RMP v065 p1331

Download as pdf or txt
Download as pdf or txt
You are on page 1of 62

The analysis of observed chaotic data in physical systems

Henry D. I. Abarbanel
Department of Physics and Marine Physical Laboratory, Scripps institution of Oceanography
University of California at San Diego, La Jolla, California 92093-0402
Reggie Brown, John J. Sidorowich, and Lev Sh. Tsimring
Institute for Nonlinear Science, University of California at San Diego, Mail Code 0402,
La Jolla, California 92093-0402
Chaotic time series data are observed routinely in experiments on physical systems and in observations in
the field. The authors review developments in the extraction of information of physical importance from
such measurements. They discuss methods for (1) separating the signal of physical interest from contamination ("noise reduction"), (2) constructing an appropriate state space or phase space for the data in
which the full structure of the strange attractor associated with the chaotic observations is unfolded, (3)
evaluating invariant properties of the dynamics such as dimensions, Lyapunov exponents, and topological
characteristics, and (4) model making, local and global, for prediction and other goals. They briefly touch
on the effects of linearly filtering data before analyzing it as a chaotic time series. Controlling chaotic
physical systems and using them to synchronize and possibly communicate between source and receiver is
considered. Finally, chaos in space-time systems, that is, the dynamics of fields, is briefly considered.
While much is now known about the analysis of observed temporal chaos, spatio-temporal chaotic systems
pose new challenges. The emphasis throughout the review is on the tools one now has for the realistic
study of measured data in laboratory and field settings. It is the goal of this review to bring these tools
into general use among physicists who study classical and semiclassical systems. Much of the progress in
studying chaotic systems has rested on computational tools with some underlying rigorous mathematics.
Heuristic and intuitive analysis tools guided by this mathematics and realizable on existing computers
constitute the core of this review.

CONTENTS
I. Introduction
A. Observed chaos
B. Outline of the review
II. Signals, Dynamical Systems, and Chaos
III. Analyzing Measured SignalsLinear and Nonlinear
A. Signal separation
B. Finding the space
C. Classification and identification
D. Modelinglinear and nonlinear
E. Signal synthesis
IV. Reconstructing Phase Space or State Space
A. Choosing time delays
B. Average mutual information
C. Choosing the embedding dimension
1. Singular-value analysis
2. Saturation of system invariants
3. False nearest neighbors
4. True vector
fields
D. TanddE
V. Invariants of the Dynamics
A. Density estimation
B. Dimensions
1. Generalized dimensions
2. Numerical estimation
C. A seguefrom geometry to dynamics
D. Lyapunov spectrum
E. Global exponents
F. Ideal case: Known dynamics
G. Real world: Observations
1. Analytic approach
2. Trajectory tracing method
3. Spurious exponents
H. Local exponents from known dynamics and from
observations
Reviews of Modern Physics, Vol. 65, No. 4, October 1993

1331
1333
1335
1335
1339
1339
1340
1341
1342
1342
1343
1344
1344
1346
1347
1347
1348
1351
1351
1352
1352
1353
1354
1355
1356
1358
1358
1359
1359
1360
1361
1362
1362

I. Topological invariants
VI. Nonlinear Model Building: Prediction in Chaos
A. The dimension of models
B. Local modeling
C. Global modeling
D. In between local and global modeling
VII. Signal Separation"Noise" Reduction
A. Knowing the dynamics: manifold decomposition
B. Knowing a signal: probabilistic cleaning
C. Knowing very little
VIII. Linearly Filtered Signals
IX. Control and Synchronization of Chaotic Systems
A. Controlling chaos
1. Using nonlinear dynamics
2. More traditional methods
3. Targeting
B. Synchronization and chaotic driving
X. Spatio-temporal Chaos
XI. Conclusions
A. Summary
B. Cloudy crystal ball gazing
Acknowledgments
References
Additional References

1365
1366
1367
1368
1370
1370
1372
1373
1375
1376
1377
1378
1378
1378
1380
1381
1383
1384
1386
1386
1387
1388
1388
1340

I. INTRODUCTION
Interpretations of measurements on physical systems
often turn on the tools one has to make those interpretations. In this review article we shall describe new tools
for the interpretation of observations of physical systems
where the time trace of the measured quantities is irregular or chaotic. These time series have been considered
bothersome "noise" without the tools we shall describe

0034-6861/93/65(4)/1331(62)/$11.20

1993 The American Physical Society

1331

1332

Abarbanel et al.: Analysis of observed chaotic data

and, as such, have traditionally been taken as contamination to be removed from the interesting physical signals
one sought to observe. The characterization of irregular,
broadband signals, generic in nonlinear dynamical systems, and the extraction of physically interesting and useful information from such signals is the set of achievements we review here. The chaotic signals that we shall
address have been discarded in the past as "noise," and
among our goals here is to emphasize the valuable physics that lies in these signals. Further we want to indicate
that one can go about this characterization by exploiting
the structure in these chaotic time series in a systematic,
nearly algorithmic, fashion. In a sense we shall describe
new methods for the analysis of time series, but on another level we shall be providing handles for the investigation and exploitation of aspects of physical processes that
could be simply dismissed as "stochastic" or random
when seen with different tools. Indeed, the view we take
in this review is that chaos is not an aspect of physical
systems that is to be located and discarded, but is an attribute of physical behavior that is quite common and
whose utilization for science and technology is just beginning. The tools we discuss here are likely also to be just
the beginning of what we can hope to bring to bear in the
understanding and use of this remarkable feature of physical dynamics.
The appearance of irregular behavior is restricted to
nonlinear systems. In an autonomous linear system of /
degrees of freedom, u(t) = [u {(t),u2(t), . . . ,Uf(t)], one
has
- = A-u(r) ,
(1)
at
where A is a constant fXf
matrix. The solution of this
linear equation results in (1) directions in /-space along
which the orbits uU) shrink to zeronamely, directions
along which the real part of the eigenvalues of A are
negative, (2) directions along which the orbits unstably
grow to infinitynamely, directions along which the real
part of the eigenvalues of A are positive, and (3) directions where the eigenvalues occur in complex-conjugate
pairs along with zero or negative real part. In any situation where the eigenvalue has a positive real part, this is
a message that the linear dynamics is incorrect, and one
must return to the nonlinear evolution equations that
govern the process to make a better approximation to the
dynamics.
Much of the work in nonlinear dynamics reported in
the past has concentrated on learning how to classify the
nonlinear systems by analyzing the output from known
systems. These efforts have provided, and continue to
provide, significant insights into the kinds of behavior
one might expect from nonlinear dynamical systems as
realized in practice and has led to an ability to evaluate
now familiar quantities such as fractal dimensions,
Lyapunov exponents, and other invariants of the nonlinear system. The capability for doing these calculations
is strongest when we know the differential equations or

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

mappings of the dynamical system. When inferring such


quantities from data alone, we find the issues more challenging, and this is the subject matter covered in this article.
At the outset it is appropriate to say that this is not,
and actually may never be, an entirely settled issue.
However, a certain insight and the establishment of certain general guidelines has been achieved, and that in itself is important. Only recently has the question of inferring from measured chaotic time series the properties of
the system that produced those time series been addressed with any real vigor. The reason is partly that,
until the classification tools were clearly established, it
was probably too soon to tackle this harder problem.
Moreover, by its nature, this challenge of "inverting"
time series to deduce characteristics of the physical
systemthat
is, model building for
nonlinear
dynamicshas required the development of tools
beyond those of direct classification, and these too have
taken time to evolve. While ^some general results are
known and will be described here, much of the analysis of
a physical system may be specific to that system. Thus
the tools here should be regarded as a guide to the
reader, not a strict set of rules to be followed.
This article is designed to bring to scientists and engineers a familiarity with developments in the area of
model building based on signals from nonlinear systems.
The key fact that makes this pursuit qualitatively
different from conventional time series analysis is that,
because of the nonlinearity of the systems involved, the
familiar and critical tool of Fourier analysis does very little, if any, good in the subject. The Fourier transform of
a linear system changes in what might be a tedious set of
differential equations into an algebraic problem where decades of matrix analysis can be fruitfully brought to bear.
Fourier analysis of nonlinear systems turns differential
equations in time into integral equations in frequency
space involving convolutions among the Fourier transforms of the dependent variables. This is rarely an improvement, so Fourier methods are to be discounted at
the outset, though as an initial window through which to
view the data, they may prove useful.
Having set aside at the outset the main analysis tool
for linear signal processing, we must more or less reinvent signal processing for nonlinear systems. In rethinking the problem it is very useful to address the essential
set of general issues faced by signal processors. We envision that the physicist will be presented with observations of one or perhaps, if fortunate, a few variables from
the system and then be interested in deducing as much as
possible about the characteristics of the system producing the time seriesthis is system classification.
The
next step would be creating a framework in which to predict, within limits determined from the data itself, the future behavior of the system or the evolution of new
points in the system state or phase spacethis is model
building and parameter estimation, since the models that
allow prediction are usually parametric.

Abarbanel et at.: Analysis of observed chaotic data

1333

Even before beginning the analysis, the practitioner


has to address the question of how to deal with data contaminated by other signals or by the measurement process itself. This contamination can come from environmental variables unmeasured by the instruments or from
properties of the observation instruments. In a general
sense the contamination has broadband Fourier spectra
with possible sharp lines embedded in the continuum.
This general description will also be true of the Fourier
spectrum of the chaotic signal, since it arises from nonperiodic motion and has a continuous broadband spectrum with possible sharp spectral lines in it. Separating
these two spectrally similar phenomena is not a task for
which standard time series analysis is well suited, but one
which, using differing characteristics of the signal and
contamination, we may address in the time domain. We
shall indicate in sections of this article how one goes
about this initial task in signal processing in chaos.
Broadly speaking, this article will follow the directions
pointed out in Table I. This indicates the requirements
on the physicist for creating some order from the time
series presented, and it is clear that the same problems,
with different solutions, are faced whether the system is
linear or nonlinear.
In an ambitious fashion, once one can clean up dirty
signals, classify the systems that give rise to those signals,
and reliably predict the behavior of the system in the future, within limits set by the dynamics themselves, then
consideration of control strategies for nonlinear systems
is in order. This topic will be touched on in later sections
of this review.
One important insight into dynamical systems is the
role played by information theory. There is an intuitive
notion that a dynamical system that has chaotic behavior
is precisely a realization of Shannon's concept of an ergodic information source (Gallager, 1968; Shaw, 1981,
1984). This source is assumed to produce a sequence of
symbols Xt\. . . ,X_l,X0,Xl,
. . . drawn from an alphabet x 0,1, . . . , J 1. This is for a discrete source with
a finite number of states available to it, namely, any
source we would realize on a finite-state digital computer
or in an experiment. The distribution of orbit points in
the alphabet P(x) gives the probability for realizing the
symbol x by measurements of the Xt. The entropy (in
bits) of a joint set of measurements is denned using the
joint probability P(xlfx2,.
. . ,xN) via

Clearly the source is no more than an abstract statement of the system producing our measurements, and the
concept of entropy of a deterministic dynamical system,
where no notion of probability is immediately evident,
seems much the same as in the information-theoretic setting. This idea was realized over thirty years ago by Kolmogorov (1958), who gave a definition of this entropy
which, in practice, is precisely the same as h (X) above.
Further, this entropy is an invariant of the orbits of the
system, that is, independent of the initial conditions or
the specific orbits observed. Additional work on this
quantity was done by Sinai (1959), and the quantity h (X)
has become known as the Kolmogorov-Sinai (or KS) entropy of a dynamical system. Strengthening this connection is a theorem of Pesin (1977) equating the K S entropy
to the sum of the positive Lyapunov exponents. If a
dynamical system has no positive Lyapunov exponents, it
does not possess chaos in the usual sense.
These comments indicate the connection between
dynamical systems and information theory, but do not
explain the interest for someone concerned with the
analysis and modeling of chaotic data. The KS entropy
gives a precise statement of the limits to predictability for
a nonlinear system. Think of a phase space for the system which is reconstructed or, if one actually knows the
differential equations or mapping, precise. Any realizable statement about the system is known only to some
precision in this state space. Within that precision or
resolution we cannot distinguish at a given time t = 0 between two state-space points lying in the resolution cell.
A nonlinear system with positive h (X) is intrinsically unstable, however, and the two points unresolvable at t = 0
will move after some time T to differing parts of the
phase space. 2^-h{X)T^ is a measure of the number of states
occupied by the system after the passage of a time T
(Gallager, 1968; Rabinovich, 1978). When this number
of states is approximately the total number of states
available for the orbits of the system, then all prediction
of the future of an orbit becomes untenable. This occurs
in approximately T^l/h
(X). One still has statistical information about the system that is quite useful and interesting, but knowledge of the evolution of a specific orbit is lost. Thus, h (X) becomes an object of considerable
interest for the physicist. Since it is the sum of the positive Lyapunov exponents, one can use the time series itself to determine how predictable it is.

HN{X)=-^P(xux2,

A. Observed chaos

. . . ,xN)log2[P(xl,x2,.

. . ,xN)]

(2)
As N becomes large this entropy, divided by N, has a
finite limit,
lim
N-+oo

HN(X)
"

=h(X),

(3)

and this is the amount one learns about the source of the
symbols by the sequence of measurements on the set X.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Actual observations of irregular evolution that we consider are typically measurements of a single scalar observable at a fixed spatial point: call the measured quantity
s(t0 + nrs)=s{n),
where tQ is some initial time and rs is
the sampling time of the instrument used in the experiment. sin) could be a voltage or current in a nonlinear
circuit or a density or velocity in a fluid dynamic or plasma dynamic experiment or could be the temperature at
some location in a laboratory container or an atmospher-

1334

Abarbanel et a/.: Analysis of observed chaotic data

ic or oceanic field experiment. The full structure of some


of these measurements lies in the observation of a field of
quantities, namely, s(x,t0-\-nrs)
with x = (x,y,z) the spatial coordinates. Most of this review will focus on
methods for the analysis of scalar measurements at a
fixed spatial point. This is done for the simple reason
that the subject has matured to a point where one can
productively review what is known with the hope that
the methods can then become part of every physicist's experimental analysis toolkit in much the same way that
Fourier analysis now is part of that toolbox. At the end
of our discussion we shall have some things to say about
spatio-temporal systems, but the developments there
which parallel what we know about purely temporal
measurements remain to be completed.
The first task will be to explain how to begin with this
set of scalar measurements and to reconstruct the system
phase space from it. This requires some explanation. In
fluid dynamics, for example, the system phase space is
infinite dimensional when the description of the fluid by
the Navier-Stokes partial differential equations is correct.
We do not reestablish from the s (n) an infinite number of
degrees of freedom, but we take the point of view that,
since the observations of a dissipative system such as a
fluid lie on a geometric object of much smaller dimension
than that of the original state space, we must seek
methods for modeling the evolution on the attractor itself
and not evolution in the full, infinite-dimensional, original phase space. In some circumstances this dimension
may be quite small, as low as five or so. The techniques
we shall discuss are then directed toward the description
of the effective time-asymptotic behavior of the observed
dynamics. In rare circumstances this might be expected
to yield the original differential equations.
The measurements s(n) and the finite number of quantities formed from them in the course of the analysis provide coordinates for the effective finite-dimensional space
in which the system is acting after transients have died
out. This gives rise to two approaches to modeling the
dynamics from the point of view of the physicist interested in understanding the system producing the measured
signals:

the Lyapunov exponents is zero, as determined by the


data using methods that we shall describe later, then one
could construct differential equations to capture the dynamics. This suggests we are confident that extrapolation to r 5 - > 0 , implicit in the idea of a differential equation rather than a finite time map, captures all the
relevant frequencies or time scales or degrees of freedom
of the source. In either casemap or differential
equationno rigid rules appear to be available to guide
the modeler.
The second approach is more traditional and consists
of making models of the experiment in the usual variables of velocity, pressure, temperature, voltage, etc. The
job of the physicist is then to explore the implications of
the model for various invariant quantities under the dynamics, such as the dimensions of the attractor and the
Lyapunov exponents of the system. These invariants,
which we shall discuss below, are then the testing
grounds of the models. The critical feature of the invariants is that they are not sensitive to initial conditions or
small perturbations of an orbit, while individual orbits of
the system are exponentially sensitive to such perturbations. The exponential sensitivity is the manifestation of
the instability of all orbits in the phase space of the dynamics. Because of this intrinsic instability no individual
orbit can be compared with experiment or with a computed orbit, since any orbit is effectively uncorrelated
with any other orbit, and numerical roundoff or experimental precision will make every orbit distinct. What
remains, and what we shall emphasize in this review, is a
statistics of the nonlinear deterministic system which produces the orbits. An additional set of invariants, topological invariants, are also quite useful with regard to this
task of characterizing physical systems and identifying
which models of the system are appropriate (Mindlin
et ah, 1991; Papoff et ah, 1992), and we shall discuss
them as well. Topological invariants are distinct from
the better known metric invariants mentioned above in
that they do not rest on the determination of distance or
metric properties of orbit points on an attractor. They
appear to be quite robust in the presence of "noise" and
under changes in system parameters.

