kernel X × Y → [0, 1], (x, y) → Pr (Y = y|X = x).
Information Flow in the Perception-Action Loop 2395

Figure 1: Perception-action loop as a Bayesian network. S—state of the sensor;


A—action performed by the actuator; M—state of the memory of the controller;
R—state of the rest of the agent-environment system. The diagram can be read
as follows: action At is picked given sensor state St and memory state Mt , sensor
state St is obtained from the state of the rest of the agent-environment system
Rt , and Rt+1 is obtained from Rt and At .

The causal Bayesian network model of the perception-action loop is use-


ful for several reasons. First, the relations between the components of the
perception-action loop are visualized. Second, various paths of the informa-
tion flow (see section 3.2) and constraints on its bandwidth are visualized.
For example, it is clear that information flows between the memory M and
the rest of the system R are mediated by the sensors and actuators, and,
hence, the bandwidth of the sensors and actuators sets a limit on the rate of
information flows between the agent’s memory and the rest of the system.
Third, the model, being a causal Bayesian network, makes it possible to
analytically calculate various probability distributions as well as the results
of interventions (see Pearl, 2001).

3.2 Information Flow. In the material world, the concept of matter flow
is natural and widespread owing to the conservation laws and additivity
of matter. For example, in nuclear medicine, radioisotope tracers are in-
jected into a patient, and their flow through the organism is tracked. Once
one tries to formulate an analogous concept of information flow, one en-
counters a number of differences, one being the nonconservative nature of
classical information unlike matter. Classical information can be duplicated
or destroyed. Furthermore, information can be “encrypted,” making it dis-
appear unless one knows another “piece” of information that serves as the
“decryption” key.6
As an illustration, consider an organization where confidential infor-
mation is leaked to the outside and there are several suspects who have
access to the information. To identify which one is leaking the information,
one would provide each suspect with a different and deliberately made-up
piece of information and see what information gets out. The presence of the

6 Here encryption is used in the broad sense of the word. The one-time pad encryption
scheme (Schneier, 1995) is one example. Another example is the use of an address as a
key to access information on a storage medium.
2396 A. Klyubin, D. Polani, and C. Nehaniv

particular piece of injected information outside the organization will have


to be solely due to the information flowing from the confidential source via
one of the suspects.
Analogously, in this letter, information flow is defined for an arbitrary
system modeled as a causal Bayesian network. Given a causal Bayesian
network information flow from a random variable X to a random variable
Y is defined as the amount of information about X causally transmitted from
X to Y. Analogous to the information leakage example, if X is an input node
of the network, then any information about X in a node Y is necessarily
due to the causal effect (Pearl, 2001) of X on Y. In that case, the amount of
information flow from X to Y is I (X; Y), the information-theoretic mutual
information between X and Y. Information flow is conceptually different
from mutual information. Mutual information is a correlational and thus
symmetric quantity: I (X; Y) = I (Y; X) for any X and Y. Information flow
is a causal quantity and is therefore directional and asymmetric.7
It may be possible to quantify the information flow in the general case,
where X is an arbitrary node of the network. One needs to distinguish be-
tween information shared between X and Y due to the causal effect of X
on Y and the information arriving from common predecessors. One could
“block” unwanted paths of information flow to Y, leaving only those pass-
ing via X (see Klyubin, Polani, & Nehaniv, 2004b, section 8.2, for an example
and Ay & Krakauer, 2007, for a more general treatment). A causal Bayesian
network model helps identify the suitable sets of blocking variables.

4 Information Capture Experiment

4.1 Motivation. Linsker’s infomax principle for neural networks is con-


cerned with the maximization of the information flow from the inputs
of a perceptual layer of to its outputs (Linsker, 1988). As a consequence
of preserving information about inputs, layers of center-surround cells,
orientation-sensitive cells, and orientation columns appeared. An extended
and generalized version of the infomax experiment can be performed using
the causal Bayesian model of the perception-action loop. Instead of maxi-
mizing the information flow across just one layer of neurons, one can max-
imize the information flow through the perception-action loop of an agent
through an arbitrary number of time steps. Moreover, since the model is
more general, one need not assume linearity or a particular information
processing substrate such as locally connected layers of neurons.
Chemotaxis, phototaxis, or gradient following in general is a good sce-
nario because information about the position of the gradient’s point source
must flow through the agent’s perception-action loop. In the following

7 In fact, information flow is nonzero only in the direction of the causal future (descen-

dants) of X.
Information Flow in the Perception-Action Loop 2397

experiment, an agent controller is evolved to capture the agent’s initial


position relative to the source of the gradient. The goal may also be inter-
preted as the agent reconstructing the position of the gradient source at the
beginning of the experiment.

4.2 Outline of the Experiment. An agent moves in an infinite two-


dimensional grid world.8 The position of the agent corresponds to R, the
rest of the system, in the Bayesian model of the perception-action loop.
The agent’s controller has no direct access to the agent’s position. The
controller has access to a gradient sensor that provides only a small amount
of information about the position.
The agent’s controller is evolved to maximize the information flow from
the initial position R0 into M15 , the state of the controller’s memory at time
step 15, under the condition that the controller is always started in the same
initial state. Similar to the infomax experiment, the features of the initial
position captured by the memory of the evolved controllers are analyzed. It
turns out that most of the features have strong spatial continuity properties
due to the way the agent perceives the world and acts in it. Therefore,
small changes in the sensorimotor apparatus of the agent are made, and the
impact on the captured features is studied.

4.3 Test Bed. The environment is a two-dimensional grid of infinite


size. A source is located at the origin of the grid. The source emits  a signal,
the strength P of which in any cell of the grid is P(r ) = r −2 P(0) = 2 ,
where r is the distance to the source. The exact relation is not important for
the experiments. It is only important that the decrease is strictly monotonic
with distance.
An agent is situated in a single cell at a time. The agent has a gradient
sensor. The gradient points to the cell with the highest signal strength
among the four adjacent cells (north, east, south, west). If there are several
cells with highest signal strength, the gradient randomly points to one of
these with equal probability (see Figure 2). The agent also has an actuator:
at each time step the agent performs one of the four available actions: move
north, east, south, or west. The agent has a controller, which takes sensoric
input and then moves the agent.
The global state of the system consists of just the agent’s position on the
grid and the state of the agent’s controller. Following the perception-action
loop model from section 3, denote the sensoric input (the gradient) with the
random variable S taking values from the set S = {s N , s E , s S , sW }; the action
with the random variable A taking values from the set A = {a N , a E , a S , a W };

8The grid is large enough so that the agent never experiences the boundary conditions.
Without loss of generality, the grid can be considered infinite; however, only a finite
number of cells is ever reached.
2398 A. Klyubin, D. Polani, and C. Nehaniv

Figure 2: Sensoric input S depending on the position R. In cells with only one
arrow the sensor input is constant. In cells with multiple arrows it is randomly
chosen from the one of the arrows shown in the cells. The source is located at
the center of the grid.

the internal state or the memory of the agent’s controller with the random
variable M taking values from the set M = {1, 2, . . . , N}, where N is the
number of states; and the two-dimensional position of the agent with the
random variable R taking values from the set Z2 . S, A, M, and R completely
describe the state of the whole system at any instant.
At and Mt+1 depend directly only on (St , Mt ) for any t (see Figure 1).
The agent’s controller, regardless of its implementation, is performing a
probabilistic mapping (St , Mt ) → (At , Mt+1 ). The mapping is assumed to be
time invariant.
The probabilistic mapping can be implemented by a nondeterministic
Mealy finite state automaton operating with input set S, output set A,
and state set M. Preliminary experiments showed that the evolved map-
pings tend to be deterministic; hence, the experiments are constrained to
deterministic automata, allowing only deterministic mappings. However,
without any loss of generality, the approach can be used with stochastic
mappings. Determinism of the controller is an experimental choice, not a
limitation of the model.

4.4 Initial Position Capture. For this experimental setup it can be


shown theoretically that, when an agent follows the gradient and ends
up near the source, most of the information about the agent’s initial posi-
tion passes through the gradient sensor (see appendix E for a proof). Can
this flow be directed into the agent’s memory to capture information about
Information Flow in the Perception-Action Loop 2399

Figure 3: Initial position capture as optimization of the communication channel


between R0 and Mt . Only the mappings represented by wavy lines can be
modified.

the initial position of the agent, using a controller with limited memory and
given limited time?
In finding a controller that maximizes the information flow9 from R0
into Mt , where t  is the duration of the run (see Figure 3), the controller
can be seen as constructing a temporally extended communication channel
between these two temporally separated parts of the system, passing not
only through the sensor but also via the memory, the actuator, and the
environment.
Two types of controllers are evolved:10 (1) unconstrained, which are not
constrained in terms of the implemented mapping (see Figure 3a) and (2)
gradient following (see Figure 3b), which are constrained to follow the gra-
dient. Clearly, gradient-following controllers are a subset of unconstrained
controllers where the information flow from M to A is blocked.
To evaluate a controller, the agent’s initial position R0 is randomly and
uniformly distributed over a square with side d centered at the source. At
the beginning of the run, the controller is always in state 1. The agent moves
for t  time steps, after which its fitness I (R0 ; Mt ) is evaluated.
I (R0 ; Mt ) is the amount of information about the agent’s initial position
captured by the state of controller at time step t  . It quantifies how well one

9 See appendix C for a brief description of how the flow is measured.


10 The details of the evolutionary algorithm employed can be found in appendix D.
2400 A. Klyubin, D. Polani, and C. Nehaniv

Figure 4: Information flow for best controllers. The amount of information


flow from R0 into Mt (ordinate) is plotted against the size of the controller’s
memory (|M|, abscissa). Both quantities are measured in bits. Dotted line:
gradient-following controllers; solid line: unconstrained; dashed line: upper
bound dictated by the IB method for gradient-following controllers. Each point
is the maximum of five evolutionary runs. d = 5, t  = 5, 1000 generations.