The first approach is that of making physical models


in the coordinates s(n) and quantities made from the
s(n)and
doing all this in the finite-dimensional space
where the observed orbits evolve. This description of the
system differs markedly from the original differential
equations, partial or ordinary, which one might have
written down from first principles or from arguments
about the physics of the experiment. Verification of the
model by future experiment or different features of the
experiment proceeds as usual, but the framework of the
models is quite different from the usual variations of
Newton's laws, which constitute most physics model
making. The net result of this will be a nonlinear relationship or map from values of the observed variables
s (n) to their values s (n -\~T) some time T later. If one of

The statistical aspect of chaotic dynamics was discussed and its importance stressed in an earlier article in
this journal by Eckmann and Ruelle (1985). Our review
now is in many ways an update of features of that earlier
article with an emphasis on practical aspects of the
analysis of experimental data and model building for prediction and control. Of course, we also include many developments unanticipated by the earlier review, but we
certainly recommend it as an introduction to many of the
topics here, often from a rather more mathematical
viewpoint than we project.
Many items we shall mention here, especially in the
sections on signal separation, control of chaotic systems,
and synchronization of chaotic systems, have direct practical implications which move beyond the usual preoccu-

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Abarbanel et al.: Analysis of observed chaotic data


pation of physicists with model building, prediction, and
model verification. The authors expect that many of
these topics will find further development in the engineering literature, and we cannot anticipate the full
spectrum of those developments. We mention some
throughout the review, since we think it quite important
to underline the utility of chaotic states of a physical system in the hopes that physicists will seek new phenomena
suggested by chaos and uncover further interesting physical behavior connected with this disordered, but not random, feature of deterministic systems.
Just a note before we move ahead: we interchangeably
denote the space in which orbits of dynamical systems
evolve as state space or phase space. State space is more
commonly used in the engineering literature, while phase
space is more common in the physics literature. In physics, phase space is also used to denote the canonical set of
coordinates of Lagrange or Hamilton for conservative
systems. We do not use phase space in this restricted
sense, nor does our use of "phase" refer to the argument
of a complex variable.
B. Outline of the review
The main body of the review begins in the next section
with some details of approaches to the analysis of time
series, first from the traditional or linear point of view.
Then we move on to the idea of reconstructing the phase
space of the system by the use of time delays of observed
data, s(n). In the subsequent section we discuss methods
of determining the time delay to use in the practical
reconstruction of phase space, and in doing so introduce
the ideas of information theory in greater detail. Having
established the time delay to use, we move on to a discussion of the choice, from analysis of the observed data, of
the dimension of the phase space in which we must work
to unfold the attractor on which the dynamics evolves.
With that we end our discussion of the state space in
which we are to view the dynamics and we move on to
the issue of classifying the dynamics that underlies the
observations. The classifiers in a linear system are the
resonant Fourier frequencies of the system. In a nonlinear system we have statistical aspects of the deterministic chaotic evolution which we can use for this same
purpose. The quantities invariant under the dynamics
are used as classifiers, and we shall discuss the physical
and mathematical meaning of some of the most widely
utilized of these invariants.
After we have established ways to classify the physical
system leading to the observations s (n), we move on to a
discussion of model building to allow the codification of
the measurements into a simple evolution rule. This is an
easy task, as we shall see, if all we desire is numerically
accurate predictors. If we wish insight into the physical
processes needed to produce, reproduce, and predict the
signals we observe, then the task is harder, and no clearcut method has emerged, or for that matter may ever
emerge. Our discussion will cover various ideas on this
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1335

problem, all of which have limitations or unattractive


features, however successful they may be in a numerical
sense.
The next topic we address is that of separating the
desired "clean" chaotic signal from the effects of contamination by the environment through which it passes on its
way from the source to the measurement device or contamination by properties of the measurement process itself. This contamination is often called "noise," and the
task of signal separation goes under the label of "noise
reduction" in common parlance. The ideas involved in
signal separation carry one far beyond traditional views
of noise reduction, and we shall explore some of these as
well. In any case, the methods one must employ are to be
used in the time domain of reconstructed phase space,
and they differ in detail and conception from Fourierbased methods, which are much more familiar. The
penultimate topic we discuss is that of controlling chaotic systems by utilizing properties of the structure of the
attractor to move the system from its "natural" autonomous chaotic state to motion about an unstable periodic motionunstable from the point of view of the autonomous system, but stable from the point of view of the
original system plus a control force.
Finally we turn to aspects of the experimental analysis
of spatio-temporal systems with disorder or chaotic
behavior. In this area, as noted earlier, we know much
less, so much of what we present is speculative. Clearly a
review of this subject five or so years from now will,
without question, supplant this aspect of the review.
Nonetheless, we felt it important to end the review with a
glimpse, however imperfect, of the problems that will
dominate the future of this kind of work. In any case,
the analysis of chaos of fields is essential for doing real
physics in chaotic systems, so the problems we outline
and the potential solutions we discuss certainly will affect
our ability to work with chaotic fields.

II. SIGNALS, DYNAMICAL SYSTEMS, AND CHAOS


In the analysis of signals from physical systems we assume from the outset that a dynamical system in the
form of a differential equation or a discrete-time evolution rule is responsible for the observations. For
continuous-time dynamics we imagine that there is a set
of / ordinary differential equations for variables
u(t) = [ux(t),U2(t),

...

9Uf(t)]9

^ = G ( u U ) ) ,
(4)
dt
where the vector field G ( u ) is always taken to be continuous in its variables and also taken to be differentiate as
often as needed. When time is discrete, which is the realistic situation when observations are only sampled every
rs, the evolution is given by a map from vectors in R-f to
other vectors in R^, each labeled by discrete time:
u(n) = u(t0-\-nTs),

Abarbanel et al.: Analysis of observed chaotic data

1336

(5)

u(n+l)=F(u(/i))

The continuous-time and discrete-time views of the dynamics can be connected by thinking of the time derivative as approximated by
du(t)
dt

uUp + U +l)Ts)-u(t0
^

+ nrs)

(6)

Ts

(7)

as the relation between the continuous and discrete dynamics.


Discrete dynamics can also arise by sampling the continuous dynamics as it passes through a hyperplane in the
/-dimensional space. This method of Poincare sections
leads to an ( / 1 )-dimensional discrete system that is
not easily related in an analytic fashion to the continuous
system behind it. Often qualitative properties of the continuous system are manifest in the Poincare section: an
orbit that is a recurrent point (fixed point) in the section
is a periodic orbit in the continuous flow.
The number of degrees of freedom or, equivalently, the
dimension of the state space of a dynamical system is the
same as the number of first-order differential equations
required to describe the evolution. Partial differential
equations generalize the number of degrees of freedom
from a finite number, / , to a continuum labeled by a vector x = ( x 1 , x 2 , . ,xD). For a field with K components
u(x,t) = [ul(xyt),u2(x,t),
. . . ,u K {x,t)] 9 we write the
differential equation
9u(x,f)
dt

= G(u(x,f))

(8)

and the labels of the dynamical variable now are both


continuous (namely, x and t) and discrete. Broadly
speaking the dynamics of fields differs from that of lowdimensional dynamical systems only because the number
of degrees of freedom has become continuous. Many new
phenomena can appear because of this, of course, but
when we think of the analysis of real data it will certainly
be finitely sampled in space x as well as in time, and we
may consider the partial differential equation to have
been reduced to a very large number of ordinary
differential equations for purposes of discussion.
There is one important distinction between discrete dynamics that comes from a surface of a section of a flow
(continuous-time dynamics) and that from finite time
sampling of the flow. In the latter case we can associate
the direction of the vector field with evolution which can
be explored by the system. In the former case evolution
along the direction of the vector field is off the surface of
the section and is not available for study. This difference
will be manifest when we discuss Lyapunov exponents
below.
The vector field F ( u ) has parameters that reflect the
external settings of forces, frequencies, boundary conditions, and physical properties of the system. In the case
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

(9)

u ( / i + l ) = F(u(w))

which would lead to


F(u())u(yi) + rsG(u(/i)) ,

of a nonlinear circuit, for example, these parameters


would include the amplitude and frequency of any driving voltage or current, the resistance or capacitance or
inductance of any lumped linear element, and the IV or
other characteristics of any nonlinear elements.
The dynamical system

is generally not volume preserving in /-dimensional


space. In evolving from a volume of points d^u ( n ) t o a
volume of points d^u(n + 1 ) the volume changes by the
determinant of the Jacobian
DF(u) = det

3F(u)
9u

(10)

When this Jacobian is unity, which is a very special occurrence and is usually associated with Hamiltonian evolution in a physical setting, phase-space volume remains
constant under the evolution of the system. Generally
the Jacobian has a determinant less than unity, which
means volumes in phase space shrink as the system
evolves. The physical origin of this phase-space volume
contraction is dissipation in the dynamics.
As the physical parameters in the vector field are
varied, the behavior of u(t) or u(), considered as an
initial-value problem starting from u U 0 ) or u(0), will
change in detail as t or n becomes large. Some of these
changes are smooth and preserve the geometry of the set
of points in R f visited by the dynamics. Some changes
reflect a sudden alteration in the qualitative behavior of
the system. To have an explicit example to discuss, we
consider the three differential equations of Lorenz (1963)1
dx(t)
dt

= a(y(t)-x(t))

dy(t)

= x{t)z(t)Jrrx{t)y{t)

,
,

(11)

dt
dz(t)
dt

=x(t)y(t)

bz(t) .

These equations are derived from a finite mode truncation of the partial differential equations describing the
thermal driving of convection in the lower atmosphere.
The ratio of thermal to viscous dissipation is called a , the
ratio of the Rayleigh number to that critical Rayleigh
number where convection begins is called r, and the dimensionless scale of a convective cell is called b. The parameters a and b are properties of the system and the
geometry-planar herewhile the parameter r is deter-

^ n a later paper (Lorenz, 1984) there is a delightful description of how he came to consider this model. He was concerned
with linear prediction of nonlinear models and used this threedegree-of-freedom model as a testing ground for this study. He
concluded that the idea of linear prediction for this kind of system would not do.

Abarbanel et al.: Analysis of observed chaotic data


mined by geometry, material properties, and the strength
of the driving forces. The physically interesting behavior
of the Lorenz system comes as r is varied, since that
represents variations in the external forces on the "atmosphere." The derivation of the three ordinary differential
equations shows that they provide a representation of
thermal convection only near r l (Salzmann, 1962),2
but we, as have many others, have adopted them as a
model of low-dimensional chaotic behavior, and we shall
push the parameters completely out of the appropriate
physical regime.
When r is less than unity, all orbits tend to the fixed
point of the vector field,
Gl(x>y,z) = ax + ay ,
G2(x,y,z)=

y +rx

G3(x,y,z)=bz

+xy

xz

(12)

at (x,j;,z) = (0,0,0). This represents steady thermal conduction. As r increases beyond unity, the state
(0,0,0) is linearly unstable, and the vector field has two
symmetric
linearly
stable
fixed
points
at
(x,y,z)
= {Vb(r-l),Vb(r-\),r-l).
For all
r < 1, the geometry of the time-asymptotic state of the
system is the same, namely, all initial conditions
tend to (0,0,0). For r > 1 and until r reaches
rca{a + b -\-3)/(a b 1), orbits end up at
(x,y,z)
depending on the value of (x(0),y(0),z(0)).
The
geometry of the final state of the orbits is the same,
namely, all volumes of initial conditions shrink to zerodimensional objects. These zero-dimensional limit sets
change with parameters, but retain their geometrical nature until rrc. When r>rc, the situation changes
dramatically. The two fixed points of the vector field at
(x,yyz)
become linearly unstable, and no stable fixed
points remain. As r is increased from rc, the timeasymptotic state emerging from an initial condition will
either perform irregular motion characterized by a broad
Fourier spectrum or undergo periodic motion.
The analysis of the sequence of allowed states in a
differential equation or a map as parameters of the system are varied is called bifurcation theory. We shall not
dwell on this topic as it is covered in numerous monographs and literature articles (Guckenheimer and
Holmes, 1983; McKay, 1992; Crawford, 1991; Crawford
and Knobloch, 1991). What is important for us here is
that as we change the ratio of forcing to damping,
represented by the parameter r in the Lorenz equations,
the physical system undergoes a sequence of transitions
from regular to more complex temporal evolution. The
physical picture underlying these transitions is the requirement of passing energy through the system from
whatever forcing is present to the dissipation mechanism
for the particular system. As the amount of energy flux
2

The same equations were derived in the context of laser physics by Orayevsky et al. (1965).
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1337

increases, the system chooses different geometrical


configurations in phase space to improve its ability to
transport this energy. These configurations in the early
stages frequently involve the appearance of higher
Fourier modes than would be dictated by the symmetry
of the driving forces and the geometry of the container.
In a fluid heated from below the transition from conduction of heat to convection is an example of this (Chandrasekhar, 1961). The sequence of bifurcations that
would be suggested by these first instabilities and changes
in geometry does not persist. After a few new frequencies have been induced in the system by increasing the
forcing to dissipation ratio, the system generically undergoes a structural transition to "strange" behavior in
which the entire phase space is filled with unstable orbits.
These unstable orbits do not go off to infinity in the phase
space, as a rule, since in addition to the instabilities one
has dissipation which limits the excursions of physical
variables. The resulting stretching by instabilities and
folding by dissipative forces is at the heart of the
"strange" behavior we call chaos. When the forcing-todissipation ratio is small and regular behavior is observed, the time-asymptotic orbits evolve from whatever
is the initial condition to motion on an attracting set of
points that is regular and is composed of simple
geometric figures such as points or circles or tori. The
dimension of these geometric figures is integer and, because of the dissipation, smaller than the dimension / of
the underlying dynamics. When the system becomes unstable, the orbits are attracted from a wide basin of initial
conditions to a limiting set, which is fractional dimensional and has an infinite number of points. The complex
appearance of the time series in such a situation is due to
the nonperiodic way in which the orbit visits regions of
phase space while moving along the geometric object we
now call a strange attractor.
In Fig. 1 we display the time series for x(t) which results from solving the Lorenz equations using a fourthorder Runge-Kutta integrator for time step
dT=0.0\.
Lorenz 63
Time series, 8 000 points

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

FIG. 1. Chaotic time series x{t) produced by Lorenz (1963)


equations (11) with parameter values r =45.92, 6 = 4 . 0 ,
a = 16.0.

1338

Abarbanel et al.: Analysis of observed chaotic data

The parameter values were r = 4 5 . 9 2 , 6 = 4 . 0 , and


cr = 16.0. In Fig. 2 we plot the orbit coming from the
three degrees of freedom [x(t)yy(t),z(t)]
in a threedimensional space. The familiar and characteristic structure of the long-term orbit of the Lorenz systemthe
Lorenz attractoris apparent.
Much of the work on dynamical systems over the past
fifteen years has focused on the analysis of specific physically, biologically, or otherwise interesting differential
equations or maps for purposes of illuminating the various kinds of behavior allowed by nonlinear evolution
equations. The sets of points in R f visited by the orbits
of such systems is probably not yet completely classified
from a mathematical point of view, but the broad and
useful description suggested above, as the ratio of forcing
to dissipation is varied, has emerged. The details of the
bifurcations involved in the transition from regular
behavior, where the limit sets are points or periodic orbits, to irregular behavior such as that exhibited by the
Lorenz model are often complicated. They are of interest
in any specific physical setting, of course, but we shall
deal with situations in which the system has arrived in a
state where the parameter settings lead to irregular
behavior characterized by a broad Fourier power spectrum, perhaps with some sharp lines in it as well.
This broad power spectrum is a first indication of
chaotic behavior, though it alone does not characterize
chaos. For the Lorenz model in the parameter regime indicated, the Fourier power spectrum is shown in Fig. 3.
It is broad and continuous and falls as a power of frequency. Throughout this review we shall also work with
data from a nonlinear circuit built by Carroll and Pecora
at the U.S. Naval Research Laboratory (Carroll and
Pecora, 1991) as an example for many of our algorithms
and arguments. The time series for this circuit is taken

Lorenz 63: Power spectrum


of x(t) time series, 30 000 points

10 [

io-2

10"4

L
10

___

..
10

. , . . . Ill
10

FIG. 3. Power spectrum of the Lorenz data x(t).


as a tapped voltage from a location indicated in their
original paper with rs=0A
ms. The voltage as a function of time is shown in Fig. 4 and the corresponding
Fourier spectrum in Fig. 5. Anticipating things to come,
we plot in Fig. 6 a reconstruction of the phase space or
state space of the hysteretic circuit orbits. The coordinates are the measured voltage, the same voltage 0.6 ms
later and the same voltage 1.2 ms later. In contrast to
the irregular voltage time trace in Fig. 4 and the Fourier
spectrum of this voltage in Fig. 5, the phase-space plot
demonstrates simplicity and regularity. These data indicate that the circuit is a candidate for a chaotic system,
and we shall systematically investigate other features of
the system from this voltage measurement. It is important to note that the time series alone does not determine
whether the signal is chaotic. A superposition of several
incommensurate frequencies will result in a complicated
time trace but a simple line spectrum in Fourier space.
Linear analysis is not well tuned to working with signals with continuous, broad power spectra. Since stable
linear dynamics produces either zero frequency, that is,
Hysteretic circuit
4.0 ,

100

Time series, 5 000 points


,
r
,

200

300

400

500

t, ms

FIG. 2. Lorenz attractor in three-dimensional phase space


(x{t),y{t),z{t)).
Rev. M o d . Phys., V o l . 65, No. 4, October 1993

FIG. 4. Time series of voltage V{t) produced by a hysteretic


circuit (Carroll and Pecora, 1991).

Abarbanel et a/.: Analysis of observed chaotic data


Hysteretic circuit: Power spectrum
of V(t) time series, 30 000 points
104

< -

- r r -

1339

sions from global linear modeling can significantly improve performance in modeling nonlinear signals.
III. ANALYZING MEASURED SIGNALSLINEAR
AND NONLINEAR

io 2 I

10 [

The tasks faced by the analyst of signals observed from


linear or nonlinear systems are really much the same.
The methods for the analysis are substantially different.
This section is devoted to a discussion of the similarities
and differences in these requirements. The discussion is
centered around Table I.

A. Signal separation
FIG. 5. Power spectrum of the time series V(t) from the hysteretic circuit.
fixed points, or oscillations at a few frequencies, it is not
surprising that the use of linear models is not likely to
give much insight into the behavior of signals
arising from nonlinear systems. Traditional autoregressive/moving-average models (Oppenheim and
Schafer, 1989), then, are not going to do very well in the
description of chaotic signals. Indeed, if one tries to use
linear models to fit data from chaotic sources and
chooses to fit the unknown parameters in the model by a
least-squares fit of the model to the observations, the rms
error in the fit is about the same size as the attractor; that
is, nothing has been accomplished. If one tries a local
linear fit connecting neighborhoods in the state space of
the system (this is certainly one of the simplest nonlinear
models), the rms error can essentially be made as small as
one likes depending on the quantity of data and size of
the neighborhoods. This is not to suggest that local
linear models answer all questionsindeed, we shall see
below this is not the casebut that even small excur-

FIG. 6. Hysteretic time series embedded in a three-dimensional


phase space: [V(t)9 V(t + 6), V(t + 12)]. See Fig. 4.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

The very first job of the physicist is to find the signal.


This means identifying the signal of interest in an observation that is possibly contaminated by the environment,
by perturbations on the source of the signal, or by properties of the measuring instrument. In a graphic way we
can describe the process of observing as (1) production of
the signal by the source, potentially disturbed during the
transmission, (2) propagation of the signal through some
communications channel, potentially contaminated by
the environment, and then (3) measurement of the signal
at the receiver, potentially disturbed during the reception
and potentially acting as a filter on the properties of the
transmitted signal of interest.
The problems in finding the signal are common to the
linear and the nonlinear observer. The linear observer
has the advantage of being able to assume that the source
emits spectrally sharp monochromatic signals which
might be contaminated by broadband interference. The
separation of the signal of interest from the unwanted
background becomes an exercise in distinguishing narrowband signals in a broadband environment. Methods
for this (Van Trees, 1968) are easily fifty years old and
quite well developed.
Indeed, the main issue is really signal separation, and
in order to carry out the processing required we must establish some difference between the information-bearing
signal and the interference. In the case of a narrowband
signal in a broadband environment, the difference is clear
and Fourier space is the right domain in which to perform the separation. Similarly if the signal and the contamination are located in quite different bands of the
spectrum, the Fourier techniques are certainly indicated.
In the case of signals from chaotic sources, both the signal of interest and the background are typically broadband, and Fourier analysis will not be of much assistance
in making the separation. In the first entry of the nonlinear side of Table I we identify several methods for
separating the signal from its contamination. All of them
rest on the idea that the signal of interest is low- or fewdimensional chaos with specific geometric structure in its
state space. The geometric structure in the phase space
is characteristic of each chaotic source, so we can use

Abarbanel et al.: Analysis of observed chaotic data

1340

this to effect the separation of that signal from others.


When the contamination is "noise," which we shall see as
an example of high-dimensional chaos, the signal separation problem is often called "noise reduction," but it
remains at heart a separation problem. In nonlinear signal separation, three cases seem easy to distinguish.
Each of these will be discussed in some detail in Sec. VII:

we are "flying blind" and must make models of the signal


which allow us to establish what part of our observations
are deterministic in low-dimensional phase space and
what part a result of processes in higher-dimensional
state space and thus to be discarded. It should be clear
that this is much more problematic than the two other
scenarios, but it may often be the realistic situation.

We know the dynamics.


We know the evolution
equations u( + l ) = F ( u ( n ) ) and can use this knowledge
to extract the signal satisfying the dynamics from the
contaminating signals.
We know a signal from the system. We have measured some signal from the system of interest at some
earlier time. We can use this earlier measurement as a
reference signal to establish the statistics of the evolution
on the attractor and use these statistics to separate a subsequent transmission of the chaotic signal from contamination of the kind indicated just above.
We know nothing. We know only a single instance of
measuring a signal from the physical source. In this case

B. Finding the space


If we now suppose we have found the signal of interest,
we may move on to establishing the correct space in
which to view the signal. If the source of our signal is a
linear physical process, perhaps a drum being beaten in a
linear range or a seismic measurement far from the epicenter, so the displacements are very small, then the natural space in which to view the signal is Fourier space. If
the source is invariant under translations in time, then
sines and cosines are the natural basis functions in which
to expand the problem, and the display of the information in the measurement will be most effective in Fourier

TABLE I. Connection between the requirements in linear and nonlinear signal processing. The goals are much the same in each
case; the techniques are quite different.
Linear signal processing

Nonlinear signal processing


Finding the signal
Noise reduction; detection
Separate broadband signal from broadband
noise using deterministic nature of signal. If
system is known or observed, make
"matched filter" in time domain. Use dynamics or invariant distribution and Markov
transition probabilities.

Noise reduction; detection


Separate broadband noise from narrowband signal using different spectral
characteristics. If system is known, make
matched filter in frequency domain.

Finding the space


Fourier transforms
Use Fourier space methods to turn
differential equations or recursion relations
into algebraic forms.
x(n) is observed;
x(f) = ^x(n)exp[i27rnf]
is used.

Phase-space reconstruction
Using timelagged variables, form coordinates for the phase space in d dimensions:
y{n)=[x(n),x(n+T),

. . . ,x(n + ( d -

\)T)]

How are we to determine d and 77 Mutual


information; False neighbors.
Classify the signal
Invariants of orbits. Lyapunov exponents;
Various fractal dimensions; Topological invariants; Linking numbers of unstable
periodic orbits

Sharp spectral peaks

Resonant frequencies of the system


Quantities independent of initial conditions

Quantities independent of initial conditions


Make models, predict

x(n +\) = ^CjX{n

j)

Find parameters Cj consistent with invariant classifiers (spectral peaks)

y (n) +y (n 4- 1) as time evolution


y(n+l)=F[y(n)9al9a2,

. . . ,ap]

Find parameters a7 consistent with invariant classifiers (Lyapunov exponents, dimensions). Find dynamical dimensions dL from
data.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Abarbanel et a/.: Analysis of observed chaotic data


space. If the source has transients or significant highfrequency content with many localized events (in the
time domain), then a linear transform, such as wavelets,
that is adapted to such localized events will be more useful. Under our assumption that the underlying physical
processes are governed by differential equations, we
would expect them to have time-independent coefficients
when the process is stationary, so Fourier space is again
suggested, as it will reduce the differential equations to
algebraic equations, which are much more tractable.
Much of the contemporary toolkit of the signal processor
rests on this ability to move to frequency domain and
perform matrix manipulations on that representation of
the measurements.
In the case of a nonlinear source, we are not likely to
find any simplification from Fourier-transforming the
scalar measurements s{n), since the processes that give
rise to chaotic behavior are fundamentally multivariate.
So we need to reconstruct the phase space of the system
as well as possible using the information in s(n). This
reconstruction process will result in vectors in a ddimensional space. How we choose the components of
the d-dimensional vectors and how we determine the
value of d itself now become central topics. The answer
will lie in a combination of dynamical ideas about nonlinear systems as generators of information and geometrical ideas about how one unfolds an attractor using coordinates established on the basis of their informationtheoretic content. The result of the operations is a set of
d-dimensional vectors y(n) which replace the scalar data
we have observed.

C. Classification and identification


With a clean signal and the proper phase space in
which to work, we can proceed to ask and answer interesting questions about the physical source of the observations. One critical question we need to address is
that of classifying or identifying the source. In the case
of linear systems we have Fourier analysis to fall back on
again. The locations of the sharp lines in the Fourier
spectrum are characteristic of the physical system being
measured. If we drive the linear system harder, this will
result in more power under each of these lines, and if we
start the system at a different time, the phase of the signal will be altered, but the location of the Fourier peaks
will remain unchanged. Thus the characteristic frequencies, usually called resonant frequencies for the linear system, are invariants of the dynamics and can be used to
classify the physics. If we see a particular set of frequencies emerging from a driven linear system, we can recognize that set of frequencies at a later date and distinguish
the source with confidence. Recognizing the hydrogen
spectrum and codifying it with the Rydberg law is a classic (no pun intended) example of this. Similarly classifying acoustic sources of their spectral content is a familiar
operation.

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1341

Nonlinear systems do not have useful spectral content


when they are operating in a chaotic regime, but there
are other invariants that are specific in classifying and
identifying these sources. These invariants are quantities
that are unchanged under various operations on the dynamics or the orbit. The unifying ingredient is that they
are all unchanged under small variations in initial conditions. Some of the invariants we discuss have the property of being unchanged under the operation of the dynamics x>F(x). This guarantees that they are insensitive to
initial conditions, which is definitely not true of individual orbits. Some of the invariants are further unchanged
under any smooth change of coordinates used to describe
the evolution, and some, the topological invariants, are
purely geometric properties of the vector field describing
the dynamics. Among these invariants are the various
fractal dimensions andeven more useful for physics
the local and global Lyapunov exponents. These latter
quantities also establish the predictability of the nonlinear source. Chaotic systems are notorious for the
unpredictability of their orbits. This, of course, is directly connected with the instabilities in phase space we discussed earlier. The chaotic motion of a physical system
is not unpredictable even with all the instabilities one encounters in traversing state space. The limited predictability of the chaos is quantified by the local and global
Lyapunov exponents, which can be determined from the
measurements themselves. So in one sense, chaotic systems allow themselves to be classified and to have their
intrinsic predictability established at the same time.
One of the hallmarks of chaotic behavior is the sensitivity of any orbit to small changes in initial condition or
small perturbations anywhere along the orbit. Because of
this sensitivity, which is quantified in the presence of positive Lyapunov exponents, it is inappropriate to compare
two orbits of a nonlinear system directly with each other;
generically, they will be totally uncorrelated. The invariants of the geometric figure called the attractor, which
each orbit visits during its evolution, will be the same.
These are independent of initial conditions and provide a
direct analogy to the Fourier frequencies of a linear
dynamical system. So these invariants can be compared
if we wish to establish whether the observations of the
two orbits mean they arise from the same physical system. System identification in nonlinear chaotic systems
means establishing a set of invariants for each system of
interest and then comparing observations to that library
of invariants.
As noted above there are also topological invariants of
any dynamical system which can be used to the same
effect. These are not dynamical in the same fashion as
the Lyapunov exponents, but can serve much the same
function as regards classification. Metric invariants such
as Lyapunov exponents or the various fractal dimensions
are unlikely to provide a "complete" set of invariants, so
using them and the topological invariants together may
provide sufficient discrimination power to allow for the
practical separation of observed physical systems.

Abarbanel et a/.: Analysis of observed chaotic data

1342

D. Modelinglinear and nonlinear


With all this analysis of the source of the measurements now done, we may address the problem of building
models of the physical processes acting at the source.
The general plan for this was outlined in the Introduction, and here we focus only on the first kind of model
making: working within the coordinate system established by the analysis of the signal to build up local or
global models of how the system evolves in the reconstructed phase space. In linear systems the task is relatively easy. One has observations s(n) and they must
somehow be linearly related to observations at earlier
times and to the driving forces at earlier times. This
leads to models of the form
N

s(n)=

<*ks(n -k)+

k=\

*/g(/i ~ ' ) ,

(13)

1=1

where the coefficients {ak} and {hl} are to be determined


by fits to the data, typically using a least-squares or an
information-theoretic criterion, and the gin) are some
deterministic or stochastic forcing terms, which are
specified by the modeler. This linear relationship becomes simpler and algebraic when we take the "z transform" by defining
+ 00

S(z)=

zns(n),

(14)

n = oo

which leads to

i b,z'
S(z) = ^

G(z) ,

(15)

k= \

with G(z) the z transform of the forcing. This is just a


discrete form of Fourier transform, of course.
Once the coefficients are established by one criterion or
another the model equation (13) is then used for prediction or as part of a control scheme. The choice of
coefficients should be consistent with any knowledge one
has of spectral peaks in the system, and the denominator
in the z transform holds all that information in terms of
poles in the z plane. Much of the literature on linear signal processing can be regarded as efficient and thoughtful
ways of choosing the coefficients in this kind of linear
model in situations of varying complexity. From the
point of view of dynamical systems this kind of model
consists of simple linear dynamicsthe terms involving
the {ak}and simple linear averaging over the forcing.
The first part of the modeling, which involves dynamics,
is often called autoregressive (AR) and the second moving average (MA), and the combination labeled A R M A
modeling.
This kind of model will always have zero or negative
Lyapunov exponents, zero KS entropy, and will never be
chaotic. It is, as a global model of interesting nonlinear
physics, not going to do the job.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Nonlinear modeling of chaotic processes revolves


around the idea of a compact geometric attractor on
which our observations evolve. Since the orbit is continually folded back on itself by the dissipative forces and
the nonlinear part of the dynamics, we expect to find in
the neighborhood of any orbit point y() other orbit
points y{r\n);r = 1,2, . . . ,NRy which arrive in the neighborhood of y(n) at quite different times than n. One can
then build various forms of interpolation function, which
account for whole neighborhoods of phase space and how
they evolve from near y(n) to the whole set of points near
y ( + l ) . The use of phase-space information in the
modeling of the temporal evolution of the physical process is the key innovation in modeling chaotic processes.
The general method would work for regular processes,
but is likely to be less successful because the neighborhoods are less populated.
The implementation of this idea is to build
parametrized nonlinear functions F ( x , a ) which take y(n)
into y(n + 1 ) = F ( y ( ) , a ) and then use various criteria to
determine the parameters a. Further, since one has the
notion of local neighborhoods, one can build up one's
model of the process neighborhood by neighborhood and,
by piecing together these local models, produce a global
nonlinear model that captures much of the structure in
the attractor itself.
Indeed, in some sense the main departure from linear
modeling is to realize that the dynamics evolves in a multivariate space whose size and structure is dictated by the
data itself. N o clue is given by the data as to the kind of
model that would be appropriate for the source of chaotic data. Indeed, since quite simple polynomial models
and quite complex models can both give rise to timeasymptotic orbits with strange attractors, the critical
part of the modelingthe connection to the physics
remains in the hands of the physicist. It is likely there is
no algorithmic solution (Rissanen, 1989) to how to
choose models from data alone, and that, of course, is
both good news and bad news.