can predict the state of the controller at time step t  from the initial position
of the agent; or, alternatively, how well one can deduce the initial position
of the agent knowing only the state of its controller at time step t  .
Evolutionary search is not guaranteed to find optimal solutions. How-
ever, an upper bound for achievable performance is the amount of in-
formation about R0 in principle extractable from the cumulative sensoric
input Stc −1 = (S0 , S1 , . . . , St −1 ) into Mt . Since R0 , Stc −1 , and Mt form a data
processing chain, it follows from the data processing inequality that no
controller can extract more information into Mt about R0 than it obtains via
Stc −1 . However, since the cumulative sensoric input contains in the (tempo-
ral) limit all the information about the initial position, the main constraint
derives from the agent’s controller having access only to the momentary
sensoric input, one at a time, and having limited memory.
The bound can be estimated for gradient-following controllers. Since
all gradient-following controllers by definition employ exactly the same
actuation strategy, that is, follow the gradient, the probability distribution
of the random variable Stc −1 is exactly the same for all the controllers.
To calculate the bound, the information bottleneck (IB) principle (Tishby,
Pereira, & Bielek, 1999) can be used to find a mapping Stc −1 → Mt that
maximizes the mutual information I (R0 ; Mt ). The latter is an upper bound
for the amount of information about R0 any gradient-following controller
can capture in Mt .
In Figure 4, the performance of best-evolved controllers is compared
to the size of their memory in bits. The controllers manage to extract in-
formation that is temporally “smeared” in sensoric input. Unconstrained
Information Flow in the Perception-Action Loop 2401

Figure 5: Best evolved four-state unconstrained controller as a finite state au-


tomaton. Circles are states. Arrows are transitions between states. Every trans-
action is annotated with the sensoric input from {n, e, s, w}, which triggers the
transition, and this is followed by the performed action from {n, e, s, w}.

controllers clearly outperform gradient-following ones. However, as the


performance of unconstrained controllers never exceeded the upper bound
for the gradient-following controllers, nothing can be concluded about the
nonoptimality of gradient-following controllers. However, the results sug-
gest that this upper bound obtained using the IB principle may also apply
to unconstrained controllers.

5 Representation of Positional Information

5.1 Introduction. In section 4.4 controllers were evolved to maximize


the amount of information they capture about the agent’s initial position.
What information about the initial position is captured? How do the con-
trollers capture the information? How is it represented in the state of the
controllers?
The evolved controllers are finite state automata. One such automaton is
shown in Figure 5 as an example. It is possible to analyze the static structure
of the automata, for example, through hierarchical decomposition (Dömösi
& Nehaniv, 2005). However, since the controllers were evolved to maximize
information flow, it is natural to analyze them from the perspective of
Shannon information, demonstrating that information-theoretic tools can
be used not only for evolving systems but also for analyzing them.
The following sections compare and supplement the insights obtained
using information-theoretic tools with prior knowledge about the analyzed
system. A later goal, however, is to analyze systems by relying solely on the
information-theoretic tools.
2402 A. Klyubin, D. Polani, and C. Nehaniv

In this section in particular, the representation of the agent’s initial po-


sition by the state of the controller at the end of the run is studied. The
analysis extends the results of Klyubin et al. (2004a) by also studying how
the representations change with small changes in the agent’s embodiment
(what the sensors capture about the environment and what the actuators
do in the environment).

5.2 Representations as Mappings. Representations are treated as prob-


abilistic mappings. To study the representation of the agent’s initial position
R0 in the state of its controller at the end of the run Mt , the mapping of
Mt onto R0 is visualized. The evolutionary experiment is rerun with d = 11
and t  = 15 to obtain spatially larger maps.
Mappings implemented by some of the best-evolved controllers are
shown in Figure 6. The controllers perform lossy compression of the agent’s
initial position: some of the information about the agent’s initial state R0 is
contained in the controller’s state M15 at the end of the run. Unconstrained
controllers successfully retain nearly the maximum amount of informa-
tion that fits in their memory. Gradient-following controllers retain less
information.
In some cases factorizable codes are produced: it is possible to find
projections of M that independently code for independent features of po-
sition, such as odd and even cell checkerboard pattern or left and right
half (see Figure 6, gf ++, |M| = 8). This suggests interesting parallels
with the results obtained for information flow maximization in neural
networks (Nadal & Parga, 1995; Bell & Sejnowski, 1995), where factorial
codes are produced as a result of maximizing information transmission
by a network of neurons. (Factorization is discussed in more detail in
section 8.)
Maximizing the information flow from R0 to M15 tends to create maps
that induce near-hard partitions over R0 . This is not surprising, since infor-
mation maximization between any two variables will tend to create hard
partitions. However, any permutation of a map with respect to the 11 × 11
set of initial position will preserve the mutual information between R0 and
M15 . Also, the form of the partitions is not important for capturing informa-
tion. Theoretically, there is no preferred grouping of initial positions. The
employed method of evolving the controller mappings is not biased in this
respect either. One could thus expect that the form of the partitions of initial
position would be completely different between different controllers due to
random factors of the evolutionary process.
However, it turns out that the evolved maps are not arbitrary. There are
certain repeating motifs in the maps: checkerboard patterns for gradient-
following controllers and spatially solid tiles for unconstrained controllers.
These regularities arise because not all of the maps from R0 to M15 are
possible. In short, the maps are constrained by and thus reflect the agent’s
embodiment and its information processing capabilities.
Information Flow in the Perception-Action Loop 2403

Figure 6: Representations of initial position (M15 → R0 ) used by best-evolved


controllers with different memory sizes. Types: g f —gradient following, u—
unconstrained, ++—cross-sensor and cross-actuator, +×—cross-sensor and
diagonal actuator, ××—diagonal sensor and diagonal actuator, ++ (××)—
evolved for ++ but mappings obtained with ××. Each box shows a mapping
from a controller’s state at time step 15 onto the agent’s initial position (11 ×
11 cells centered at the source). The intensity of each cell represents the proba-
bility of the agent having started in that cell: black—highest, white—zero. For
example, state 1 of the unconstrained controller (u + +, |M| = 2) at time step 15
means that the agent started in the lower-right half of the world. State 2 means
the opposite.

To illustrate where the constraints arise, it is useful to briefly study the


map R0 → M15 , the inverse of M15 → R0 . The map is a product of identical
time-invariant maps (Rt , Mt ) → (Rt+1 , Mt+1 ) (time evolution operator) pro-
jected onto the Mt component. Already this fact is a constraint. The form of
the time evolution operator is also constrained: the agent can move only into
2404 A. Klyubin, D. Polani, and C. Nehaniv

adjacent cells, the sensor captures only certain features of the world, and,
in the case of following the gradient, the agent’s controller is constrained to
strictly following the gradient.
The evolved maps R0 → M15 , and also M15 → R0 , are thus not arbitrary.
They reflect the structure of the agent’s sensorimotor apparatus, the corre-
lations between the actuator and the sensor—in other words, the agent’s
embodiment. The maps also reflect constraints on the agent’s information
processing capabilities. This is illustrated by the qualitative differences in
maps of gradient-following (constrained) and unconstrained controllers
(see Figure 6), as well as by quantitative differences between maps created
by controllers with different number of states.

5.3 Different Embodiment. In the previous experiment the agent’s


body/embodiment was kept constant, and it was the controller that was
allowed to change. To illustrate the idea that the evolved maps R0 → M15
reflect both the embodiment and the constraints on the controllers, an evo-
lutionary experiment was performed with different sensors and actuators
but with the same constraints on the controller.
The simplest change to sensors and actuators is to permute them. This
can be seen as changing the meaning of the sensoric states and actions. For
instance, the action “go west” can be changed to move the agent one cell
north. Correspondingly, the former “go north” action can be changed to
move the agent west. The evolutionary search is not biased toward a par-
ticular meaning of sensoric states and actions. It treats sensoric inputs and
actions as two unordered sets. Similarly, the fitness function, information
flow, being an information-theoretic quantity, is not influenced by the per-
mutations either. As a result, the evolutionary search is completely immune
to permutations of meaning externally assigned to elements of the sensor
and action sets.
What would happen if, instead of permuting sensoric states and ac-
tions, a new meaning were assigned to them in terms of what features
of the position the sensors capture and how the actions change the posi-
tion? In the previous experiment, the sensor was measuring the gradient
by comparing the signal strength in four cells, located north, east, south,
and west from the agent. Due to the layout, the sensor is called a cross
sensor. In addition to the cross sensor, a diagonal sensor, which compares
signal strength in four diagonally adjacent cells (northeast, southeast, south-
west, northwest), will be used in the following experiments. Similar to the
two types of sensors, the agent can have a cross (original) or a diagonal
actuator.
Two more evolutionary experiments were run optimizing the informa-
tion flow from R0 to M15 for the case of cross sensor and diagonal actua-
tor, and for the case of diagonal sensor and diagonal actuator. The result-
ing maps implemented by the best-evolved controllers are visualized in
Figure 6, type +× and ××, respectively.
Information Flow in the Perception-Action Loop 2405

The second case, diagonal sensor and diagonal actuator, is interesting


because the experiment is almost isomorphic to the original one rotated
by 45 degrees. The difference comes from the fact that the initial position
of the agent is still distributed around the source in a square exactly as
in the original experiment. The best-evolved mappings M15 → R0 for the
diagonal sensor and diagonal actuator case (see Figure 6, ××) are indeed
quite similar to the best-evolved mappings rotated by 45 degrees obtained
in the original experiment (see Figure 6, ++) where cross sensor and cross
actuator were used.
Due to this observation, it was interesting to see how the controllers
evolved for the original scenario (cross sensor, cross actuator) behaved in
the agent with diagonal sensor and diagonal actuator. It turns out that for
a small number of states (|M| ≤ 6) they produce original maps rotated by
45 degrees (see Figure 6, ++ (××)) and actually capture a little bit more
information about R0 than in the original scenario. For larger number of
states, the performance of gradient-following controllers drops compared
to the original embodiment, and the maps no longer represent original
features in a clean way (see Figure 5.2, |M| = 8 gf ++ ××).

5.4
. Discussion

5.4.1 Experiments. The mappings from the controller’s state at the end
of the run M15 to the agent’s initial position R0 show what information
about the agent’s initial position gets captured by different controllers.
Gradient-following and unconstrained controllers as a class capture dif-
ferent features of the initial position (see Figure 6). The main qualitative
difference is that unconstrained controllers divide the initial position into
spatially solid chunks. Neighboring initial positions are generally coded
by the same state of the controller at the end of the run. Representa-
tions facilitated by gradient-following controllers lack the solidity of the
chunks. These representations tend to have stripe or checkerboard patterns.
The differences in representations derive from the differences in informa-
tion processing capabilities (unconstrained versus gradient-following) and
embodiment.
The results may thus be viewed as having extended and generalized
Linsker’s work on the infomax principle (Linsker, 1988) (see section 4.1)
but with fewer assumptions: this letter does not assume linearity, it does
not assume a particular information processing substrate for the agent’s
controller, and it allows for more than just one time step for information
processing. The unconstrained controllers were even allowed to perform ac-
tions based on the past inputs. The results, however, are similar to Linsker’s:
without being told what is important about the agent’s initial position, evo-
lution found controllers that extract various features of the initial position
as a result of maximizing information preservation. Despite the agent hav-
ing no concept of geometry, most of the features, especially the ones created
2406 A. Klyubin, D. Polani, and C. Nehaniv

by unconstrained controllers, have strong spatial continuity properties due


to the way the agent perceives the world and acts in it.
5.4.2 Scalability. The main aim here was to find basic principles—hence
the minimal bias and high inefficiency of the method. A natural question is
how to scale the method up. More efficient methods of approximating infor-
mation flow could be devised for particular environments or computational
structures. A promising way of improving efficiency could be to factorize
(see section 8) the information flows, thus creating small and manageable
building blocks from which more complex hierarchical systems could be
composed.
5.4.3 Biological Relevance. It may be argued that natural evolution is in-
fluenced by many factors and constraints that have no relation to informa-
tion processing. However, assuming that the constraints take precedence,
Linsker’s infomax (Linsker, 1988) is a useful guide because more informa-
tion is likely to be advantageous given the constraints: information may
provide selective advantage (Howard, 1966; Polani, Nehaniv, Martinetz, &
Kim, 2006). Thus, if nothing is known (assumed) about the system in ad-
vance, infomax is a maximally unbiased approach: transmit as much infor-
mation as possible, since it is not known what information will be important
in later stages. If, however, there is bias in terms of what information is im-
portant, then that information should be transmitted, and hence evolution
will adapt accordingly by removing or reusing the unnecessary capacity.
The information capture experiment based on the perception-action loop
model demonstrates how under very limited constraints (limited sensors
and actuators), the infomax principle results in self-organization of the in-
formation processing in an unstructured system. Additional constraints are
expected not to reduce but rather modify the effect (cf. unconstrained ver-
sus gradient-following controllers). The minimum-assumption perception-
action loop model here, combined with infomax, provides an organizing
principle for information processing. However, infomax may be evolution-
arily more important, resulting in trade-offs with other constraints. In fact,
information may be a primary resource on par with energy: in the resting
state the human brain is responsible for around 20% of the total oxygen
consumption of the body (Kandel, Schwartz, & Jessell, 1991), and the fly
photoreceptors consume around 10% (Laughlin et al., 1998) of the total en-
ergy consumed. Whether information is a primary resource or whether it
is superseded by other constraints, the above indicates that infomax-like
principles may help in understanding the organization of information pro-
cessing in evolved natural and also artificial systems.

6 Representation of Time

Assume that during the experiment the state of the controller was ob-
served without knowing the time. Would the state of the controller contain
Information Flow in the Perception-Action Loop 2407

Figure 7: Relation of MT , T, and the Bayesian model of the perception-action


loop. MT is a readout of the process (Mt )t=0,1,...,15 at an unknown time step T.
Neither T nor MT has an effect on the dynamics of the loop.

information about the time step? It turns out that it would, provided the
probability distribution of the controller’s state depends on time.
Time is modeled as a random variable T with values t in T =
{0, 1, . . . , 15}. A special case is a peaked distribution where p(t) = 1 only for
a single t and is 0 for all others, resulting in the observed state of memory
Mt . However, in general, the distribution of T may be arbitrary. Denote the
readout of the process (Mt )t=0,1,...,15 observed at an unknown time step T by
a random variable MT (see Figure 7) with values mT . Here p(mT ) is given
by

p(mT ) = Pr (MT = mT ) = Pr (T = t)Pr (MT = mT |T = t), (6.1)
t∈T

where the probability distribution of the state of the readout for a particular
time step t is inherited from the memory variable at the time step

Pr (MT = mT |T = t) = Pr (Mt = mT ). (6.2)

The relation between MT , T, and the process (Mt )t=0,1,...,15 trivially follows
from equations 6.1 and 6.2:

Pr (MT = mT ) = Pr (T = t)Pr (Mt = mT ). (6.3)
t∈T

The amount of information captured by MT about the time step T is


I (MT ; T). Sensoric input can also be included, in which case the amount
of information about time available to the controller is I ((S, M)T ; T) ≥
I (MT ; T), where (S, M)T is defined analogous to MT . Informally, these quan-
tities can be interpreted as the amount of information the controller has on
the average about the time step it is at.
2408 A. Klyubin, D. Polani, and C. Nehaniv

Figure 8: p(t|(s, m)T )—representation of time by the best two- and nine-
state controllers. Vertical axis—(s, m)T − 2 × 4 states; horizontal axis—t −
{0, 1, . . . , 15}. The darker the cell, the higher the conditional probability
p(t|(s, m)T ). I = I ((S, M)T ; T) measured in bits.

Assuming a uniformly distributed T, the controllers evolved in the initial


position capture experiment capture from 0.1 to 0.84 bits of information
about time in their sensors and memory. For comparison, the global state of
the system (R, M)T captures between 0.93 and 2.00 bits of information about
time. The small amounts of the former are not surprising considering that
the controllers were not evolved for capturing information about time.11
To investigate how time is represented by the controllers, the mapping
(S, M)T → T will now be studied.12 For every tuple of sensoric input and
state of the controller, the mapping tells the probability of being at a par-
ticular time step. To illustrate the idea, Figure 8 shows the conditional
probability distribution p(t|(s, m)T ) for some of the controllers.
Different states (s, m)T capture different features of time. One type of
state captures local features of time. In Figure 8b, rows 2, 3, 6 and 7 (starting

11 If the controllers were evolved to capture as much information about time as possi-

ble, they would evolve into clocks. The simplest solution is to shut out the environment
and permute the controller’s state.
12 Strictly speaking, the reconstructed time is not metric time but an ordinal sequence.
Information Flow in the Perception-Action Loop 2409

from the top) are such examples. These states differentiate between odd
and even time steps. In Figure 8d there are states that capture time modulo
2, 3, and 4.
Another type of states captures global properties of time, such as being
close to the beginning, the middle, or the end of the experiment. Rows 1
and 4 in Figure 8 are an example.
There are also hybrid states, which capture both local and global proper-
ties of time. Row 1 and middle rows in Figure 8 are such states. They locate
the system in time in terms of being at a particular modulo 4 time step and
also in terms of being at the beginning or the end of the experiment.
The controllers were evolved to capture information about the agent’s
initial position at the end of 15 time steps. Why do the controllers capture
information about time? Is capturing information about time an evolution-
ary neutral side effect, or is it crucial to capturing information about the
agent’s initial position? For example, it may be that the processing and in-
tegration of information by the controller depend on time, or that different
methods of obtaining the information about the agent’s position are used
at different time steps, or that the controller represents the information in
a time-dependent way. The issue of time-dependence of representation is
addressed in the next section.

7 Time Dependence of Representation

Section 5 showed how M15 , the state of controller at the end of the experi-
ment, represents R0 , the agent’s position at the beginning of the experiment.
It turns out that the representation of the initial position in the controller’s
memory throughout the experiment is time dependent. For illustration
purposes the representation used by the best eight-state unconstrained con-
troller is shown in Figure 9.
The degree of time dependence of the representation of the initial posi-
tion R0 in the process (Mt )t=0,1,...,15 is quantified as the probabilistic depen-
dence of the mapping MT → R0 on the time step T, where MT is the readout
of the process at an unknown time T. (See section 6 for the definition of T
and MT .)
For every observed state mT of the controller, the mapping MT → R0 ,
characterized by the conditional probability distribution p(r0 |mT ), describes
where and with what probability the agent could have started given that
the controller was found in that state at an unknown time step.
If, in addition to knowing mT , the time step t is known, the mapping

p(r0 |mT , t) = Pr (R0 = r0 |MT = mT , T = t) = Pr (R0 = r0 |Mt = mT )


(7.1)
2410 A. Klyubin, D. Polani, and C. Nehaniv

Figure 9: Time dependence of the representation of the initial position R0 by


the process (Mt )t=0,1,...,15 of the best eight-state unconstrained controller. MT is
the readout of the process at an unknown time step T. Vertical axis: eight states
of the controller; horizontal axis: time step (0 to 15). Darker cells code for higher
probability of having started there. If the observation time step T is uniformly
distributed, then knowing whether the time step is odd or even greatly helps
determine what the controller’s states mean in terms of initial position.

can be used to find out where the agent could have started. For example,
in the case of the best evolved eight-state unconstrained controller (see
Figure 9) knowing whether the time step is odd or even helps to significantly
reduce the uncertainty in the agent’s initial position given the state of the
controller. The difference in how the two conditional distributions (time-less
and time-dependent) describe the initial position can be used to quantify
the dependence of the mapping MT → R0 on the time step T.
Here the degree of time dependence of the representation is quantified as
the Kullback-Leibler distance from the conditional distribution p(r0 |mT , t)
to the conditional distribution p(r0 |mT ):13
  

Dp(mT ,t) p(r0 |mT , t) p(r0 |mT ) = H(R0 |MT ) − H(R0 |MT , T)

= I (R0 ; T|MT ).