E. Signal synthesis
The main emphasis of this review and actually of most
work on nonlinear signals has been on their analysis; that
is, given an observed signal, can we tell if it came from a
low-dimensional chaotic attractor, can we characterize
that attractor, can we predict the evolution of state-space
points near the attractor or into the future, etc.
Another direction now being taken concerns the synthesis of new signals using chaotic sources. Can one utilize these signals for communication between distant locations or synchronization of activities at these locations,
etc. This, in our opinion, is likely to be the main direction of engineering applications of the knowledge we
have garnered about nonlinear systems operating in a
chaotic mode. We shall touch upon these ideas as we
proceed here, but their review will have to await a more

Abarbanel et a/.: Analysis of observed chaotic data


substantial set of developments, and this we are confident
will be available over the next few years.

IV. RECONSTRUCTING PHASE SPACE


OR STATE SPACE
We now turn to methods for reconstructing the state
space of a system from information on scalar measurements s(n)=s(t0
+ nrs). If more than just a single scalar
were measured, that would put us in an advantageous position, as will become clear, but for the moment we
proceed with one type of measurement.
From the discussion of dynamical systems we see that
if we were able to establish values for the time derivatives
of the measured variable, we could imagine finding the
connection among the various derivatives and the state
variables, namely, the differential equation, which produced the observations. To create a value for

we need to approximate the derivative, since we have


measurements only every rs. The natural approximation
would be
()

s(t0 + {n +\)Ts)-s(t0

+ nrs)

(17)

With finite rs this is just a crude high pass filter of the


data, and we are producing a poor representation of the
time derivative. When we next try to estimate s(n) with
second differences, we are producing an even poorer version of the second derivative, etc. If we examine the formula for the derivatives, we see that at each step we are
adding to the information already contained in the measurement s(n) measurements at other times lagged by
multiples of the observation time step rs. In approximately 1980 both a group at the University of California,
Santa Cruz (Packard et aL, 1980) and David Ruelle apparently simultaneously and independently introduced
the idea of using time-delay coordinates to reconstruct
the phase space of an observed dynamical system. The
Santa Cruz group (Packard et aL, 1980) implemented the
method in an attempt to reconstruct the time derivatives
of s(t). The main idea is that we really do not need the
derivatives to form a coordinate system in which to capture the structure of orbits in phase space, but that
we
could
directly
use
the
lagged
variables
s(n -\-T)=s(t0-\-(n
+T)rs)9 where T is some integer to
be determined. Then using a collection of time lags to
create a vector in d dimensions,
y(n) = [s(n),s(n+T),s(n

1343

comprise the source of the measurements. While we do


not know the relationship between the physical variables
u() and the s(n +jT), for purposes of creating a phase
space, it does not matter, since any smooth nonlinear
change of variables will act as a coordinate basis for the
dynamics. This idea was put on a firmer footing by Mane
and Takens (Mane, 1981; Takens, 1981) and further
refined relatively recently (Sauer et al.y 1991).
To see this idea in action we take data from the Lorenz
attractor described earlier. We take points from the variable x(n),
and in Fig. 7 we plot the points
[x(n),x(n
-\-dT),x{n -\-2dT)} in R3. Comparing this to
Fig. 3, we see that from the complicated time trace in
Fig. 1 we have created, by using 3-vectors of timedelayed variables, a geometric object similar in appearance, though somewhat distorted, to the original
geometric portrait of the strange attractor. The distortion
is
not
surprising
since
the
coordinates
[x{n)9x(n +dT),x(n
+2dT)]
are
different
from
[x(n),y(n),z(n)].
Now how shall we choose the time lag T and the dimension of the space d to go into the timelagged vectors
y(), Eq. (18)? The arguments of Mane and Takens suggest that it does not matter what time lag one chooses,
and a sufficient condition for d depends on the dimension
of the attractor dA. They argue that if d, which is an integer, is larger than 2dA, which can be fractional, then
the attractor as seen in the space with lagged coordinates
will be smoothly related to the attractor as viewed in the
original physical coordinates, which we do not know. In
practice, their theorem tells us that, if we have chosen d
large enough, physical properties of the attractor that we
wish to extract from the measurements will be the same
when computed on the representative in lagged coordinates and when computed in the physical coordinates.
The procedure of choosing sufficiently large d is formally
known as embedding, and any dimension that works is
called an embedding dimension dE.
Once one has
achieved a large enough d=zdE, then any d>.dE will also

+2T), . . . ,s(n +(d - 1 ) D ] ,


(18)

we would have provided the required coordinates. In a


nonlinear system, the s (n +jT) are some (unknown) nonlinear combination of the actual physical variables that
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

FIG. 7. Lorenz time series x(t) (Fig. 1) embedded in a threedimensional phase space (x(t)yx(t -\rdT),x(t +2dT)),dT =0.2.

Abarbanel et a/.: Analysis of observed chaotic data

1344

provide an embedding.
Time-delay embedding is certainly not the only
method for embedding scalar data in a multivariate
dynamical environment. Other methods (Landa and
Rosenblum, 1989; Mindlin et al.y 1991) are available and
have their advantages. Time-delay embedding is,
perhaps, the only systematic method for going from scalar data to multidimensional phase space, and it has certainly been the most explored in the literature.
A. Choosing time delays
The statement of the embedding theorem that any time
lag will be acceptable is not terribly useful for extracting
physics from the data. If we choose T too small, then the
coordinates s(n+jT)
and s(n -\-(j + l)T) will be so
close to each other in numerical value that we cannot distinguish them from each other. From any practical point
of view they have not provided us with two independent
coordinates. Similarly, if T is too large, then s(n +jT)
and s(n +{j +1)T) are completely independent of each
other in a statistical sense, and the projection of an orbit
on the attractor is onto two totally unrelated directions.
The origin of this statistical independence is the ubiquitous instability in chaotic systems, which results in any
small numerical or measurement error's being amplified
exponentially in time. A criterion for an intermediate
choice is called for, and it cannot come from the embedding theorem itself or considerations based on it, since
the theorem works for almost any value of T.
One's first thought might be to consider the values of
sin) as chosen from some unknown distribution. Then
computing the linear autocorrelation function
~
"

CL(r)--

[s(m+r)-s][s(m)-s]

m = l

(19)
[s(m)-s]2

771)2
+r)-s-CL(r)[s(n)-s]}

{s(n

(21)

n= l

with respect to CL(r), immediately leads to the definition


of CL(T) above.
Choosing Tto be the first zero of CL(r) would then, on
average over the observations, make s (n) and s (n +7")
linearly independent. What this may have to do with
their nonlinear dependence or their utility as coordinates
for a nonlinear system is not addressed by all this. Since
we are looking for a prescription for choosing T7, and this
prescription must come from considerations beyond
those in the embedding theorem, linear independence of
coordinates may serve, but we prefer another point of
view, one that stresses an important aspect of chaotic
behaviornamely the viewpoint of information theory
(Shaw, 1984; Fraser and Swinney, 1986; Fraser, 1989a,
1989b)and leads to a nonlinear notion of independence.
B. Average mutual information
In the Introduction we discussed how the idea of information created by chaotic behavior is suggested in a
qualitative fashion by the KS entropy and the growth in
time of the number of states of a system available to a
small piece of phase space as 2^th{x)\
The same idea can
be used to identify how much information one can learn
about a measurement at one time from a measurement
taken at another time. In a general sort of context, let us
imagine that we have two systems, call them A and B,
with possible outcomes in making measurements on them
ai and bk. We consider there to be a probability distribution associated with each system governing the possible
outcomes of observations on them. The amount one
learns in bits about a measurement of at from a measurement of bk is given by the arguments of information
theory (Gallager, 1968) as

m=\

where

^AB^i^k)

= l0g2

m 1

and looking for that time lag where CL(r) first passes
through zero, would give us a good hint of a choice for T.
Indeed, this does give a good hint. It tells us, however,
about the independence of the coordinates only in a
linear fashion. To see this, recall that if we want to know
whether two measurements sin) and s(n + r ) depend
linearly on each other on the average over the observations, we find that their connection, in a least-squares
sense, is through the correlation matrix just given.
That is, if we assume that the values of s(n) and
s (n + r ) are connected by
s(n+T)-J=CL(r)[s(n)-J]
then minimizing
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

(20)

PAB^i^k)

(22)

PA^i)PB(bk)

where the probability of observing a out of the set of all


A is PA(a)9 and the probability of finding b in a measurement of B is PB(b), and the joint probability of the measurement of a and b is PAB(a,b).
This quantity is called
the mutual information of the two measurements at and
bk and is symmetric in how much one learns about bk
from measuring at. The average mutual information between measurements of any value at from system A and
bk from B is the average over all possible measurements
ofIAB(ai9bk),
*AB(T)=

PAB(aiybk)IABiat,bk)

(23)

To place this abstract definition into the context of observations from a physical system s(n)> we think of the
sets of measurements s (n) as the A set and of the mea-

Abarbanel et al.: Analysis of observed chaotic data


Lorenz 63; 8192 points

surements a time lag T later, s (n +T), as the B set. The


average mutual information between observations at n
and n +T, namely, the average amount of information
about s(n +T) wt have when we make an observation of
s(n), is then
I(T)=

P(s(n),s(n+T))\og2

1345

Average mutual information

P(s(n),s(n
+;))
P(s(n))P(s(n+T))
(24)

and / (T) > 0 (Gallager, 1968).


The average mutual information is really a kind of generalization to the nonlinear world from the correlation
function in the linear world. It is the average over the
data or equivalently the attractor of a special statistic,
namely, the mutual information, while the correlation
function is the average over a quadratic polynomial
statistic. When the measurements of systems A and B
are completely independent, PAB(a,b)=Pa(a)Pb(b),
and
Evaluating I(T) from data is quite straightforward.
To find P(s(n)) we need only take the time trace s(n)
and project it back onto the s axis. The histogram
formed by the frequency with which any given value of s
appears gives us P(s(n)).
If the time series is long and
stationary, then P(s (n + 70) is the same as P(s (n)). The
joint distribution comes from counting the number of
times a box in the s (n) versus s (n +T) plane is occupied
and then normalizing this distribution. All this can be
done quite easily and rather fast on modern PC's or
workstations.
In his dissertation (Fraser, 1988) Fraser gives a very
clever recursive algorithm, which is basically a search algorithm for neighbors in d-dimensional space, for
evaluating I (T) and its generalizations to higher dimensions. The more appropriate question to answer when
creating d-dimensional vectors y(n) out of time lags is
this: what is the joint mutual information among a set of
measurements s(n)9s(n -f-F),. . . ,s(n +(d l)T) as a
function of 77 For our purposes of giving a rule for an
estimation of a useful time lag T, it will suffice to know
I (T) as we have defined it.
Now we have to decide what property of I(T) we
should select, in order to establish which among the various values of T we should use in making our data vectors
y(n). If T is too small, the measurements s(n) and
s (n + T) tell us so much about one another that we need
not make both measurements. If T is large, then J (T)
will approach zero, and nothing connects s(n) and
s (n + T), so this is not useful. Fraser and Swinney (1986)
suggest, as a prescription, that we choose that Tm where
the first minimum of I(T) occurs as a useful selection of
time lag T. As an example of this we show in Figs. 8 and
9 the average mutual information for the Lorenz attractor and for the data from the Carroll-Pecora nonlinear
circuit. Each has a first minimum at some Tm> and this
minimum can be used as a time lag for phase-space
reconstruction. The lag Tm is selected as a time lag
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

0.2

0.3
Time lag

FIG. 8. Average mutual information / ( D a s a function of time


lag T for Lorenz data.

where the measurements are somewhat independent, but


not statistically independent:
P(s(n)9s(n

+Tm))=P(s(n))P(s(n

+Tm))

Recognizing that this is a prescription, one may well


ask what to suggest if the average mutual information
has no minimum. This occurs when one is dealing with
maps, as the I(T) curve from x(n) data taken from the
Henon map (Henon, 1976)

x(n+l)=l+y(n)-1.4xM2

,
(25)

y(n+l)

0.3x(n)

shows in Fig. 10. Similarly, if one integrates the Lorenz


equations with a time step dt in the integrator, which is
then increased, the minimum seen in Fig. 8 fills in as dt
grows and eventually disappears. This does not mean
that J (T) loses its role as a good grounds for selection of
Tm, but only that the first minimum criterion needs to be
replaced by something representing good sense. Without
much grounds beyond intuition, we use Tm = 1 or 2 if we
Hysteretic circuit
Average mutual information

FIG. 9. Average mutual information for hysteretic circuit data.

Abarbanel et a/.: Analysis of observed chaotic data

1346

Henon map, 8192 points


Average mutual information
7.0

6.0

ormation

5.0

4.0

Mutual

g
3.0

2.0

1.0

0.0

FIG. 10. Average mutual information for data generated by the


Henon attractor [Eq. (25)].
know the data comes from a map, or choose Tm such
that J C T m ) / J ( 0 ) j . This is clearly an attempt to
choose a useful Tm in which some nonlinear decorrelation is at work, but not too much.
Since this is prescriptive, one may compare it to the
prescription used in linear dynamics of choosing a time
lag r m such that CL(rm ) = 0 for the first time. Why not
the second zero or the fifth? Grassberger et ah (1991)
scrutinize the use of a simple prescription for choosing T.
The authors point out that for high-dimensional data one
must use a more sensitive criterion than the minimum
average mutual information between x(n) and x (n +7 1 ),
especially if one allows differing lags in each component
of the data vector y(n). Recognizing, as we have
stressed, that the choice of T is prescriptive, we agree
with their caution that "we do not believe that there exists a unique optimal choice of delay." Nonetheless, it is
useful to have a general rule of thumb as a guide to a delay T that is workable; seeking the optimum is likely to
be quite unrewarding.
Liebert and Schuster (1989) have examined the first
minimum criterion by adding another criterion involving
the so-called correlation integral, which we shall discuss
below, and concluded that the Tm of the first minimum is
that Tm in which the correlation integral is also numerically well established. This is yet another prescription,
and we have no special bent toward any particular
prescription, though when the I(T) curve does have
minima, we find that to be a practical and useful choice.
The main idea is to use criteria based on correlation
properties tuned to aspects of chaotic systems, namely,
properties that are attuned to chaotic dynamics as information generators, and average mutual information certainly does just that.
C. Choosing the embedding dimension
The embedding theorem of Marie and Takens rests on
topological results discussed sixty years ago (Whitney,
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1936). It is really quite independent of dynamics and is a


generalization of ideas on regular geometries. The goal
of the reconstruction theorem is to provide a Euclidean
space R d large enough so that the set of points of dimension d A can be unfolded without ambiguity. This means
that if two points of the set lie close to each other in some
dimension d they should do so because that is a property
of the set of points, not of the small value of d in which
the set is being viewed. For example, if the attractor is
one dimensional and it is viewed in d = 2 , it may still be
that the one-dimensional line crosses itself at isolated
points. At these points there is an ambiguity about
which points are neighbors of which other points. This
ambiguity is completely resolved by viewing the same
one-dimensional attractor in d =3, where all possible
zero-dimensional or point crossings of the set on itself are
resolved. Of course, if one wants to view the set of points
in d = 4 , that is fine too. N o further ambiguities are introduced by adding further coordinates to the viewing or
embedding space. When all ambiguities are resolved, one
says that the space R d provides an embedding of the attractor.
An equivalent way to look at the embedding theorem
is to think of the attractor as comprised of orbits from a
system of very high dimension, even infinite dimensional
as in the case of delay differential equations or partial
differential equations. The attractor, which has finite dA,
lies in a very small part of the whole phase space, and we
can hope to provide a projection of the whole space down
to a subspace in which the attractor can be faithfully captured. Most of the large space, after all, is quite empty.
If this projection takes us down to too small a space, as
for example the scalar one-dimensional measurements
themselves are almost sure to do, then the evolution as
seen in this too small space will be ambiguous because of
the presence of incorrect neighboring points. In this too
small space the appearance of the orbit will be unduly
complicated, and part of the bad name of chaos or irregular behavior comes from looking at the orbits in the
wrong space. The appearance of the attractor when fully
unfolded in the larger embedding space is much more
regular and attractive than when projected down to one
dimension as in the scalar measurements. The embedding theorem provides a sufficient condition from geometrical considerations alone for choosing a dimension dE
large enough so that the projection is goodi.e., without
orbit crossings of dimension zero, one, two, etc.
The sufficient integer dimension dE is not always
necessary. For the Lorenz attractor, for example, we
have dA=2.06
(we discuss later how to establish this).
This sufficient condition would suggest, since
dE>2dA,
that if we choose to view the Lorenz attractor in dE = 5
we can do so without ambiguity. This is not at all inconsistent with our knowledge of the Lorenz differential
equations, which provide three coordinates for an R3 in
which to view the attractor, since we are speaking now
about a time-delayed version of one of the Lorenz variables, and these are nonlinear combinations of the origi-