13 The method for measuring time dependence of a mapping between two random

variables is quite general. In fact, equation 7.2 applies to the general case of having three
arbitrary random variables X, Y, Z and measuring the dependency of the mapping
X → Y on Z:
 
D p(x,z) p(y|x, z)|| p(y|x) = I (Y; Z|X)
I (Y; Z|X) + I (X; Z) = I (X; Z|Y) + I (Y; Z).
Information Flow in the Perception-Action Loop 2411

The distance can be interpreted as the average increase in the information


about the initial position R0 due to knowledge of the time step T when the
controller’s state MT is known. For example, for the eight-state controller
shown in Figure 9, the time dependence is 0.57 bits, assuming a uniform
distribution for T.
The time-dependence measure is nonnegative. It is zero if and only if
knowing the time step does not improve the knowledge about the initial
position, that is, when the state of controller represents the initial position
in a time-independent way. The higher the measure, the higher the time
dependence of the representation.
The time-dependence measure is not symmetric with respect to R0 and
MT . However, the dependence of MT → R0 and its inverse, R0 → MT on T,
is related:

I (R0 ; T|MT ) + I (MT ; T) = I (MT ; T|R0 ) + I (R0 ; T). (7.2)

Since the initial position of the agent is not dependent on time, the last term,
I (R0 ; T), vanishes. The time dependence of the two representations
 and the
amount of information captured by the controller about time I (MT ; T) are
then related as

I (R0 ; T|MT ) + I (MT ; T) = I (MT ; T|R0 ). (7.3)

See appendix F for further discussion of time dependence.

8 Factorization of Representations

8.1 Rationale. In section 5, it was mentioned that some of the represen-


tations of the agent’s initial position by the state of the controller at the end
of the run shown in Figure 6 are factorizable. In this letter, factorization of
a representation means that the corresponding probabilistic mapping can
be represented as a product of two or more component mappings, such
that they capture independent pieces of the information captured by the
original mapping, creating an orthogonal coordinate system. Another way
of looking at the factorization of the representation of the initial position is
to think of it as factorization of the initial position itself. However, only cer-
tain features of the initial position are factorized. The mapping between the
initial position and the state of the controller at the end of the run defines
what features are relevant. This is more in line with the relevance-based
compression view adopted in the parallel bottleneck method (Friedman
et al., 2001) discussed in more detail in appendix G.

8.2 Problem Statement. Assume two random variables X and Y with a


given joint probability distribution p(x, y). Y represents X in some way. The
goal is to factorize this representation, that is, factorize Y with respect to X.
2412 A. Klyubin, D. Polani, and C. Nehaniv

Figure 10: Factorization of Y with respect to X into Z1 and Z2 . X and Y are


given. The task is to create two new variables, Z1 and Z2 , based on Y.

Y is to be split into two new random variables, Z1 and Z2 (see Figure 10),
such that they meet the following criteria.14
First, the two new variables should capture as much information about
X as possible. Hence, I (Z1 , Z2 ; X) should be as large as possible. Second,
the two variables should comprise an orthogonal coordinate system for X.
Hence, the two variables should be as independent as possible (I (Z1 ; Z2 )
close to zero), and the conditional probability distribution p(z1 , z2 |x) de-
scribing the projection of X onto Z1 and Z2 should be closely approximated
by the product of partial projections p(z1 |x) and p(z2 |x). The distance be-
tween p(z1 , z2 |x) and the product p(z1 |x) p(z2 |x) can be measured as the
Kullback-Leibler distance:
  

Dp(x) p(z1 , z2 |x) p(z1 |x) p(z2 |x) = I (Z1 ; Z2 |X) . (8.1)

The distance should be as close to zero as possible. Note that although the
distance is for conditional distributions, it still depends on the distribution
of X.
Minimizing I (Z1 ; Z2 |X) makes Z1 and Z2 as conditionally independent
given X as possible. This avoids synergetic or XOR-like coding schemes
where, taken separately, Z1 and Z2 contain less information about X than
when they are combined.15 (See section G.3 for more details.)
To summarize the task, given a joint distribution p(x, y), find two con-
ditional distributions p(z1 |y) and p(z2 |y) satisfying the following three
criteria:
1. I (Z1 , Z2 ; X) ≤ I (X; Y) is maximal.
2. I (Z1 ; Z2 ) ≥ 0 is minimal.
3. I (Z1 ; Z2 |X) ≥ 0 is minimal (goal: p(z1 , z2 |x) = p(z1 |x) p(z2 |x)).

14 Factorizing into three or more variables places more constraints on the solution. Two

variables are sufficient to start with. Later, each of the new variables can be factorized
further if necessary.
15 Synergy colloquially refers to the case when the whole is more than the sum of its

parts.
Information Flow in the Perception-Action Loop 2413

Appendix G discusses the details of the factorization method used in


this letter, as well as the relation to synergy and the parallel information
bottleneck method.

8.3 Factorization of the Representation of the Initial Position. The


mappings M15 → R0 created by the best controllers evolved for capturing
information about the initial position of the agent (see section 5) are factor-
ized here. In terms of the factorization task, R0 is X, and M15 is Y. Thus,
the information about R0 available in M15 was factorized into two new
variables, Z1 and Z2 , by means of mappings M15 → Z1 and M15 → Z2 . The
original mappings, as well as mappings from Z1 , Z2 , and Z back to initial
position R0 , are shown in Figure 11.
The factorization of the representation of initial position used by the best-
evolved eight-state gradient-following controller (see Figure 11, |M| = 8)
is a good example of the previous claims that these representations are
factorizable. In the first factorization, the two new variables are constrained
to be binary (|Z1 | = |Z2 | = 2). The first variable captures the initial position
in terms of being left or right from the source. The second variable captures
the position roughly in terms of being on either white or black cells of
a checkerboard pattern. Note that there is a vertical discontinuity in the
pattern in the middle.
The factorization is not at the ideal point (d = (0.008, 0.000, 0.000)) (see
appendix G) because 0.008 bit of the information about R0 captured by M
are not captured by the new variables. To capture this residual information,
one of the new variables was grown to have three states. The resulting
best factorization is shown in the middle of Figure 11 (8 → 3 × 2). There
only 0.001 bit of original information is lost. The representation of the initial
position employed by the second variable is exactly the same as used by the
first variable in the previous case. However, the first variable uses a proper
checkerboard coordinate system, without the vertical fault in the middle.
The controllers were originally evolved to capture at the end of the run
as much information about the agent’s initial position as possible. Alterna-
tively, this task may be seen as establishing a temporally extended infor-
mation flow from R0 to M15 . Factorization of the initial position through
the state of the controller at the end of the run can be seen as splitting
the information flow, albeit only at the end, into two or more independent
components. Later, each component can be processed separately.
Such factorization has several advantages. First, the state space of the
variables representing information is reduced due to compression. Second,
information is split into independent parts or features. The state spaces of
features are normally smaller than the state space of the original variables.
Compact representation may allow for more efficient processing, requiring
less space to store and transmit. Third, the fact that individual components
are independent of each other allows for their parallel independent pro-
cessing. At a later stage, the results of the parallel processing can be again
2414 A. Klyubin, D. Polani, and C. Nehaniv

Figure 11: Factorization of information about the initial position captured by


the best-evolved two-, three-, four-, and eight-state controllers into pairs of
binary (2 × 2) or ternary (3 × 3) variables. Mapping M15 → R0 is factorized
into Z1 → R0 and Z2 → R0 , and then reconstructed as (Z1 , Z2 ) → R0 . Each box
shows the mapping from a state (m or z) onto 11 × 11 cells of initial position.
The darker the cell, the higher the probability.

combined. Moreover, if individual components can be factorized further,


they could serve as starting points for hierarchies.

8.4 Factorization of Representations of Time. In this section, the infor-


mation about time available to the agent is factorized. To be more specific,
Information Flow in the Perception-Action Loop 2415

the information about T as seen through (S, M)T is factorized. T is assumed


to be uniformly distributed over {0, 1, . . . , 15}. In terms of the original def-
inition of the factorization task, T is X, and (S, M)T is Y. (S, M)T can be
thought of as containing information available to the agent’s controller.
The amount of information about time contained in internal state (S, M)T
of the best-evolved gradient-following controllers is quite small. The best
factorizations into two binary variables are shown in Figure 12a. Most of
the resolution of the new variables is concentrated on the first half of the
run. Because of the very low level of information about time in (S, M)T , the
factorized variables contain even less information. Consequently, it is hard
to interpret them.
Factorizations of the information about time in (S, M)T of the best-
evolved unconstrained controllers produce features distinguishing mostly
between odd and even time steps (see Figure 12b), although in the case
of the best three-state controller the first feature distinguishes between the
beginning and the rest of the run.

8.5 Discussion. Why is factorization of the agent’s initial position as


seen through the state of the controller at the end of the run (see section
8.3) much more successful than factorization of time as seen through the
agent’s internal state and the sensor (see section 8.4)?
To quickly rehash the structure of the factorization problem, random
variables X and Y are given, and two new random variables, Z1 and Z2 ,
are individually mapped from Y so that they capture independent pieces of
information about X. The factorization method used in this section can be
approximately thought of as consisting of two stages. First, similar to the
information bottleneck principle, it compresses the information contained
in Y about X into new variables, Z1 and Z2 , which have fewer states than
Y, hence resulting in a compact representation of the information. Second,
factorization splits this information into two independent variables.
The initial position of the agent as seen through the state of its controller
at the end of the run was factorized first (see section 8.3). The controllers
were evolved so that the state of the controller captured as much informa-
tion about the initial position as possible. Thus, the state of the controller
at the end of the run already is a compact, compressed representation of
the initial position. Factorization simply represents the captured informa-
tion in terms of two independent variables. In most cases the results of the
factorization algorithm were very close to the ideal point.
The factorization of the information about time as seen through the
sensor-controller state tuple (see section 8.4) was much further away from
the ideal point. In most cases the algorithm failed to capture a significant
share of the information about time that was to be factorized. The expla-
nation may be quite simple. Although the amount of information about
time captured by the agent’s internal state or the system’s global state is not
high, the information cannot be squeezed into two binary variables. It is
2416 A. Klyubin, D. Polani, and C. Nehaniv
Information Flow in the Perception-Action Loop 2417

“smeared” over a large number of states (|T | · |S| · |M|) and, being an av-
erage quantity, cannot be extracted without noticeable loss into two binary
variables. This hypothesis was confirmed by relaxing the factorization con-
straints and simply trying to extract as much of the information as possible
into just four states. The amount of the extracted information is a lower
bound for the distance from the ideal point the factorization algorithm can
achieve. Thus, the original factorization may have been less successful sim-
ply because the controllers were not evolved to capture information about
time in a convenient compressed way.
The factorization of the highly compressed information about the initial
position offers the advantage of viewing the captured features in terms
of two smaller independent features. The case of the best evolved eight-
state gradient-following controller in Figure 11 is a good example. There,
factorization creates two natural variables: left versus right and black versus
white cell of the checkerboard.
When factorizing the relatively scarce and widely “smeared” informa-
tion about time, the most useful feature of factorization is that relevance-
based compression is performed. This projects the information about time
from a high number of states to just a few. These are easier to interpret.
For example, it becomes obvious that the two most common features of
time captured are roughly (1) whether the time step is odd or even and (2)
whether roughly half of the run has elapsed.

9 Conclusion

An information-theoretic approach to analysis and construction of


perception-action loops was presented. The approach enables measuring
information flows between various parts of the loop through time, con-
structing perception-action loops with desired properties, and analyzing
the resulting behavior at the level of information.
To show the approach in action, a two-dimensional grid world with
an agent controlled by a finite state automaton and having access only
to a gradient sensor was presented. The agent’s sensor captures partial
information about the agent’s position. It was shown that it is possible to
capture more of the information by integrating sensoric input over time
and employing an evolved movement strategy.

Figure 12: Factorization of internal state (S, M)T of best-evolved controllers


with respect to time. Each section shows the pair of mappings Z1 → T and
Z2 → T created from (S, M)T → T. Four boxes show the probabilistic mappings
described by p(t|z1 ) and p(t|z2 ) from states of Z1 and Z2 onto the 16 states of
time t (left to right, from 0 to 15). Darker color codes for higher probability. Since
new variables Z are binary, p(t|Zn = 1) = 1 − p(t|Zn = 0)∀t∀ p.
2418 A. Klyubin, D. Polani, and C. Nehaniv

Controllers with a limited amount of memory were evolved to capture


as much information as possible about the agent’s initial position under
the constraints of the agent’s embodiment. To use the memory efficiently,
the controllers performed lossy compression, which resulted in near-hard
partitioning of the position with respect to the controller’s state.
From the information-theoretic perspective, these controllers were
evolved to create a temporally extended communication channel of maxi-
mum bandwidth between two temporally separate parts of the system: the
agent’s position at the beginning of the run and the controller’s state at the
end of it. The channel is created through interaction of the agent with the
environment. Compression, partitioning, and the movement strategy em-
ployed by the controllers are all induced by maximization of information
transfer (infomax) through the perception-action loop through time. This
is related to results for neural networks, where information transfer maxi-
mization resulted in phenomena such as feature detectors (Linsker, 1988),
factorial codes (Nadal & Parga, 1995; Bell & Sejnowski, 1995), and source
separation (Bell & Sejnowski, 1995), but is more general, since it allows ar-
bitrary temporally extended channels and removes assumptions about the
internal structure of the underlying information processing.
In the second part of this letter, a way to use information-theoretic tools
to analyze representations of space (initial position) and time by the best-
evolved controllers was presented. The intention was twofold: to provide
more insight into what information the controllers capture and to demon-
strate the power of several information-theoretic tools.
The representations of the agent’s initial position by the state of con-
trollers at the end of the experiment were studied. The representations
reflect some properties of the agent’s embodiment, the interaction and
constraints of the world, and the agent’s sensorimotor apparatus. Slight
changes to the agent’s actuator produce similarly slight changes in the
representations. In some cases, even old representations are suitable for
a slightly changed actuator. A method for factorizing representations was
demonstrated. It turned out that some of the representations factorize nicely
into smaller independent features. Factorized features, being independent
by definition, can serve as starting points for hierarchies if factorized further.
The amount of information captured by the agent about time was quan-
tified by treating time as a random variable, and the representation of time
in the agent’s controller was studied. A method for information theoreti-
cally measuring the time dependence of representations was demonstrated,
linking time dependence with the amount of information captured about
time. Last but not least, the representations of time were factorized to un-
derstand what is captured. The most common features of time captured are
(1) whether the time step is odd or even and (2) whether the first half of the
run has elapsed.
The applications of information-theoretic tools to the perception-action
loop are quite promising. They allow one to analyze information processing
Information Flow in the Perception-Action Loop 2419

even when one does not understand the underlying substrate. For example,
there is no requirement to assume finely grained components (e.g., neurons)
that perform the information processing. The model of computation is thus
more abstract and accommodating than that of traditional infomax. The
various information-theoretic tools presented here can provide a peek into
systems that do information processing in a nonobvious way.
Viewing information as the currency or essence of computation in the
perception-action loop provides a plausible mechanism, where a few as-
sumptions and constraints, such as embodiment with various bottlenecks
coming from the limited capacity of sensory and motor channels, and lim-
ited memory, can give rise to structures in a self-organized way. This paves
the way toward a principled and general understanding of the mechanisms
guiding the evolution of sensors, actuators, and information processing
substrates in nature and provides insights into the design of mechanisms
for artificial evolution.
The main contributions of this letter are (1) building up an information-
theoretic picture of the perception-action loop with a minimal set of assump-
tions, (2) maximizing information flow through the full perception-action
loop of an agent through time, as opposed to the well-established maxi-
mization of the flow through layers in neural networks, and (3) introducing
the use of information-theoretic tools to analyze the representations arising
from the maximization of information flow through the perception-action
loop.

Appendix A: Information Theory

In this section, the central information-theoretic notions used in this letter


are briefly introduced. For a detailed introduction to information theory
(Shannon, 1948), consult, for example, Cover and Thomas (1991).
A random variable can assume various values with various probabilities.
In this letter, exclusively discrete random variables are considered. Denote
random variables with uppercase letters (e.g., X), their sets of values with
calligraphic letters (e.g., X ), and their values with lowercase letters (e.g.,
x). Denote composite random variables by listing their elements inside
parentheses (e.g., (X, Y)) with values (x, y) from the set X × Y. By abuse of
notation, denote Pr (X = x), the probability that X assumes the value x, by
p(x). Similarly, the joint probability of X and Y is denoted by p(x, y) and
the conditional probability of X given Y by p(x|y).
The entropy of X, denoted by H(X), is defined as a measure of the
uncertainty of the probability distribution of X:


H(X) := − p(x) log2 p(x).
x∈X
2420 A. Klyubin, D. Polani, and C. Nehaniv

Entropy as well all other information-theoretic measures used in this let-


ter are measured in bits. Note that all of the information-theoretic measures
presented in this appendix are nonnegative.
The conditional entropy of X given Y, denoted H(X|Y), is defined as the
uncertainty of X knowing Y weighted by the probability of Y:

H(X|Y) := p(y)H(X|Y = y)
y∈Y
 
=− p(y) p(x|y) log2 p(x|y).
y∈Y x∈X

The mutual information between X and Y, denoted I (X; Y), is defined


as the average reduction in the uncertainty of X given Y:

I (X; Y) := H(X) − H(X|Y).

Kullback-Leibler distance (Kullback & Leibler, 1951) from a distribution


p1 (x) to a distribution p2 (x), denoted D( p1 || p2 ), is a generalized measure of
distance between the two distributions:
 p1 (x)
D( p1 || p2 ) = p1 (x) log2 .
x∈X
p2 (x)

D( p1 || p2 ) = 0 if and only if p1 is identical to p2 . The measure can assume


infinity if there exists x for which p1 (x) = 0 and p2 (x) = 0. Kullback-Leibler
distance is not actually a distance in the mathematical sense, in particular
since it is not symmetric with respect to p1 and p2 .
Kullback-Leibler distance from a conditional distribution p1 (x|y) to a
conditional distribution p2 (x|y) is defined as the sum of Kullback-Leibler
distances from p1 (x|Y = y) to p2 (x|Y = y) weighted by the probability of Y:

   

Dp(y) ( p1 || p2 ) = p(y)D p1 (x|Y = y) p2 (x|Y = y)
y∈Y

  p1 (x|y)
= p(y) p1 (x|y) log2 .
y∈Y x∈X
p2 (x|y)

Appendix B: Bayesian Network Model of the Agent’s Controller

The agent’s controller selects actions based on sensoric input and its own
memory. The assumptions about the agent’s controller are minimal: the
controller is finite time and finite state. Given the sets of sensoric states S,
actions A, and the memory states of the controller M, any such controller
Information Flow in the Perception-Action Loop 2421

Figure 13: Causal Bayesian network models of an agent’s controller. S—state


of the sensor; A—action performed by the actuator; M—state of the memory of
the controller. (a) General model. (b) Model with an extra causality assumption.
The new state of memory is influenced by the choice of action.

can be viewed as performing a probabilistic mapping S × M → A × M.


The interpretation is that the controller picks an action and the next state of
its memory based on the current sensoric input and the current state of its
memory.
The causal Bayesian network model for the general case of any such
probabilistic mapping is as shown in Figure 13a. To simplify the diagrams,
in the main body of the letter a different model shown in Figure 13b was
used. This model, although simpler to understand at first glance, contains
an extra assumption that the new state of memory can be influenced by
the choice of action. In general, this need not be the case. For example, it
is easy to envisage a controller where it is the action that is influenced by
the new state of memory. The general model, shown in Figrue 13a, on the
other hand, does not assume any particular causality between actions and
the new state of the controller’s memory.