Abarbanel et air. Analysis of observed chaotic data


nal x (t),y (t),z (t) which my twist the attractor onto itself
in only three dimensions. The embedding theorem
guarantees us that, however complicated these twists
may be, if the nonlinear transformation is smooth, dE = 5
will untwist them globally in phase space. The theorem
is silent on local phase-space properties.
Why should we care if we always work in a dimension
dE that is sufficient to unfold the attractor but may be
larger than necessary? Would it not be just fine to do all
analysis of the properties of the Lorenz system in dE = 5?
F r o m the embedding theorem point of view, the answer
is always yes. For the physicist more is needed. Two
problems arise with working in dimensions larger than
really required by the data and time-delay embedding: (i)
many of the computations we shall discuss below for extracting interesting properties from the data, and that is
after all the goal of all this, require searches and other
operations in ^ ^ whose computational cost rises exponentially with d; and (ii) in the presence of "noise" or
other high-dimensional contamination of our observations, the "extra" dimensions are not populated by dynamics, already captured by a smaller dimension, but entirely by the contaminating signal. In too large an
embedding space one is unnecessarily spending time
working around aspects of a bad representation of the observations which are solely filled with "noise."
This realization has motivated the search for analysis
tools that will identify a necessary embedding dimension
from the data itself. There are four methods we shall discuss: (i) singular-value decomposition of the sample covariance matrix; (ii) "saturation" with dimension of some
system invariant; (iii) the method of false nearest neighbors, (iv) the method of true vector fields.

1. Singular-value analysis
If our measurements y(n) are composed of the signal
from the dynamical system we wish to study plus some
contamination from other systems, then in the absence of
specific information about the contamination it is plausible to assume it to be rather high dimensional and to assume that it will fill more or less uniformly any fewdimensional space we choose for our considerations. Let
us call the embedding dimension necessary to unfold the
dynamics we seek dN. If we work in dE>dN, then in an
heuristic sense dE dN dimensions of the space are being
populated by contamination alone. If we think of the observations embedded in dE as composed of a true signal
yT(n) plus some contamination c: y(n) = y r ( n ) + c(n),
then the dE X dE sample covariance matrix
C O V =

4F
iV

[y(")-y a v][yU)-y a v] 7 \

(26)

n= 1

with
y a v ^ i ; y(*),
iV

n= 1

Rev. Mod. Phys., Vol. 65, No, 4, October 1993

(27)

1347

will, again in an heuristic sense, have dN eigenvalues arising from the variation of the (slightly contaminated) real
signal about its mean and dEdN eigenvalues which
represent the "noise." If the contamination is quite high
dimensional, it seems plausible to think of it filling these
extra dE dN dimensions in some uniform manner, so
perhaps one could expect the unwelcome dEdN eigenvalues, representing the power in the extra dimensions, to
be nearly equal. If this were the case, then by looking at
the eigenvalues, or equivalently the singular values of
COV, we might hope to find a "noise floor" at which the
eigenvalue spectrum turned over and became flat
(Broomhead and King, 1986). There are dE eigenvalues,
and the one where the floor is reached may be taken as
d

N-

This analysis can also be carried out locally (Broomhead and Jones, 1989), which means that the covariance
matrix is over a neighborhood of the NB nearest neighbors, y{r)(n) of any given data point y(n):
1

COV(n) = ~

[y{r)(n)-yJn)][y{r)(n)-ysiV(n)]T
(28)

(r,

y.v<*>=^iy <)The global singular-value analysis has the attractive


feature of being easy to implement, but it has the downside of being hard to interpret on occasion. It gives a
linear hint as to the number of active degrees of freedom,
but it can be misleading because it does not distinguish
two processes with nearly the same Fourier spectrum
(Fraser, 1989a, 1989b) or because of differing computers
the anticipated "noise floor" is reached at different numerical levels. Using it as a guide to the physicist, to be
looked at along with other tools, can be quite helpful.
Local covariance analysis can provide quite useful insights as to the structure of an attractor and, as emphasized by Broomhead and Jones (1989), can be used to
distinguish degrees of freedom in a quantitative fashion.
The local analysis is also useful in certain signalseparation algorithms when one of the signals to be
separated has substantial resemblance to white noise
(Sauer, 1991; Cawley and Hsu, 1992).
2. Saturation of system invariants
If the attractor is properly unfolded by choosing a
large enough dE, then any property associated with the
attractor which depends on distances between points in
the phase space should become independent of the value
of the embedding dimension once the necessary dE has
been reached. Increasing dE beyond this dN should not
affect the value of these properties, and, in principle, the
appropriate necessary embedding dimension dN can be
established by computing such a property
for
dE 1,2, . . . until variation with dE ceases.
A familiar example of this comes from the average

Abarbanel et aL: Analysis of observed chaotic data

1348

Ikeda map: correlation integrals

over the attractor of moments of the number density


n(r,x):
the number of points on the attractor or
equivalently on the orbit within a radius of r of points x
in the phase space, n (r,x), is defined by
(M)=-^2

0(r-\y(k)-x\),

7 500 points, nT=1, dE=1 ,...,8

(29)

W k= i

with 6(u) = 0, u < 0 ; 0 ( M ) = 1 , u > 0 , and the average over


all points of powers of n(r>x) is (Grassberger and Procaccia, 1983a, 1983b)
M

Cq(r)-

[n(r,y{j))iq-

(30)

7= 1

We shall show in Sec. V that averages such as these are


independent of initial conditions and are thus good candidates for characterizing an attractor. We shall also elaborate on the discussion of the correlation integrals Cq(r)
in Sec. V.D. Now we want to concentrate on the
fact that this quantity depends on the embedding
dimension dE
used to make the vectors y(k)
= [s(n),s(n+T),.
. . ,s(n +
(dE-\)T)}.
We can evaluate Cq(r) as a function of dE and determine when the slope of its logarithm as a function of
log(r) becomes independent of dE. In Fig. 11 we show a
log-log plot of C2{r) for data from the Lorenz attractor
as a function of dE. It is clear that for dE 3, or perhaps
dE=z4y the slope of the function C2(r) becomes independent of embedding dimension, and thus we can safely
choose dE = 3 or dE=49 which is less than the sufficient
condition from the embedding theorem of dE 5.
While what we have described is quite popular, it does
not give any special distinction, from the point of view of
the embedding theorem, to Cq(r) as the function whose
independence of dE we should establish. As we shall see
in the next section, there is a very large class of functions
one could consider, all of which are equivalent from the
point of view of embedding.
Indeed, it is our own opinion that determining dN by
looking for the independence of some function on the attractor, whether it be Cq(r) or another choice, is somewhat of an indirect answer to the direct geometric question posed by the embedding theorem: when have we unfolded the attractor by our choice of time-delay coordinates?
Rd(k)2

= [s(k)sNN(k)]2+[s(k+T)-sNN(k+T)]2+

[s{k+dT)--s(k+dT)]

(33)

If Rd + l{k) is large, we can presume it is because the near


neighborliness of the two points being compared is due to
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

False nearest neighbors

The third method for determining dN comes from asking, directly of the data, the basic question addressed in
the embedding theorem: when has one eliminated false
crossings of the orbit with itself which arose by virtue of
having projected the attractor into a too low dimensional
space?
Answers to this question have been discussed in various ways. Each of the ways has addressed the problem of
determining when points in dimension d are neighbors of
one another by virtue of the projection into too low a dimension. By examining this question in dimension one,
then dimension two, etc. until there are no incorrect or
false neighbors remaining, one should be able to establish, from geometrical considerations alone, a value for
the necessary embedding dimension dE=dN.
We describe the implementation of Kennel et al. (1992).
In dimension d each vector
y(k) = [s(k),s(k

+T),.

. . ,s(k +(d

-DT)]

(31)

NN

has a nearest neighbor y (k) with nearness in the sense


of some distance function. Euclidean distance is natural
and works well. The Euclidean distance in dimension d
between y(k) and yNN(k) we call Rd(k):
+[s(k +T(d

Rd(k) is presumably small when one has a lot of data,


and for a data set with N entries, this distance is more or
less of order \/Nl/d.
In dimension d + 1 this nearestneighbor distance is changed due to the {d -f l) 5 ' coordinates s(k+dT)
and sNN(k + dT) to
Rj+x{k)=Rj(k)

FIG. 11. Sequence of correlation integrals C 2 (r) for data from


the Ikeda map in embedding dimensions dE ~ 1>. > 8.

-l))-sNN(k

+T(d

-l))]2

(32)

the projection from some higher-dimensional attractor


down to dimension d. By going from dimension d to dimension d + l, we have "unprotected" these two points
away from each other. Some threshold size RT is required to decide when neighbors are false. Then if

\s{k+Td)-s""(k+Td)
Rd(k)

>R

T >

(34)

the nearest neighbors at time point k (or t0 4- krs) are de

Abarbanel et al.: Analysis of observed chaotic data


clared false. In practice, for values of RT in the range
10<RT < 5 0 the number of false neighbors identified by
this criterion is constant. With such a broad range of independence of R T one has confidence that this is a workable criterion.
When this criterion is applied to various standard
models, as shown in Figs. 1 2 - 1 4 , one finds results that
are quite sensible. For the Lorenz attractor, one finds
that the number of false nearest neighbors drops to zero
at dN = 3, while the sufficient condition from the embedding theorem is at dE = 5. For the Henon map, we find
dN = 2, while the embedding theorem only guarantees us
success in unfolding the attractor at dE=39
since
dA = 1.26 for this map.
For the Ikeda map (Ikeda, 1979; Hammel et a/., 1985)
of the complex plane z (k)=x (k) + iy (k) to itself,
z(fc+!)=/>+JBz(fc)exp

IK-

ia

(l + |z(fc)|2)

Henon map; False Nearest Neighbors


7 000 points, T=1

Embedding dimension

FIG. 13. Percentage of false nearest neighbors as a function of


embedding dimension for Henon map data.

(35)

which describes the evolution of a laser in a ring cavity


with a lossy active medium, we find, with "standard" parameters / > = 1 . 0 , 2? = 0 . 9 , fc=0.4, and a = 6.0, where
d ^ 1 . 8 , that we require dN = A to eliminate all false
neighbors. This is the same dimension that the embedding theorem tells us is sufficient. In a sense this example
tells us that the false-nearest-neighbor test is really quite
powerful. While one needs only a phase space of dimension two for the original map, the test reveals that the
choice of time delays as coordinates for a reconstructed
state space twists and folds the attractor so much that in
these coordinates these features of the coordinate system
require a larger dimension to be undone. Comparing the
actual [x(k)9y(k)]
phase plane with the [x(k),x(k
-hi)]
phase portrait as in Figs. 15 and 16 makes this twisting
quite apparent.
The criterion stated so far for false nearest neighbors
has a subtle defect. If one applies it to data from a veryhigh-dimensional random-number generator, such as is

Lorenz 63; False Nearest Neighbors


7 000 points, T=0.1

1349

found on any modern computer, it indicates that this set


of observations can be embedded in a small dimension. If
one increases the number of points analyzed, the apparent embedding dimension rises. The problem is that
when one tries to populate uniformly (as "noise" will try
to do) an object in d dimensions with a fixed number of
points, the points must move further and further apart as
d increases because most of the volume of the object is at
large distances. If we had an infinite quantity of data,
there would be no problem, but with finite quantities of
data eventually all points have "near neighbors" that do
not move apart very much as dimension is increased. As
it happens, the fact that points are nearest neighbors does
not mean they are close on a distance scale set by the approximate size RA of the attractor. If the nearest neighbor to y(fc) is not close, so Rd(k)^RA>
then the distance
JRrf + 1(fc) will be about 2Rd(k). This means that distant,
but nearest, neighbors will be stretched to the extremities
of the attractor when they are unfolded from each other,
if they are false nearest neighbors.

False Nearest Neighbors


Ikeda Map; 7500 Points

40.0

Embedding dimension

FIG. 12. Percentage of false nearest neighbors as a function of


embedding dimension for clean Lorenz data.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

FIG. 14. Percentage of false nearest neighbors as a function of


embedding dimension for data from the Ikeda map.

Abarbanel et a/.: Analysis of observed chaotic data

1350

Ikeda Map
12 000 Points

t>

*"'

\\ *

<-*>"<*

J*?+*>

x(k)

FIG. 15. Ikeda attractor in two-dimensional phase space


(x(k)fy(k)).

This suggests as a second criterion for falseness of


nearest neighbors that if
>2,

(36)

then y( k) and its nearest neighbor are false nearest neighbors. As a measure of RA one may use the rms value of
the observations
R

2 _

[s(k)-s]2

*k%

1
N

(37)

k= \

The choice of other criteria for the size of the attractor


works as well. When nearest neighbors failing either of
these two criteria are designated as false, uniform noise
from a random-number generator is now identified as
high dimensional and scalar data from known low-

dimensional chaotic systems is identified as low dimensional.


This is a good occasion to emphasize a point that we
have more or less made implicitly during the discussion.
Noise, as commonly viewed, is thought of as completely
unpredictable and "random." When a practical characterization of noise is sought, one states that it has a
broadband Fourier spectrum and an autocorrelation
function that falls rapidly to zero with time. Since the
power spectrum and autocorrelation function are just
Fourier transforms of one another, these are more or less
the same statement. Just these properties are possessed
by low-dimensional chaos as well, and chaos cannot be
distinguished from noise on the basis of Fourier criteria.
Instead, we see in the false-nearest-neighbor method or
the other methods indicated for determining embedding
dimension a true distinguishing characteristic: noise appears as high-dimensional chaos. Qualitatively they are
the same. Quantitatively and practically one may not be
able to work with dimensions higher than ten or twenty
or whatever. It may be useful, then, in the sense of being
able to work with problems in physics, to adopt any of
the usual statistical techniques developed for random
processes for signals with dimension higher than this
threshold.
When one uses the false-nearest-neighbor test on measurements from a dynamical system that have been corrupted with "noise" or with signals from another dynamical system (low- or high-dimensional), there is a very
useful robustness in the results. Figure 17 shows the
effect of adding uniform random numbers lying in the interval [ L,L] to a x{n) signal from the Lorenz attractor. For the Lorenz attractor RA 1 2 , and the contamination level is given in units of L /R A in the figure. Until L /R A0.5
we see a definite indication of lowdimensional signals. When the contamination level is
low, the residual percentage of false neighbors gives an
indication of the noise level. In any case, when the perLorenz63; False Nearest Neighbors
Noise 0.0, 0.005, 0.01,0.05,0.1,0.5,1.0; RMS UR

FIG. 16. Time series from the Ikeda map embedded in a twodimensional space (x (k),x (k 4- 1)).
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

FIG. 17. Percentage of false nearest neighbors as a function of


embedding dimension for noisy Lorenz data: O, CT = 0.0; ,
o- = 0.005; 0 , a = 0 . 0 1 ; A, <r=0.05; *, a = 0.1; , cr=0.5; ,
(7=1.0.