Appendix C: Measuring Information Flows

The goal of the information capture experiment in section 4 is to evolve


agent controllers that maximize the information flow from the agent’s initial
position to the memory of the controller at a later time. To calculate the
amount of information flow, the joint probability distribution of the parts
of the system that are the source and the destination of information flow, is
required. Here the relevant parts are the agent’s initial position R0 and M15 ,
the state of memory of the agent’s controller at time step 15.
Although the joint distribution could be obtained from the Bayesian
network analytically, here Monte Carlo simulations are used for estimating
the probability distributions. The system is started in an initial state drawn
from the probability distribution of the initial state (see section 4.4 for an
example), then iterated for the required number of time steps after which
data of interest are gathered. The process is repeated until the estimates of
probability distribution of interest are stable enough.
2422 A. Klyubin, D. Polani, and C. Nehaniv

To spot and avoid undersampling, for each of the possible initial states of
the system, 256 or more samples are produced depending on the quantities
measured. The number of samples is then increased by a factor of at least
16 to check whether the quantities of interest remain stable.

Appendix D: Method of Evolution

In the information capture experiment in section 4, evolution is used as a


search in the space of controllers. The aim is not to develop new evolution-
ary computation techniques but to search for controllers with the maximum
information flow from the agent’s initial position to its controller’s memory
at a fixed time step. To do that in a straightforward, transparent, and unbi-
ased way, a minimal setup is employed. This is to emphasize that nothing
in the general approach is specific to the particular model or the search
methods employed.
The controller is evolved. The controller is described by a probabilistic
time-invariant mapping (St , Mt ) → (At , Mt+1 ) (see section 4.3). Evolving the
controller means evolving the mapping for fixed sizes of state (memory),
sensoric input, and action sets.
The search is limited to deterministic mappings only. To any pair of
sensoric input and state (st , mt ) there corresponds exactly one action and
new state (a t , mt+1 ). A mutation is performed picking a pair of (st , mt ) at
random with uniform probability from the set S × M and setting its image
to a pair (a t , mt+1 ) picked at random with uniform probability from the
set A × M. This type of mutation is unbiased in the sense that it is not
designed for a particular fitness function. Mutation is the only evolutionary
operator employed to generate variation.
Fitness is evaluated by letting the controller control the agent. The fitness
function is the amount of information flow from the agent’s initial position
R0 to the memory of the controller Mt at a later time step. The population is
initialized with five randomly generated controllers. After each generation,
five best controllers are kept in the population and produce five offspring,
each by mutation. Thus, the size of the population is between 5 and 30.
To speed up the search, ideas from integer programming (Glover, 1986)
are used: (1) random shakeup—the number of mutations applied to each
offspring is uniformly distributed between 1 and 1 + (G mod 20), where G
is the generation, and (2) permanent tabu list—offspring controllers that
have been evaluated before or are present in the population are discarded.
Additionally, at least five separate evolutionary runs are performed to sam-
ple different solutions.

Appendix E: Gradient Following as Initial Position Capture

This section proves the assertion in section 4.4 that when the agent follows
the gradient, most of the information about the agent’s initial position passes
through its sensors.
Information Flow in the Perception-Action Loop 2423

Denote the agent’s initial position with a random variable R0 . The en-
tropy of initial position is H(R0 ). This is the maximum amount of informa-
tion there is to know about the initial position.
Assume that the agent follows the gradient for t time steps. Denote the
agent’s position after t time steps with a random variable Rt .
Denote the sequence of sensoric input up to but excluding the time step
t with a random variable St−1c
= (S0 , S1 , . . . , St−1 ).
The goal is to prove that
 c 
I St−1 ; R0 = H(R0 ) − , (E.1)

where  is a small quantity.


Denote the sequence of actions the agent performed up to but excluding
the time step t with a random variable Act−1 = (A0 , A1 , . . . , At−1 ). Using
standard information-theoretic equalities, the amount of information the
sequence of actions Act−1 contains about the agent’s initial position R0 can
be expressed as
     
I Act−1 ; R0 = H(R0 ) − I R0 ; Rt |Act−1 − H R0 |Rt , Act−1 . (E.2)

All actions move the agent deterministically over the grid. Given a final
c
position rt , the sequence of actions performed, a t−1 , identifies the exact
starting position r0 . This means that the initial position is a deterministic
function of the final position and the sequence of actions:
 
H R0 |Rt , Act−1 = 0. (E.3)

Equation E.2 becomes


   
I Act−1 ; R0 = H(R0 ) − I R0 ; Rt |Act−1 . (E.4)

When the agent follows the gradient, it does not end up exactly at the
source. Because there is no “do nothing” action, once the agent is at the
source, it cannot stay there and moves into one of the four adjacent cells at
the next time step. One time step later, it comes back following the gradient.
Thus, after a certain amount of time, the agent cycles through two states:
(1) being in the center and moving in one of the four randomly chosen
directions and (2) being in one of the four cells adjacent to the central cell
and moving back to the central cell.
Due to this fact, given enough time, the tails of sequences of actions will
consist of a repeating pattern: a random action, then the action opposite to
it—for example, a tc = (. . . , ↑, ↓, →, ←, ↑). When the agent is at the central
cell, it enters the loop of performing an action followed by its counteraction.
When not in the central cell, the agent performs an action followed by its
2424 A. Klyubin, D. Polani, and C. Nehaniv

counteraction only in one situation: when the agent is in a cell adjacent to


the center.
The following two sequences are examples of what the agent can per-
form if it starts in the cell adjacent from the left to the central cell:
(→, ←, →, ←, →, ←), (→, ←, →, ↑, ↓, ↑). The first sequence consists of
the same repeating tuple of action and its counteraction. In principle, the
same sequence of actions can be performed by the agent starting from the
central cell. The second sequence of action cannot, because the agent must
follow the gradient consistently. Thus, only a tuple of action and its counter-
action followed by a tuple of a different action and its counteraction signify
that the agent has reached the center. Sequences of one and the same tuple
of action and its counteraction do not mean that the agent has reached the
center. However, the probability of such sequences decreases exponentially
with their length.
The above observation means that the higher the t, the more the sequence
of actions determines where the agent finished, and, hence, where it started.
This means that H(R0 |Act−1 ) tends to zero. Since

     
I R0 ; Rt |Act−1 = H R0 |Act−1 − H R0 |Act−1 , Rt (E.5)
 
≤ H R0 |Act−1 , (E.6)

it follows that
 
I R0 ; Rt |Act−1 → 0 (E.7)

and from equation E.4,


 
I Act−1 ; R0 → H(R0 ). (E.8)

When the agent follows the gradient, its actions are completely deter-
mined by the sensoric input. The sequence of actions is in one-to-one cor-
respondence with the sequence of sensoric input. Hence, the sequence of
actions contains the same amount of information about the initial position
as the sequence of sensoric input:
 c   
I St−1 ; R0 = I Act−1 ; R0 . (E.9)

Thus,
 c 
I St−1 ; R0 → H(R0 ). (E.10)

with increasing t.
Information Flow in the Perception-Action Loop 2425

Appendix F: Examples of Time Dependence

F.1 Clock Example. It is possible to imagine a case where the mapping


MT → R0 is time independent (first term in equation 7.3 is 0), whereas the
reverse mapping R0 → MT is time dependent. Assume a controller with
pos
two substates: M = (Mpos , Mclock ). Substate MT permanently encodes some
information about the agent’s initial position (mpos is constant in time), and
substate MTclock captures some information about time, for example, counts
modulo some number.
The combined state (Mpos , Mclock ) of the controller will be changing with
time because Mclock does. Nevertheless, the mapping (Mpos , Mclock )T → R0
pos
will be time independent, with all pairs (mT , ∗) mapping to the same subset
of initial position. The reverse mapping R0 → (Mpos , Mclock )T will be time
dependent because the substate Mclock changes with time. Knowing only
pos
the initial position predicts only MT . Without knowledge of time, MTclock
remains completely random. In this scenario, the time dependence of R0 →
(Mpos , Mclock )T is exactly the amount of information about time T contained
in MTclock :

   
I (Mpos , Mclock )T ; T|R0 = I (Mclock )T ; T . (F.1)

F.2 XOR Example. At the other extreme of the time-dependence scale


lies the case where both mappings, MT → R0 and R0 → MT , are time de-
pendent, though the state of the controller does not contain any information
about time. Consider any two-state controller that is started in state 1 if the
agent is above the signal source, and in state 2 otherwise. At each time step,
regardless of the sensoric input, the controller permutes its state: 1 → 2,
2 → 1.
In this case, the state of the controller contains no information about time
I (MT ; T) = 0. For instance, the fact that the state is 1 means that the agent
started above the source provided the time step is odd, and that the agent
started elsewhere provided the time step is even. Thus, without knowledge
of the time step, the state of the controller does not tell anything about the
initial position of the agent.
The coarsely grained initial position (above the source or not), time
step T modulo 2, and state of the controller M are in an XOR rela-
tionship. Knowing only one of the variables does not provide informa-
tion about any of the other two. Knowing the states of any two vari-
ables provides full information about the third. The mapping MT → R0
maps every state mT uniformly onto initial position Z2 . However, with
the knowledge of time, the mapping becomes more deterministic, map-
ping state 1 onto one half of Z2 and state 2 onto the other. Simi-
larly, the reverse mapping R0 → MT is made completely deterministic by
knowing whether the current time step is odd or even. Agreeing with
2426 A. Klyubin, D. Polani, and C. Nehaniv

equation 7.3, the time dependence of both mappings is identical since


I (MT ; T) = 0:

I (R0 ; T|MT ) = I (MT ; T|R0 ) = 1. (F.2)

F.3 Factoring Time Out. In the light of the two examples above, it is
possible to give an interpretation to equation 7.3, linking time dependence
of the mappings between R0 and MT and the amount of information about
time contained in MT . The state of the controller MT can be thought of as
containing two pieces of information about time. One is available without
reference to any other variable; its amount is quantified by I (MT ; T). The
other is “encrypted” through R0 . Knowledge of R0 provides I (R0 ; T|MT )
more bits of information about time. The total amount of information about
time extractable from MT with the help of R0 is then I (MT ; T|R0 ).
One could attempt to factor MT into two variables: one containing infor-
mation about time, the other containing information about the rest of MT . In
the clock example above, this type of factorization is easy. The new variables
pos
correspond to MTclock and MT by definition. In the XOR example, though,
it is not possible to split MT due to the fact that all information about time
in MT is “encrypted” through R0 and is thus not “freely” available without
the knowledge of R0 . The term I (MT ; T) in equation 7.3 is the upper bound
to the amount of information about time that can be factored out of MT .
Factoring out or removing information about time from MT reduces
the time dependence of the R0 → MT mapping by exactly the amount of
information about time removed. However, this still does not guarantee that
the mapping will become time independent. At best the time dependence
of R0 → MT can be lowered to the time dependence of its inverse, MT →
R0 . However, the time dependence of the latter cannot be influenced by
removing information about time from MT .