Abarbanel et al.\ Analysis of observed chaotic data


Hysteretic circuit; False Nearest Neighbors
7 000 points, T=6
40.0 ,

r-

1351

ten, say, then at the present time we can think of dealing


with the problem using statistical methods.
To test for the deviation of the vector field from what
would be expected for a random motion, taken to be random walks, the statistic
(Vf)2-(R*)2

2
o
XJ

o> 20.0

Embedding dimension

FIG. 18. Percentage of false nearest neighbors as a function of


embedding dimension for hysteretic circuit data.

centage of false nearest neighbors has dropped below


1 % , one may choose that dimension as dN and have
some real sense of the error one might make in any subsequent calculations.
As an illustration of the method applied to real data,
we present in Fig. 18 the percentage of false nearest
neighbors for data from the hysteretic circuit of Carroll
and Pecora. Clearly, choosing dN = 3 would result in errors of less than 1 % in computing distances or in prediction schemes. This is also commensurate with the error
level in the observations.
4. True vector fields
Another geometrical approach to determining the
embedding dimension for observed time series data is
given by Kaplan and Glass (1992), who examine the
uniqueness of the vector field responsible for the dynamics. If the dynamics is given by the autonomous rule
x>-F(x), and F(x) is smooth, that is, differentiable, then
the tangents to the evolution of the system are smoothly
and uniquely given throughout the state space. Kaplan
and Glass establish the local vector field by carving up
phase space into small volumes and identifying where orbits enter and exit the volumes. This defines the local
flow under the assumption that the volumes are small
enough. "Small enough" is something one determines in
practice, but there seems no barrier to making the
volumes adequately small.
If one is in too small an embedding dimension, the vector fields in any location will often not be unique, as they
will overlap one another because of the projection of the
dynamics from a larger embedding space. As one increases the embedding dimension the frequency of overlap will fall to zero and the vector field will have been unfolded. If the dynamics is very high dimensional, then
the vector field will not be unfolded until this high dimension has been reached. If this dimension is above
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

is considered. The vectors Vj are the average directions


of the line segments from entry to exit of the trajectory as
it passes through box number j . R is the known average displacement per step for a random walk in dimension d after rij steps. This quantity will be averaged over
all boxes.
If the data set is drawn from a random walk, the statistic will be 0, and if it is effectively random by being a
high-dimensional deterministic system viewed in too low
a dimension, it will be small. For a low-dimensional system viewed in a large enough dimension this will be near
unity. The test appears to work quite well even with contaminated data. Indeed, the spirit of this method is quite
similar to that of false nearest neighbors. Both are
geometric and attempt to unfold a property of attractors
from its overlap on the observation axis to its unambiguous state in an appropriately high-dimensional state
space.

D.

Tan6dE

The determination of the appropriate phase space in


which to analyze chaotic signals is one of the first tasks,
and certainly a primary task, for all who wish to work
with observed data in the absence of detailed knowledge
of the system dynamics. We have spent some time on the
methods that have been devised for this, both because it
has been a subject of such intense interest, and because it
is important for the physicist to realize that not only are
there a variety of reliable techniques available for this job
but also the set of tools may be useful in combination and
as a complement to each other when the setting for the
data changes.
To determine the time lag to be used in an embedding,
one may always wish to use something nonlinear, such as
average mutual information, but the data may mitigate
against that. If one has sampled a map, achieved stroboscopically or taken as a Poincare section, there is typically no minimum in the average mutual information
function. The reason is quite simple: the time between
samples rs is so long that the orbit has become decorrelated, in an information-theoretic sense. One can see this
quite clearly by taking a flow such as the Lorenz model
and increasing the time step in the differential equation
solver step by step. At first, for a small time step, the
average mutual information has a minimum. As the time
step increases, the minimum moves in towards zero and,
since time steps are only in integer steps, disappears.

Abarbanel et al.\ Analysis of observed chaotic data

1352

What is one to do at this stage?


Typically, when one is handed data, there is nothing to
do about resampling it, but if that is an option, take it. If
not, one can turn to the autocorrelation function of the
time series to find at least an estimate of what one can reliably use for a time delay in state-space reconstruction.
While the criterion is linear, it may not be totally
misleading to use the first zero crossing of the autocorrelation function as a useful time lag. When the average
mutual information does have a first minimum, it is usually more or less the same order, in units of rs9 as the first
zero crossing of the autocorrelation, so one is not likely
to be terribly misled by this tactic.
Once a time delay has been agreed upon, the embedding dimension is the next order of business. As we have
seen, there are numerous options available, and we shall
not repeat them in detail. Our own experience is that it
is better to work with algorithms that are geometric rather than derivative from the data. Computing correlation
functions Cq(r) not only requires a large data set, it also
degrades rapidly when the data are contaminated. If one
wishes simply to know that the embedding dimension is
low, then geometric methods will suffice. If one wishes to
know whether to use dimension d or d 4-1, then
geometric methods will allow a way to start the selection,
and then after performing some signal separation as indicated below, perhaps fractal dimension estimators can
pin down the answer more precisely. In any case, robustness seems to come with methods that do not require precise determination of distances between points on the
strange attractor.
V. INVARIANTS OF THE DYNAMICS

8dU-y(k))

(g) = fddxg(x)p(x)

= -h%zWk)),

(40)

^ k= \

we see that
fddxg(F(x))p(x)

= ^-"2
^

*(y(*+l

k = i

= | [ g ( y W + D ) - g ( y ( l ) ) ] + (g> ,
(41)
and in the limit TV> oo, the averages of g(x) and g ( F ( x ) )
are the same. Since F(x) moves us along the orbit precisely one step, we see that, for long enough data sets, the
physicist's meaning of iV oo, all (g) are unchanged by
perturbations within the basin of attractionall those
points in state space which end up on an attractorof
the observations.
Quantities such as ( g ) can now be used to identify the
system that produced the observations. The (g ) are statistical quantities for a deterministic system and are the
only quantities that will be the same for any long observation of the same system. Different chaotic orbits will,
as we shall make quantitative below, differ essentially
everywhere in phase space. The issue, then, is to choose
a g (x) that holds some interest for physical questions.

(39)

Before proceeding any further, we need to discuss the


practical issues associated with numerical estimation of
the invariant distribution, Eq. (39). Expressing the measure as a sum of delta functions is supremely convenient
for defining other invariant quantities, as Eq. (40) demonstrates. But integrals of delta functions presuppose an
absolutely continuous, infinitely resolvable state space,
which is something most experimentalists are not familiar with.
Any measurement of a tangible variable necessarily imposes a partitioning of state space. The number of distinct, discrete states observable by any apparatus can be
no larger than the inverse of the measurement resolution.
Each element of this partition can be assigned a label,
which we shall denote as i (typically this label might be
some expression of the measurement value). In this
discretized space, the state-space integral
<g) = fddxg(x)p(x)

(42)

becomes

< * > = 2 *<*/)/><*/) >


1= 1

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

A. Density estimation

Classification of the physical system producing one's


observations is a standard problem for physicists. In an
earlier section we discussed general matters related to invariants of the dynamics. Now we give more detail.
The basic idea is that one wants to identify quantities
that are unchanged when initial conditions on an orbit
are altered or when, anywhere along the orbit, perturbations are encountered. As usual we assume that underlying the observations y(k) there is a dynamical rule
y(k -\-l) = (y(k)).
For purposes of defining invariants,
we need some notion of invariant distribution or invariant measure on the attractor. This issue is discussed at
some length in the earlier review article of Eckmann and
Ruelle (1985), and for us it is only necessary to agree that
only one of the many possible candidates for invariant
distribution seems to be stable under the influence of errors or noise. This is the natural density of points in the
phase space, which tells us where the orbit has been, and
whose integral over a volume of state space counts the
number of points within that volume. The definition of
this density is
PM = -^?L

in phase space of dimension d.


Now any function on phase space g ( x ) , when integrated with this density, is invariant under the dynamics
xF(x). Defining

(43)

Abarbanel et a/.: Analysis of observed chaotic data


where the probability of the zth element of the partition is
d

f d x p(x)
r
xt
d
2 fJx d xp(x)
J

, ,
nix;)
2,n(Xi)

with n(Xj) equal to the number of counts in partition element i. These discrete estimates of the partition probabilities in the state space provide our approximation for
the invariant measure. Equation (40) can be interpreted
as the sparsely populated version of Eq. (43), resulting
from taking so fine a partitioning of phase space that
each partition element contains only one point.
This type of partition-based estimator is simply a multidimensional histogram and, as such, does not provide
smooth distributions and can suffer from boundary
effects. One way of improving upon traditional histograms is to use kernel density estimators, which replace
the delta function in Eq. (39) with one's favorite integrable model for a localized, but nonsingular, distribution
function. Tight Gaussians are a popular choice for the
kernel, but scores of alternatives have been suggested on
various grounds of optimality criteria (Silverman, 1986).
Kernel density estimators illuminate an alternative to
the traditional fixed-volume methods for estimating
densities/probabilities. Instead of fixing a uniform grid
to partition the phase space and counting the number of
observations in each equal-sized partition element, it is
possible to base our estimates on the localized properties
of the given data set around each individual point.
Fixed-mass techniques perform a virtual partitioning of
the phase space with overlapping elements centered on
every point in the data set. The radius of each partition
element, rN (*), is increased until a prespecified number
NB of neighboring points have been enclosed. In this
way we accumulate a large number of overlapping partition elements, all containing the same number of points
but having different spatial sizes. The local density, or
probability, is then estimated as a function of the inverse
of the bin radius:
p(x)*rN

(x)~

dE

(45)

where dE is the embedding dimension of the reconstructed'phase space.


It will soon become clear that every time we can identify two broad classes of methods for implementing a
type of algorithm, the particular properties of any given
data set will dictate a hybrid technique for optimal performance. For example, if there is obvious clustering in
the data, it would make sense to have the different clusters lie in different partition elements.

B. Dimensions
Attractor dimension has been the most intensely studied invariant quantity for dynamical systems. Much of
the interest of the past decade was spawned by the realiRev. Mod. Phys., Vol. 65, No. 4, October 1993

1353

zation that attractors associated with chaotic dynamics


have a fractional dimension, in contrast to regular, or integrable, systems, which always have an integer dimension. Armed with this knowledge, a legion of researchers
embarked on the search for evidence of fractional dimensions in experimental time series, to demonstrate the existence of chaos in the real world.
Since there are already a large number of very good reviews of dimensions and their estimation (Farmer et al.>
1983; Paladin and Vulpiani, 1987; Theiler, 1990), we shall
not devote much space to a full discussion of all the
relevant issues. In particular, we refer the reader to
Farmer et al. (1983) for a more detailed discussion of the
basic theoretical concepts and Theiler (1990) for a
comprehensive review of many of the subtleties of estimation. Paladin and Valpiani (1987) provide a good
discussion of the inhomogeneity of points distributed on
an attractor, stating clearly how one determines quantitative measures of this inhomogeneity and how one looks
for effects in the physics of this phenomenon.
Previous sections described the basic mathematical
concept of the phase-space dimension as the (integer)
number of quantities that need to be specified to fully
identify the state of the system at any instant. In the case
of ordinary differential equations, or discrete mappings,
the phase-space dimension corresponds to the number of
equations defining the evolution of each component of
the state vectors. In many cases of physical interest, e.g.,
Navier-Stokes or any system of partial differential equations, this state-space dimension is infinite. However, in
dissipative dynamical systems it is frequently the case
that the system's asymptotic behavior relaxes onto a
small, invariant subset of the full state space. The variety
of dimensions about to be introduced describe the structure of these invariant subsets.
Beyond the simplest concept of dimension as the number of coordinates needed to specify a state is the geometrically related concept of how (hyper) volumes scale as a
function of a characteristic length parameter:
V^LD

(46)

Planar areas scale quadratically with the length of a side,


and volumes in real world space go as the cube of the side
length. Contingent upon some workable generalization
of the idea of volume, we can invert Eq. (46) to "define" a
dimension:
logL *
Early techniques for estimating the dimension used a
covering of the attractor to calculate V. Consider a partitioning of the dE -dimensional phase space with an esized grid. Count how many of the partition elements
contain at least one point from the sample data, and use
this value as the measure of V at resolution e. Then
refine the phase-space partition by decreasing e. Now
count how many of these smaller partition elements are
not empty. Continue this procedure until e has spanned

Abarbanel et a/.: Analysis of observed chaotic data

1354

a large range, hopefully at least several orders of magnitude. In theory, in the limit as e>0 the ratio in Eq. (47)
describes the space-filling properties of the point set being analyzed. It is obvious that the largest value of D calculable with this box-counting algorithm is dE. If the
embedding dimension is not large enough to unfold the
attractor's structure fully, we shall only be observing a
projection of the structure, and much information will be
obscured. Therefore the common practice is to repeat
the described box-counting calculations in a number of
increasing
embedding
dimensions.
For
lowerdimensional embeddings, the attractor's projection is expected to "fill" the space, resulting in an estimated fractal dimension equal to the embedding dimension. As the
embedding dimension increases through the minimum required for complete unfolding of the geometric structure,
the calculated fractal dimension saturates at the proper
value. Figure 19 illustrates this basic technique with data
from the Henon map.
According to our geometric intuition, Eq. (46) takes
"the amount" of an object to be equivalent to its volume.
But this is not the only measure for how much of something there is. Theiler has nicely phrased Eq. (46) in the
more general form
bulk c size'dimension

(48)

1. Generalized dimensions
Let us now return to our discussion of invariant quantities by considering the case in which bulk is measured
in terms of some moment of the invariant distribution.
In other words, take g(x)=pp(x)
in Eq. (40) and use
(g ) l / p to measure bulk. Putting everything together we
arrive at

r-*o log(size)
r-+o
r
~\l/p
= lim-log P 2 / ^
logr
r-*0

log2>*

= lim
r-+o q\

logr

logr
r

1 log 2 P " + 1

lim

logr

Dn

This definition of the generalized dimension Dq provides a whole spectrum of invariant quantities for
oo < q < oo.
We
are
already
familiar
with
Z> 0 = logA^ 0 /logr; this is just the capacity estimated by
the box-counting method.
Next we have Dx to consider. The q>\ limit results
in an indeterminate form, but one easy way to find the
limit is with L'HospitaFs rule:
_3_
1 dq

This now invites us to consider other measures of bulk.

log2>f

Dx = \im
or

j?=

(49)

2/^logA-

(50)

dq
:

^FsV

J.

/ .A^
//
^
/
f

lim
r-*0 T
i

FIG. 19. Box counting for Henon attractor in different embedding dimensions dE = 2,3 and at differing spatial resolutions.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

i = i

5>;loSA
(51)
= lim
;
r~>o logr
Note that, when pt l/N for all /, then Dl=DQ.
But
when the/?, are not all equal Dx <D0. (In fact, in general
Dt< Dj, if i>j.)
The difference between the two measures is explained
by how we use our state-space partition. For D0 the bulk
is how many partition elements are nonempty. An element containing one sample contributes as much bulk as
an element containing a million samples. Dx is calculated from a bulk in which each partition element's contribution is proportional to how many samples it contains.
So bulk in DQ is simply the volume, whereas for Dx bulk
is equivalent to the mass.
Because of the functional form of the numerator in Eq.
(51), Dx is commonly called the information
dimension.
That numerator expresses the amount of information
needed to specify a state to precision r, and the ratio
defining the dimension tells us how fast this required information grows as r is decreased.
Next we have D2, the so-called correlation dimension:

Abarbanel et al.: Analysis of observed chaotic data

-log 2 pf
D2 = \im

.
(52)
r^o
logr
We see that the numerator constitutes a two-point correlation function, measuring the probability of finding a
pair of random points within a given partition element,
just as the numerator in the definition of D{ measures the
probability of finding one point in a given element.
Consequently, we can try to estimate the numerical
value of that numerator by counting how many pairs of
points have a separation distance less than some value r.
In 1983 Grassberger and Procaccia (1983a, 1983b) suggested using

C2(r)=

fddxp(x)n(r,x)

= * ^ | ' - l y < ; > - y < / > l >

(53)

as a simple and computationally efficient means of estimating ^Ltpf.


With the introduction of the
Grassberger-Procaccia algorithm the floodgates were
opened for researchers looking for chaos in experimental
data. From fluid mechanics and solid-state physics, to
epidemiology and physiology, to meteorology and
economics, there was a staggering volume of papers publishing estimates of fractal dimensions.
2. Numerical estimation
Evaluation of the Dq has been the subject of numerous
inquiries (Smith, 1988; Ruelle, 1990; Theiler, 1990).
Much of the discussion centers on how many data are required to determine the dimension reliably. Generally
speaking, large quantities of data are necessary to achieve
accurate approximations for the density of points in the
different regions of the attractor, and a good signal-tonoise ratio is required to probe the fine structure of any
fractal sets in the attractor. Clearly, the required number
of points must scale as some function of the dimension
being estimated. The following are just a few opinions
that can be found in the literature:
# Bai-Lin H a o (1984) points out the most basic limit.
To examine a 30-dimensional attractor "the crudest
probe requires sampling two points in each direction and
that makes 2 3 0 109 points, a value not always reached in
the published calculations."
Theiler (1990) tells us N~D.
"Experience indicates
that should be of the order of 10, but experience also
indicates the need for more experience.*'
Leonard Smith (1988) claims that to keep errors
below 5 percent one must have N > A2M, where M is the
largest integer less than the set's dimension.
# On the other hand, Abraham et al. (1986) argue that
"Increasing N to the point where L/Nl/D
is less than a
characteristic noise length is not likely to provide any
more information on the attractor's structure."
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1355

Ruelle (1990) and Essex and Nerenberg (1991) argue


that if one estimates a dimension larger than 2Xlog 1 0 (A0
that one has only counted something related to the statistics of the sample size N. This would mean that, for estimating a dimension d, a data set of at least size \0d/1 is
required.
It is not uncommon to see attempts to overcome the
limitations imposed by small data sets by measuring the
system more frequently. However, as discussed earlier in
Sec. IV, this is not an effective tactic. The raw number of
points is not what matters; it is the number of trajectory
segments, how many different times any particular locale
of state space is revisited by an evolving trajectory, that
counts.
Identifying the scaling region for estimating the dimension is another significant challenge. Standard practice
calls for accumulating the data's statistics and then plotting logC(r) versus logr for a set of increasing embedding
dimensions. For embedding dimensions smaller than the
minimum required for complete unfolding of the attractor, the data will fill the entire accessible state space, and
the slope of the plot will equal the embedding dimension.
As the embedding dimension increases, the slope of the
logC(r) versus logr plot should saturate at a value equal
to the attractor's dimension. The dimension is defined as
the slope of this plot in the r >0 limit, but this region of
the plot is always dominated by noise and the effects of
discrete measurement channels. Therefore one hopes to
identify a scaling region at intermediate length scales,
where a constant slope allows reliable estimation of the
dimension by Eq. (47). Working with large quantities of
machine-precision computer-generated data, it is not
difficult to produce gorgeous plots with clearcut scaling
regions spanning many decades. But when dealing with a
limited quantity of moderately noisy data, the scaling region may not be identifiable. For example, one of the
characteristic signatures of oversampled data is the appearance of a "knee" in the correlation integral plot,
separating regions with different scaling behavior. The
knee is caused by enhanced numbers of near neighbors at
small scales due to the strong correlation between temporally successive points. We refer the reader to the review by Theiler (1990) for a more complete and detailed
explanation of the subtleties of dimension estimation,
along with a survey of proposed refinements for increasing the efficiency and reliability of the algorithms.
Motivated by the simple fact that many interesting
time series are extremely short or noisy, substantial effort
has gone into clarifying the statistical limits of the techniques (Smith, 1988; Theiler, 1990). At the other extreme
is the question of how large a dimension can be investigated with these techniques. Although many experienced investigators believe the data requirements limit
the applicability of these algorithms to systems of dimension 6 or less, several groups claim encouraging results
up to 20 dimensions with careful application of the
Grassberger-Procaccia algorithm.

Abarbanel et a/.: Analysis of observed chaotic data

1356

In addition to the difficulties in accumulating sufficient


statistics for the reliable estimation of dimensions, there
is the uncomfortable fact that one can still be tricked by
the algorithm's output. A controversial example was
pointed out by Osborne and Provenzale (1989), who
claimed that time series generated by colored stochastic
processes with randomized phase are typically identified
as finite dimensional by standard correlation integral
methods. In one sense, this can be explained as a consequence of the fact that the data are not stationary, since
the correlation length grows with the length of the time
series. Although Theiler (1991) identified the sources of
their paradoxical results and suggested several procedures for the avoidance of such anomalies, the fact
remains that the identification of systems with
\/fa
power spectra can be a frustrating task. In Sec. VIII we
shall see that filtered signals are also susceptible to misinterpretation by standard dimension-estimating algorithms.
More important to attend to is the use to which the Dq
are put for establishing that one has a low-dimensional
chaotic signal. If this is indeed one's interest, then using
the false-nearest-neighbor method or one of the variants
will do that with many fewer data and much less computation. If one wishes to see an invariant on the attractor
saturate, then choose one that is less computationally intensive. If one really wants the numerical value of Dq,
then caveat emptor.
With all this warning, we nonetheless quote some familiar D2, which come from graphs of good clean correlation functions as a function of log[r]. From these we
are able to read off for the Henon map D2 ~ 1.26, for the
Lorenz system D 2 2 . 0 6 , and for the Ikeda map
Z>2~l-8. For the hysteretic circuit of Pecora and Carroll we see the results in Fig. 20. In earlier sections we
have loosely referred to the dimension of the attractor as
dA and made no distinction between the value of D2 as
quoted here and the values of the box-counting D0 dimension and information dimension Dx which appear in
Hysteretic circuit: correlation integrals
7 500 points, nT=1, dE=1 ,...,8
8
1

10 6

514
102 I

10

10

'

10

-^

'

10
r

FIG. 20. Sequence of correlation integrals C2(r) for hysteretic


circuit data.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

more precise statements. In practice, these dimensions


are all numerically so similar that there is no advantage
in distinguishing them from a signal analysis point of
view. A clear discussion of how and when the various
fractal dimensions are similar is given by Beck (1990),
who relates this question to the singularity of the invariant distribution on the attractor.
C. A seguefrom geometry to dynamics
One frustrating aspect of dimension estimation is the
meager quantity of useful information gained for the
amount of effort expended. Fractal dimensions are a convenient label for categorizing systems, but they are not
very useful for practical applications such as modeling or
system control. The dimensions describe how the sample
of points along a system orbit tend to be distributed spatially, but there is no information about the dynamic,
temporally evolving, structure of the system.
An interesting connection between the fractal dimensions discussed in the previous section and the stability
properties of the system dynamics was conjectured by
Kaplan and Yorke in 1980.
Assume the state space can be decomposed at each
point into local stable and unstable subspaces, with a
basis set containing expanding and contracting directions. The rates at which these deformations occur in
each direction are called the Lyapunov numbers. With
the typical ordering LX>L2...
>LU>\>
Ls...
>Ld9
and the existence of an attractor requires n f = 1 ^j < * ^
dynamical system is said to be chaotic if at least one of
t h e L y > 1.
Consider a partitioning of the state space with ddimensional elements of side length e, i.e., the number of
partition elements required to cover a unit d-cube is e~d.
Along the lines of a box-counting algorithm, let us count
the number of partition elements needed to cover our attractor, and call this number N(e).
Now consider one of these elements of the partition
that covers the attractor as a fiducial volume, and let this
volume evolve according to the system dynamics. Each
axis of the fiducial volume will be scaled by a factor proportional to that direction's Lyapunov number. Directions with a Lyapunov number of magnitude less than 1
will contract, meaning one e width will still be needed to
cover these directions. Directions with a Lyapunov number greater than one will be expanded, thereby requiring
more e widths to cover the evolved fiducial volume in
these directions.
Let us use this information to estimate the number of
partition elements needed to cover the attractor if the
partition resolution is improved. Take the partition
spacing to be Ce, where C is some constant. How is
N(Ce) related to N(e)7 Following the same reasoning as
above, we expect that each direction associated with a
Lyapunov number less than C will contribute a factor of
1.0, i.e., leave N unchanged. But each direction associated with a Lyapunov number greater than C will contrib-

Abarbanel et a/.: Analysis of observed chaotic data

ute a factor of Lt /C to N:

1357

log

1=1

Dr=-

N(C)=N()Hmax

\\ max

LK + \

,1

logL K + \

(54)

,1

log
Recall the basic scaling relation that motivated our boxcounting algorithm:

K
i = l

V*N(e)~e-

(55)

So

maxLt

-n
AT, x .Vi
N(e)

N{Ce)

_ ^

(Ce) -D

--D

C~D

(56)

yielding yet another dimension measure:


d

Y[ max

log
Dr =

i = i

(57)

logC

Now what value should be used for C? Consider using


one of the Lyapunov numbers, say LK + l. Then

2*/

=K

i = l

^K + \

where the ki9 the logarithms of the Lt, are called the
Lyapunov exponents.
Which Lyapunov exponent should be picked for kkl
In the spirit of the definition of the capacity D0 we want
to choose the value that minimizes DL. Note that if all
the A,, are less than zero the attractor is a stable fixed
point and has dimension zero. If the only non-negative
exponents are zero, the attractor is a limit cycle. Multiple null exponents correspond to the number of incommensurate frequencies in a quasiperiodic system, which is
also the system's dimension. When at least one exponent
is positive, the last equation is to be minimized with
respect to kk, to provide the most efficient covering of
the invariant set.
To find the optimal kk consider the difference in DL
calculated with kk or kk + l:

2*/
&Dr = k'+l-

2*/

i = l

/=!

^k+2

** + l

i = i

^-k+2

^fc + 1

^k + \
k

= 1-

^k + \
^k+7

k
/=1

^+n

|X A:+2l

+ 2l~I^A: + ll
\i k+2\

Wt + li ^ + 2 1

i
x
and
therefore
Assuming
that
'k + u^k
UA'k+2 < 0,
|A,fc + 1 | < l^&+2l> w e s e e t n a t t n e n r s t factor is a positive
quantity less than one. The second factor is a negative
so long
as
quantity,
and DT
is decreasing,
2?-i*i>lA- fc + l l Therefore we take

2*/
DL=K~X

/=1
K+l

(58)

such that 2 f = j ^ > 0, and 2 f = V kt < 0.


An exact correspondence between DL> constructed
from measures of the stability properties of the dynamics,
and the dimensions defined in terms of moments of the
measure of the limit set is not obvious. The derivation of
the quantity is based upon a clever twist in the traditional
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

i = l

l^fc + ll
:

description of the capacity D0.


By constructing a
refinement of the phase-space partitioning that is custom
tailored to the way phase space is deformed by the dynamics, we automatically create an efficient covering for
the invariant set, and we know how the population of the
covering scales in the 6>0 limit. However, that efficient
covering is defined in terms of the Lyapunov exponents,
which are invariant quantities, defined as averages over
the invariant measure. Therefore it seems that DL might
equal D l.
But which geometric-based dimension is equal to DL is
not very relevant from the time series analyst's
viewpoint. Rarely does the real world bestow upon the
experimentalist a data set of good enough quality to distinguish among these various dimensions. The important
issue is which definition can motivate an efficient and

Abarbanel et a/.: Analysis of observed chaotic data

1358

stable numerical algorithm for dimension estimation.


One characteristic the Dq all share is that they become
unwieldy and demand too many data as the dimension of
the attractor approaches and exceeds ten. Being based
on the dynamics rather than the distribution of sample
points, the Lyapunov dimension allows us to investigate
much higher dimensions so long as we can get reliable estimates for the Lyapunov exponents.

have established by phase-space reconstruction or direct


observation an orbit y(k);k = 1,2, . . . ,N to which we
make a small change at " t i m e " unity: w( 1). The evolution of the dynamics will now be
y(fc + l) + w ( f c + l ) = F(y(fc) + w(Jfc)) ,
which will determine the w(k).
and stay small, then

(60)

If the w{k) start small

w ( f c + l ) = F ( y ( i k ) ) - y ( i k + l ) + DF(y(ik))-w(ife)+
D. Lyapunov spectrum
= DF(y(fc))-w(fc) .
Dimensions characterize the distribution of points in
the state space; Lyapunov exponents describe the action
of the dynamics denning the evolution of trajectories.
Dimensions only deal with the way measure is distributed
throughout the space, whereas Lyapunov exponents examine the structure of the time-ordered points making up
a trajectory.
To understand the significance of the spectrum of
Lyapunov exponents, consider the effects of the dynamics
on a small spherical fiducial hypervolume in the (possibly
reconstructed) phase space. Arbitrarily complicated dynamics, like those associated with chaotic systems, can
cause the fiducial element to evolve into extremely complex shapes. However, for small enough length scales
and short enough time scales the initial effect of the dynamics will be to distort the evolving spheroid into an ellipsoidal shape, with some directions being stretched and
others contracted. The primary, longest, axis of this ellipsoid will correspond to the most unstable direction of
the flow, and the asymptotic rate of expansion of this axis
is what is measured by the largest Lyapunov exponent.
More precisely, if the infinitesimal radius of the initial
fiducial volume is called r ( 0 ) , and the length of the ith
principal axis at time t is called //(f), then the /th
Lyapunov exponent can be defined as
^=Iim|log-^ .

(59)

By convention, the Lyapunov exponents are always ordered so that A^ > A,2 > A,3 .
Equivalently, the Lyapunov exponents can be seen to
measure the rate of growth of fiducial subspaces in the
phase space. kx measures how quickly linear distances
growtwo points initially separated by an infinitesimal
distance will, on average, have their separation grow as
ee l . The two largest principal axes define an area element, and the sum kx + k2 determines the rate at which
two-dimensional areas grow. In general, the behavior of
d-dimensional subspaces is described by the sum of the
first d exponents, X / = o ^jE. Global exponents
The spectrum of Lyapunov exponents is determined by
following the evolution of small perturbations to an orbit
by the linearized dynamics of the system. Suppose we
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

(61)

The evolution of the perturbation can be written as


w(L + l ) = D F ( y a ) ) - D F ( y ( L - 1 ) ) D F ( y ( l ) ) - w ( l )
= DFL(y(l))-w(l) ,

(62)
L

which defines our shorthand notation, DF (y( 1)), for the


composition of L Jacobian matrices D F ( x ) along the orbit y(k).
In 1968 Oseledec (1968) proved the important multiplicative ergodic theorem, which includes a demonstration
that the eigenvalues of the orthogonal matrix
DF L ( x) [ DF L ( x) ] T are such that the matrix
lim { D F L ( x ) - [ D F L ( x ) ] r j 1 / 2 L

(63)

L> oo

exists and has eigenvalues e x p ^ ^ e x p t A ^ ] , . . . >exp[X rf ]


for a <i-dimensional dynamical system which are independent of x for almost all x within the basin of attraction of
the attractor. The Xa are the global Lyapunov exponents. Their independence of where one starts within
the basin of attraction means that they are characteristic
of the dynamics, not the particular observed orbit. We
call them global because the limit L > oo means they are,
in a sense we shall make more precise, a property of the
global aspects of the attractor.
Oseledec also proved the existence of the eigendirections of DF L (y( 1)). These are linear-invariant manifolds
of the dynamics: points along these eigendirections stay
along those directions under action of the dynamics as
long as the perturbation stays small. Ruelle (1979) and
others extended this to the nonlinear manifolds and to
more complicated spaces. These eigendirections depend
on where one is on the attractor, but not on where one
starts in order to get there. That is, if we want to know
the eigendirections of D F L at some point y(L) and we begin at y ( l ) and take L steps, or begin at y(2) and take
only L 1 steps, the eigendirections will be the same as
long as L is large enough. In practice, on the order of ten
steps are usually required to establish the eigendirections
accurately.
The multiplicative ergodic theorem of Oseledec is a
statement about the eigenvalues of the dynamics in the
tangent space (Eckmann and Ruelle, 1985) to x>F(x)
and is motivated by a statement about linearized dynamics of the mapping. It is also a characterization via the
Xa of nonlinear properties of the system. Basically it is
an analysis of the behavior of the nonlinear system in the

Abarbanel et a/.: Analysis of observed chaotic data


neighborhood of an orbit on the strange attractor. By
following the tangent-space behavior (the linearized
behavior near an orbit) locally around the attractor, we
extract statements about the global nonlinear behavior of
the system. A compact geometrical object, such as a
strange attractor, which has positive Lyapunov exponents cannot arise in a globally linear system. The
stretching associated with unstable directions of
DF(y(A:)) locally and the folding associated with the dissipation are the required ingredients for the theorem.
If any of the ka are positive, then over the attractor
small perturbations will grow exponentially quickly.
Positive ka are the hallmark of chaotic behavior. If all
the ka are negative, then a perturbation will decrease to
zero exponentially quickly, and all orbits are stable. In a
dissipative system, the sum of the ka must be negative,
and the sum governs the rate at which volumes in phase
space shrink to zero. If the underlying dynamics is
governed by a differential equation, one of the ka will be
zero. This corresponds to a perturbation directly along
the vector field, and such a perturbation simply moves
one along the same orbit on which one started, so nothing happens in the long run. Indeed, one can tell, in principle, if the source is governed by differential equations
or by a finite time m a p by the presence or absence of a
zero global Lyapunov exponent.
A Hamiltonian system preserves phase-space volume
and respects the other invariants reflecting the symplectic symmetry (ArnoPd, 1978). This leads to two consequences: (i) the sum of all ka is zero, and (ii) the ka come
in pairs, which are equal and opposite. If ka appears in
the collection of Lyapunov exponents, so does ka. For
Hamiltonian dynamics there are two zero global
Lyapunov exponents, one associated with the conservation of energy, the second associated with the fact that
the evolution equations are differential.
An interesting result due to Pesin (1977) relates, under
fairly general circumstances, the sum of the positive exponents to the K S entropy: 3

h(X)= 2 K

(64)

xa>o

1359

comes from the fact that although each D F ( y ) has


eigenvalues more or less like exp[A, fl ], the composition of L Jacobians, D F L ( y ) , has eigenvalues approximately exp[A,jL], exp[A, 2 ], . . ,exp[kdL],
and since
^i > ^ 2 > ' ' ' >kd, as L becomes large, the matrix is terribly ill conditioned. The diagonalization of such a matrix poses serious numerical problems. Standard Q R
decomposition routines do not work as well as one would
like, but a recursive Q R decomposition due to Eckmann,
Kamphorst, Ruelle, and Ciliberto (Eckmann et aL, 1986)
does the j o b . 5 The problem in the direct Q R decomposition is keeping track of the orientation of the matrices
from step to step, so they propose to write each Jacobian
in its Q R decomposition as
DF(y(i))-Q(i ~ 1 ) = Q ( I ) - R ( I ) ,

(65)

where we begin with Q(0) as the identity matrix.