Appendix G: Factorization: Details

G.1 Ideal Solutions. An ideal situation is when (1) all of the information
about X contained in Y is captured by (Z1 , Z2 ), (2) Z1 and Z2 are indepen-
dent, and (3) the mapping X → (Z1 , Z2 ) is the product of the mappings
X → Z1 and X → Z2 . The ideal point in terms of the left-hand side values
of the three criteria above is thus (I (X; Y), 0, 0).
It should be noted that in case Z1 or Z2 is large enough to accommodate
all the states of Y, it is possible to create a trivial factorization, such as where
Y → Z1 is information preserving and Y → Z2 preserves no information
about Y. This trivial factorization meets the ideal point criterion but is
not interesting. A similarly trivial factorization is known for numbers: any
number can be factorized into the product of itself and 1. To avoid trivial
factorization, both |Z1 | and |Z2 | were made smaller than |Y|.
Information Flow in the Perception-Action Loop 2427

G.2 The Method. Each solution is a tuple of conditional distributions


p(z1 |y) and p(z2 |y). Comparing different solutions requires trading off be-
tween the three criteria above. As there is no known precise method for
finding an optimal factorization, the following method was used.
When the number of states for the two new random variables Z1 and
Z2 is given, the goodness or fitness of a tuple is defined in terms of how
well it satisfies the three criteria above. The ideal situation is where all of
the information about X contained in Y is captured by Z1 and Z2 , and the
left-hand side of the other two criteria is zero. The fitness is defined as the
Manhattan distance in the criteria space from the ideal point,

d = (c − I (Z1 , Z2 ; X)) + I (Z1 ; Z2 ) + I (Z1 ; Z2 |X), (G.1)

where c = min(I (X; Y), log2 |Z1 ||Z2 |) is the maximum amount of informa-
tion that Z1 and Z2 could in principle capture about X. c is a constant for all
tuples with same sizes of Z1 and Z2 . The goal is to find a tuple with minimal
distance from the ideal point.
Here hill climbing is employed in the space of the tuples of mappings
( p(z1 |y), p(z2 |y)) to search for the best tuple. The search is restricted to de-
terministic mappings. The best solution is chosen from five independent
runs of the hill climbing search. Each search run is started with a randomly
chosen tuple and run for 50,000 steps. At each step, the tuple is modified
1 + (t mod 20) times, where t is the step. A modification is performed by
randomly choosing one of the two mappings and then deterministically
mapping a randomly chosen state y to a randomly chosen state z1 or z2 ,
depending on which of the two mappings is modified. If after the modifi-
cations the new tuple is closer to the ideal point than the old one, the tuple
is used as a base for the next step; otherwise, the new tuple is discarded,
and the old one is used at the next step.

G.3 Relation to Synergy. Synergy colloquially refers to the case when


the whole is more than the sum of its parts. In neural networks, synergy
refers to an ensemble of neurons providing more information about a
stimulus than the sum of the information provided by neurons individ-
ually (Brenner, Strong, Koberle, Bialek, & de Ruyter van Steveninck, 2000;
Schneidman et al., 2003).
Synergy can also be applied to the factorization problem presented in
section 8. The total amount of information about X contained in Z = (Z1 , Z2 )
can be expressed as

I (Z; X) = I (Z1 ; X) + I (Z2 ; X) + I (Z1 ; Z2 |X) − I (Z1 ; Z2 ).

I (Z1 ; X) and I (Z2 ; X) is the amount of information about X contained in Z1


and Z2 , respectively. The extra term I (Z1 ; Z2 |X) − I (Z1 ; Z2 ), called synergy
2428 A. Klyubin, D. Polani, and C. Nehaniv

(Brenner et al., 2000; Schneidman et al., 2003),

Syn(Z1 , Z2 ) = I (Z1 , Z2 ; X) − I (Z1 ; X) − I (Z2 ; X)


= I (Z1 ; Z2 |X) − I (Z1 ; Z2 ),

can be seen as the amount of information about X that is “jointly” contained


in Z1 and Z2 in an “encrypted” form and can be extracted only when both
variables are known. Synergy can be negative. When synergy is negative,
information about X is contained in Z1 and Z2 in a redundant way.
An extreme example of synergetic representation of information
is the binary XOR case, when x = z1 ⊕ z2 and Z is uniformly dis-
tributed. I (Z1 ; X) = I (Z2 ; X) = 0 = I (Z1 ; Z2 ) = 0 bit, whereas I (Z; X) =
I (Z1 ; Z2 |X) = Syn(Z1 , Z2 ) = 1 bit. Knowing either Z1 or Z2 does not tell
anything about X, whereas knowing both Z1 and Z2 tells everything.
Conditions 2 and 3 of the factorization task minimize I (Z1 ; Z2 ) and
I (Z1 ; Z2 |X), respectively. When these two terms reach zero, synergy also
becomes zero. In this case, the resulting representation of X by Z1 and Z2 is
neither synergetic nor redundant. This is exactly what is desired, since Z1
and Z2 should be independent components or orthogonal coordinates of X.

G.4 Relation to Parallel Information Bottleneck Method. The factor-


ization task is similar in spirit to the parallel bottleneck case of the multi-
variate information bottleneck (Friedman et al., 2001). In this section, the
similarities and the differences between the two tasks are pointed out briefly.
To facilitate easier comparison, this letter’s notation for the parallel bottle-
neck problem is used instead of the original notation from Friedman et al.
(2001). The random variable Z = (Z1 , Z2 ) is introduced to save space.
The parallel bottleneck problem is stated as following. Given a joint dis-
tribution p(x, y), find two conditional distributions p(z1 |y) and p(z2 |y) such
that the two random variables Z1 and Z2 compress Y in “parallel,” predict-
ing X as much as possible. The Bayesian network of relations between the
random variables X, Y, Z1 and Z2 is identical to the Bayesian network from
the factorization problem in Figure 10.
The quality of a solution is quantified by the following Lagrangian, which
should be minimized in the space of tuples ( p(z1 |y), p(z2 |y)) to find a good
solution,

L(1) = −β I (Z; X) + I (Z1 ; Z2 ) + I (Z; Y), (G.2)

where β is a trade-off parameter that specifies the relative importance of


preserving information about X.
The parallel bottleneck problem is very much similar to the factoriza-
tion problem described in this article. The relations between the random
variables are identical. The goal is almost identical too. Here the aim is to
Information Flow in the Perception-Action Loop 2429

factorize the information about X captured by Y into new variables Z1 and


Z2 . In parallel bottleneck, this requirement is less explicit.
Both methods express the goodness of a solution (a tuple of mappings
p(z1 |y) and p(z2 |y)) in terms of a single measure. Here it is the Manhattan
distance from the ideal point (see equation G.1), and in the parallel bottle-
neck case, it is the Lagrangian (see equation G.2). Without the constant term
c, the distance specified in equation G.1 is

d = −I (Z; X) + I (Z1 ; Z2 ) + I (Z1 ; Z2 |X). (G.3)

X, Y, and Z form a Markov chain X → Y → Z (see Figure 10) and therefore


I (Z; Y) = I (Z; X) + I (Z; Y|X). Hence, the Lagrangian from equation G.2
can be rewritten as

L(1) = −(β − 1)I (Z; X) + I (Z1 ; Z2 ) + I (Z; Y|X). (G.4)

Now two differences between the factorization and the parallel bottle-
neck are apparent. First, the factorization method in this article can be seen
as using a constant parallel bottleneck β = 2. Second, the last term in the
two equations is different. The latter needs closer attention.
I (Z1 ; Z2 |X) = 0 requires that Z1 and Z2 are conditionally independent
given X. This avoids synergetic, XOR-like coding schemes where Z1 and
Z2 individually cannot be used to obtain information about X, where only
their combination is helpful. I (Z; Y|X) = 0 requires that Z and Y share no
information other than information about X. In other words, this requires
that Z1 and Z2 contain information only about X.
The two requirements are different. However, they are related. Keeping
in mind that Z = (Z1 , Z2 ) and that Z1 and Z2 are obtained from Y using
independent mappings,

I (Z; Y|X) = I (Z1 ; Y|X) + I (Z2 ; Y|X) − I (Z1 ; Z2 |X). (G.5)

Minimizing only I (Z1 ; Z2 |X) allows Z1 and Z2 to contain “irrelevant”


information—information about Y that is not information about X.
Parallel bottleneck, on the other hand, attempts to forbid such irrelevant
information.
Interestingly, I (Z; Y|X) ≥ I (Z1 ; Z2 |X).16 As a result of nonnegativity
of mutual information, if I (Z; Y|X) = 0, then I (Z1 ; Z2 |X) = 0. Hence, a

16 I (Z; Y|X) = I (Z ; Y|X) + I (Z ; Y|X) − I (Z ; Z |X). Since I (Z ; Z |X) ≤ I (Z ; Y|X)


1 2 1 2 1 2 1
and I (Z1 ; Z2 |X) ≤ I (Z2 ; Y|X):

I (Z; Y|X) ≥ I (Z1 ; Z2 |X) + I (Z1 ; Z2 |X) − I (Z1 ; Z2 |X)


≥ I (Z1 ; Z2 |X).
2430 A. Klyubin, D. Polani, and C. Nehaniv

minimum of the parallel bottleneck Lagrangian in the case of I (Z; Y|X) = 0


is also an optimal solution for the factorization method in this
article.
An important advantage of the multivariate bottleneck method, of which
parallel bottleneck is just an example, is that it provides an iterative opti-
mization algorithm for finding a good solution. However, the algorithm is
designed only for certain types of Lagrangians. The distance d (see equation
G.3) used in this article seems to be unsuitable for the algorithm.

Acknowledgments

We thank the Condor Team from the University of Wisconsin, whose high-
throughput computing system Condor provided a convenient way to run
large numbers of simulations on ordinary workstations.

References

Ashby, W. R. (1952). Design for a brain. New York: Wiley.


Ashby, W. R. (1956). An introduction to cybernetics. London: Chapman & Hall.
Attneave, F. (1954). Some informational aspects of visual perception. Psychological
Review, 61(3), 183–193.
Avraham, H., Chechik, G., & Ruppin, E. (2003). Are there representations in embod-
ied evolved agents? Taking measures. In W. Banzhaf, T. Christaller, P. Dittrich,
J. T. Kim, & J. Ziegler (Ed.), Proceedings of the 7th European Conference on Artificial
Life (ECAL) (pp. 743–752). Berlin: Springer-Verlag.
Ay, N. (2002). Locality of global stochastic interaction in directed acyclic networks.
Neural Computation, 14(12), 2959–2980.
Ay, N., & Krakauer, D. C. (2007). Geometric robustness theory and biological net-
works. Theory in Biosciences, 125(2), 93–121.
Ay, N., & Wennekers, T. (2003). Dynamical properties of strongly interacting Markov
chains. Neural Networks, 16(10), 1483–1497.
Barlow, H. B. (1959). Possible principles underlying the transformations of sensory
messages. In W. A. Rosenblith (Ed.), Sensory communication: Contributions to the
Symposium on Principles of Sensory Communication (pp. 217–234). Cambridge, MA:
MIT Press.
Barlow, H. B. (1963). The coding of sensory messages. In W. H. Thorpe & O. L.
Zangwill (Eds.), Current problems in animal behaviour (pp. 331–360). Cambridge:
Cambridge University Press.
Barlow, H. B. (1989). Unsupervised learning. Neural Computation, 1(3), 295–311.
Barlow, H. B. (2001). Redundancy reduction revisited. Network: Computation in Neural
Systems, 12(3), 241–253.
Bell, A. J., & Sejnowski, T. J. (1995). An information-maximization approach to blind
separation and blind deconvolution. Neural Computation, 7(6), 1129–1159.
Bialek, W. (1987). Physical limits to sensation and perception. Ann. Rev. Biophys.
Biophys. Chem., 16, 455–478.
Information Flow in the Perception-Action Loop 2431

Bouman, M. A. (1959). History and present status of quantum theory in vision. In


W. A. Rosenblith (Ed.), Sensory communication: Contributions to the Symposium on
Principles of Sensory Communication (pp. 377–401). Cambridge, MA: MIT Press.
Brenner, N., Strong, S. P., Koberle, R., Bialek, W., & de Ruyter van Steveninck, R. R.
(2000). Synergy in a neural code. Neural Computation, 12(7), 1531–1552.
Cannon, W. B. (1939). The wisdom of the body. New York: Norton.
Cover, T. M., & Thomas, J. A. (1991). Elements of information theory. New York: Wiley.
Der, R., Steinmetz, U., & Pasemann, F. (1999). Homeokinesis—a new principle to back
up evolution with learning. In M. Mohammadian (Ed.), Computational intelligence
for modelling, control, and automation (pp. 43–47). Amsterdam: IOS Press.
Dömösi, P., & Nehaniv, C. L. (2005). Algebraic theory of finite automata networks: An
introduction. Philadelphia: Society for Industrial and Applied Mathematics.
Friedman, N., Mosenzon, O., Slonim, N., & Tishby, N. (2001). Multivariate informa-
tion bottleneck. In Uncertainty in Artificial Intelligence: Proceedings of the Seventeenth
Conference (UAI-2001) (pp. 152–161). San Francisco: Morgan Kaufmann.
Gibson, J. J. (1979). The ecological approach to visual perception. Boston: Houghton
Mifflin.
Glover, F. (1986). Future paths for integer programming and links to artificial intel-
ligence. Computers and Operations Research, 13(5), 533–549.
Howard, R. A. (1966). Information value theory. IEEE Transactions on Systems Science
and Cybernetics, SSC-2, 22–26.
Hutchins, E. (1995). How a cockpit remembers its speeds. Cognitive Science, 19(3),
265–288.
Kandel, E. R., Schwartz, J. H., & Jessell, T. M. (1991). Principles of neural science (3rd
ed.). New York: McGraw-Hill.
Kirsh, D., & Maglio, P. (1994). On distinguishing epistemic from pragmatic action.
Cognitive Science, 18(4), 513–549.
Klyubin, A. S., Polani, D., & Nehaniv, C. L. (2004a). Organization of the informa-
tion flow in the perception-action loop of evolved agents. In R. S. Zebulum, D.
Gwaltney, G. Hornby, D. Keymeulen, J. Lohn, & A. Stoica (Eds.), Proceedings of
2004 NASA/DoD Conference on Evolvable Hardware (pp. 177–180). Los Alamitos,
CA: IEEE Computer Society.
Klyubin, A. S., Polani, D., & Nehaniv, C. L. (2004b). Organization of the information
flow in the perception-action loop of evolved agents (Tech. Rep. 400). Hertfordshire:
Department of Computer Science, University of Hertfordshire.
Klyubin, A. S., Polani, D., & Nehaniv, C. L. (2004c). Tracking information flow
through the environment: Simple cases of stigmergy. In J. Pollack, M. Bedau,
P. Husbands, T. Ikegami, & R. A. Watson (Eds.), Artificial Life IX: Proceedings of
the Ninth International Conference on the Simulation and Synthesis of Living Systems
(pp. 563–568). Cambridge, MA: MIT Press.
Kullback, S., & Leibler, R. A. (1951). On information and sufficiency. Annals of Math-
ematical Statistics, 22(1), 79–86.
Laughlin, S. B., de Ruyter van Stevenick, R. R., & Anderson, J. C. (1998). The metabolic
cost of neural information. Nature Neuroscience, 1(1), 36–41.
Linsker, R. (1988). Self-organization in a perceptual network. IEEE Computer, 21(3),
105–117.
Millikan, R. G. (2002). Varieties of meaning. Cambridge, MA: MIT Press.
2432 A. Klyubin, D. Polani, and C. Nehaniv

Nadal, J.-P., & Parga, N. (1995). Information transmission by networks of non linear
neurons. In Proc. of the Third Workshop on Neural Networks: From Biology to High
Energy Physics (1994, Elba, Italy). International Journal of Neural Systems (Suppl.),
153–157.
Nolfi, S., & Marocco, D. (2002). Active perception: A sensorimotor account of object
categorization. In B. Hallam, D. Floreano, J. Hallam, G. Hayes, & J.-A. Meyer
(Eds.), From Animals to Animats 7, Proceedings of the VII International Conference on
Simulation of Adaptive Behavior (pp. 266–271). Cambridge, MA: MIT Press.
Pearl, J. (2001). Causality: Models, reasoning, and inference. Cambridge: Cambridge
University Press.
Polani, D., Nehaniv, C. L., Martinetz, T., & Kim, J. T. (2006). Relevant information
in optimized persistence vs. progeny strategies. In L. M. Rocha, M. Bedau, D.
Floreano, R. Goldstone, A. Vespignani, & L. Yaeger (Eds.), Proc. Artificial Life X
(pp. 337–343). Cambridge, MA: MIT Press.
Powers, W. T. (1974). Behavior: The control of perception. London: Wildwood House.
Schmidhuber, J. (1992). Learning factorial codes by predictability minimization.
Neural Computation, 4(6), 863–879.
Schmidhuber, J., Eldracher, M., & Foltin, B. (1996). Semilinear predictability mini-
mization produces well-known feature detectors. Neural Computation, 8(4), 773–
786.
Schneidman, E., Bialek, W., & Berry II, M. J. (2003). Synergy, redundancy, and inde-
pendence in population codes. Journal of Neuroscience, 23(37), 11539–11553.
Schneier, B. (1995). Applied cryptography: Protocols, algorithms, and source code in C
(2nd ed.). New York: Wiley.
Shannon, C. E. (1948). A mathematical theory of communication. Bell System Technical
Journal, 27, 379–423.
Tishby, N., Pereira, F. C., & Bialek, W. (1999). The information bottleneck method.
In Proceedings of the 37th Annual Allerton Conference on Communication, Control,
and Computing (pp. 368–377). Urban-Champaign: Department of Electrical and
Computer Engineering, University of Illinois.
Touchette, H., & Lloyd, S. (2000). Information-theoretic limits of control. Phys. Rev.
Lett., 84, 1156.
Touchette, H., & Lloyd, S. (2004). Information-theoretic approach to the study of
control systems. Physica A, 331(1–2), 140–172.
Usher, M. (2001). A statistical referential theory of content: Using information theory
to account for misrepresentation. Mind and Language, 16(3), 311–334.
Wennekers, T., & Ay, N. (2005). Finite state automata resulting from temporal infor-
mation maximization. Neural Computation, 17(10), 2258–2290.

Received November 21, 2005; accepted October 16, 2006.

You might also like