DF(y( 1)) is then decomposed as Q( 1 )-R( 1) as usual with
the Q R decomposition of matrices. Then we write
D F ( y ( 2 ) ) - Q ( l ) = Q(2)-R(2) ,

(66)

so we have for the product


DF(y(2))-DF(y(l)) = Q(2).R(2)-R(l) .

(67)

Continuing, we have for DF (y( 1))


DFL(y(l)) = Q a ) - n

= 1 R(fc)

(68)

The problem of directions within the product of matrices


has been handled step by step; at each step of this recursive procedure no matrix R(A:) is much larger than
exp[A, t ] and the condition number is more or less
exp[kx~kd],
which is quite reasonable for numerical accuracy.
If we have differential equations rather than mappings,
or more precisely, if we want to let the step in the map
approach zero, we can simultaneously solve the
differential equation and the variational equation for the
perturbation around the computed orbit. Then a recursive Q R procedure can be used for determining the
Lyapunov exponents. Greene and K i m (1986) use a recursive singular-value decomposition, and the idea of the
method is quite the same.

F. Ideal case: Known dynamics


G. Real world: Observations
If one knows the vector field, determination of the ka
is straightforward, though numerically it poses some
challenges (Greene and Kim, 1986). 4 The main challenge

See also the paper by D. Ruelle (1978), which gives a bound


on h (X) in rather general settings.
4
This paper also presents eigendirections associated with
Lyapunov exponents. These are not the same eigendirections
mentioned above for DF L , they are the eigendirections associated with DF (y( 1 ))-[DF L (y( 1 ))]T. The eigendirections of DF L
need not be orthogonal.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

The problem of determining the Lyapunov exponents


when only a scalar observation s{n) is made is another
challenge. As always, the first step is reconstruction of
the phase space by the embedding techniques described
in Sec. IV. Once we have our collection of embedded
vectors y( n) there are two general approaches to estimating the exponents. In one case an analytic approach is

A similar recursive method was shown to us independently by


E. N. Lorenz (private communication, 1990).

Abarbanel et a/.: Analysis of observed chaotic data

1360

adopted when a predictive model is available for supplying values for the Jacobians of the dynamical rules.
Lacking such detailed information about the dynamics,
one can still try to estimate the larger exponents by
tracking the evolution of small sample subspaces through
the tangent space.
1. Analytic approach
The problem is to reconstruct the phase space and then
estimate numerically the Jacobians DF(y(&)) in the
neighborhood of each orbit point. Then the determination of the Lyapunov exponents by use of a recursive Q R
or singular-value decomposition is much the same as before. T o estimate the partial derivatives in phase space
we evidently must use t h e information we have about
neighbors of each orbit point on the attractor. The idea
(Sano and Sawada, 1985; Eckmann et al, 1986; Brown
et ah, 1990, 1991) is to make local maps from all points
in t h e neighborhood of the point y(fc) to their mapped
image in the neighborhood of y ( f c + l ) . In the early
literature (Sano and Sawada, 1985; 6 Eckmann et al.,
1986) this map was taken to to be locally linear, but this
accurately gives only the largest exponent, since the numerical accuracy of the local Jacobians is not very good
and when one makes mistakes in each of the elements of
the composition of Jacobians that one is required to diagonalize to determine the ka, those errors are exponentially compounded by the ill conditioned nature of the problem (Chatterjee and Hadi, 1988). The problem is that the
burden of making the local map and the burden of being
an accurate Jacobian were both put on the dE X dE matrix in the local linear map. The solution is to allow for a
larger class of local neighborhood-to-neighborhood
maps, and then extract the Jacobian matrix from that
map.
This means that locally in state space, that is, near a
point y(n) on the attractor, one approximates the dynamics x>F(x) by

and quite accurate values for the full Lyapunov spectrum


are thus achieved. In the papers of Brown et al. (Bryant
et al., 1990; Brown et al., 1991), local polynomials were
used, and that was the method of Briggs (1990) as well.
Various basis functions have been used for this: radial
basis functions (Parlitz, 1992), sigmoidal basis functions
(called neural networks in popular parlance) (Elmer
et al, 1991; MeGaffrey et al, 1992), and probably other
bases as well.
Typical of the results one can achieve are the exponents shown in Table I I for the Lorenz attractor,
where it is known that k x = 1.51, k2 ~ 0. > a n c *
k3=22.5.
W e achieve high accuracy in establishing
these values from data on the x (n) observable only. T h e
table comes from making a local polynomial fit from
neighborhood to neighborhood in the phase space of the
Lorenz attractor, reconstructed from t h e observations
x(n).
From this local map numerical values of the Jacobian are read off and then the eigenvalues of products of
Jacobians are found by t h e recursive method indicated.
In the data used for x (n) the parameter values are a = 4,
b = 1 0 , and r = 4 5 . 9 2 . We do not use the equations of
the Lorenz model in determining these values for the ka,
but from them we can establish a check on the results because the rate of phase-space contraction in the Lorenz
model is exp[ (a-\-b + 1 ) ] , and by the general properties of the Lyapunov exponents this must also be equal to
exp[A,! + A,2 + A,3]. This check is satisfied very accurately
by the numerical results.
All these results are quite sensitive to the presence of
contamination by other signals. This is quite natural,
since the evaluation of D F ( y ) is quite sensitive t o distances in phase space, and inaccurate determination of
distances leads to inaccurate determination of D F . In
particular, achieving an accurate value of DF(y(/c)) is a
task in extrapolating from measurements at finite distances in phase space to values on the orbit, i.e., zero distance, and that is just where noise or contamination has
the most effect. The effect of noise is shown in Fig. 21,

.FU)= 2 cB(fcty*(x) ,

Lorenz 63 with noise

(69)

Lyapunov exponents

k= \

with the <f>k(x) a set of basis functions chosen for convenience or motivated by the data. The coefficients cn(k)
are then determined by a least-squares fit minimizing the
residuals for taking a set of neighbors of y( n) to a set of
neighbors of y(n + 1 ) . The numerical approximation to
the local Jacobian then arises by differentiating this approximate local map.
The burden of making the map and achieving a numerically accurate Jacobian are separated by this approach,

5.0

IB

0.0 4

-5-0 [

-io.o \

c
^

5*

This paper also introduces the local-phase-space method for


finding DF(y(n)), but uses only local linear maps for this purpose. The method is unable to determine all exponents, just the
largest exponent.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

'

-8.0

-g____^___~-Br

**__^__

/
K

-15.0 V

-25.0

-2o.o i
6

\
\

-I

'

-6.0

'

-4.0
log(noise)

-2.0

'

0.0

FIG. 2L Global Lyapunov exponents for Lorenz attractor in


the presence of noise.

Abarbanel et at.: Analysis of observed chaotic data

1361

TABLE II. Global Lyapunov exponents computed from x (t) data for the Lorenz attractor. The order
of the neighborhood-to-neighborhood polynomial map is varied. The correct values for these exponents are ^ ! = 1.50, A,2 = 0.0, and ^3= 22.5.
Global Lyapunov exponents for the Lorenz system
Order of
polynomial used
in local fit

Linear
Quadratic
Cubic
Quartic
Quintic

1.549
1.519
1.505
1.502
1.502

-0.094 70
-0.02647
-0.005 695
-0.002 847
-0.000 387

where the three Lyapunov exponents for the Lorenz system are evaluated from x(n) data for various levels of
uniform random numbers added to the data. The sensitivity is clearly seen here.
2. Trajectory tracing method
As an alternative to finding numerical values for the
local Jacobians of the dynamics along the orbit, one can
attempt to follow small orbit differences (Wolf et aL,
1985) and small areas defined by two orbits, and then
small subvolumes defined by three or more orbits to
sequentially extract values for the Lyapunov exponents.
These methods, while popular, almost always give accurate values for only the largest Lyapunov exponent k{
due to inevitable numerical errors. Nonetheless, we describe this method and alert the reader to its sensitivity.
Assume we have a large collection of experimental
points from a long time series. What space our data is in
is not important, whether it be the natural space of the
defining dynamics or an embedding by time delays or
whatever. We want to find points in our sample library
that have close neighbors from different temporal segments of the library. Call these points x(t) and y(t). We
then keep track of how they deviate under dynamical
evolution until they become separated by a distance
greater than some threshold a n d / o r another trajectory
segment passes closer to the presently monitored one.
When either of these conditions is met, the ratio of
1
n

\\y(n+t)-x(n+t)\\
\\y(t)-x{t)\\

(70)

is calculated and a new pair of neighboring points is observed. The largest Lyapunov exponent is estimated by
the log of the long-term average of the above quantity.
Strictly speaking, we want the asymptotic value as
n > 00. In practice, we are happy to be able to trace a
trajectory in this manner for several dozen characteristic
time steps.
If a large sample of observation points is available, we
may adopt a numerical trajectory-tracing technique
motivated by an alternative, but equivalent, definition of
the largest Lyapunov exponent:
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

-14.31
-20.26
-22.59
-22.63
-22.40

^ m a x = lim - l o g | | D F ^ u | [
n>00

(71)

rl

where u is a nearly randomly oriented unit vector perturbation in the tangent space. We say nearly because if u
should just happen to fall precisely along the stable manifold, it would remain within that invariant set and not relax onto the most unstable direction.
Our task is to find points in our sample library that are
very close together and watch how trajectories specified
by following the points separate. In locating the initial
neighboring points we must not consider points that are
from the same temporal segment of the library. Let us
denote the starting point of our reference orbit as
r ( 0 ) = y(/) and call its nearest spatial neighbor
Vi(0) = y(j). We require that i and j differ by some
minimum decorrelation length.
Now consider following the evolving trajectories
emanating from these two points. Initially the points are
separated by some distance || A 0 ||, and we want to see how
this distance grows under dynamical evolution. Therefore we continue to calculate a series of separation distances
||A fc || = ||r(fc)-i; 1 (fc)|| = | | y ( i + f c ) - y ( 7 + f c ) | |

(72)

until we find a value of Afc exceeding a present threshold


value. Typically this threshold will be some intermediate
distance between the smallest resolvable distances and
the overall size of the attractor as a whole. ||A 0 || is assumed to start off at the smallest length scale, but when
||A^|| has grown to some sizable fraction of the
attractor's extent we expect other nonlinear mechanisms
of enfoldment to dominate the further evolution of A.
Therefore we attempt to locate another point v2(0) in
our sample library close to r{k), lying as near as possible
to the line segment connecting r{k) and vx(k).
Things
can get tricky here as we try to trade off between nearness to r(k) and nearness to the separation segment from
r(k) to v(k). Locating v2(0) close to \r{k) vx{k)\ is
important, because in addition to being stretched by the
unstable components of the nonlinear dynamics, A is rotated into alignment with the unstable manifold. In order to get an accurate estimate of the expansion properties of the dynamics, we want to keep all future A's

Abarbanel et al.: Analysis of observed chaotic data

1362

aligned with the unstable direction.


When we reinitialize our neighboring trajectory with
v2(0), we store the final value of Ak || for sx along with its
terminating value of k =kax. Then we iterate the procedure, following r(k) further, with its new neighbor v2,
monitoring the separation:
&krx+k = \\r(krx+k)-v2(k)\\

(73)

We continue this procedure for as long as the data or


computing resources allow. Then we calculate the ratio

^0

^max + 1

fcS^

+l

and estimate
*,,1 = ^ - l o g A .
u max
M

(75)

To determine the second Lyapunov exponent we repeat the above tracing procedure, but monitor the evolution of area elements rather than separation lengths.
Consider the same starting points used in the above example, r(0) and v{(0). We want to augment this pair
with another neighboring point wx(0), which is close to
r(0) while being in a direction orthogonal to 8 0 , the separation between r ( 0 ) and 1^(0). Call the separation vector
between r(0) and w{(0) y0. Then 8 0 and y 0 define a fiducial area element, which we can try to follow under
dynamical evolution. We again follow r(k)> vx{k), and
now w{(k) also, at each point calculating the area
spanned by 8^ and yk.
However, following this two-dimensional element introduces further complications because, in general, yk
will tend to become aligned with 8^ along the unstable
direction. Therefore, in addition to reinitializing v and w
when the spanned area grows above a threshold value, we
must carefully reorient w to keep it as orthogonal to v as
the data resources permit.
Clearly the trajectory tracing method requires a substantial quantity of data to facilitate the aligned updating
procedure. If we are unable to locate nearby update
points in the required directions, our exponents will be
underestimated, since the maximal expansion rates will
not be observed. Although the basic idea easily generalizes to higher-dimensional subspace elements for estimating the rest of the exponents, the precision and quantity
of data required, especially in larger dimensions, makes
this technique impractical for anything beyond the largest couple of exponents.

3. Spurious exponents
One difficulty associated with working in reconstructed
phase spaces is that artifacts arising from inefficiencies of
representation inherent to the ad hoc embedding can be
annoying. A good example of such an annoyance is the
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

presence of spurious Lyapunov exponents when the dimension of the reconstructed space exceeds the dimension of the dynamics' natural space dN.
Analysis of a dynamical system in a d-dimensional
space necessarily produces d Lyapunov exponents. A n d
Takens' theorem tells us that the embedding space may
have to be expanded to d =2dN + l. So it is possible,
even in the ideal noise-free case, to have dN + 1 spurious
exponents in addition to the dN real ones.
Takens' theorem assures us that dN of the exponents
we calculate in a large enough embedding space will be
the same as those we could calculate from a complete set
of measurements in the system's natural space. But little
was known about the behavior of the spurious exponents
until recent work clarified the robustness of the numerical estimation techniques.
First clues were provided by Brown et al. (Bryant
et al., 1990; Brown et al., 1991) when they investigated
the use of higher-order polynomial interpolants for modeling the dynamics in the reconstructed phase space.
Identifying the directions associated with each computed
Lyapunov exponent, they demonstrated that very few
sample points could be found along the directions of the
spurious exponents. Essentially, the directions of the real
exponents were thickly populated, indicating there was
actual dynamical structure along these directions. T h e
directions of the artificial exponents showed no structure
beyond a thickness attributable to noise and numerical
error.
Recently another method h a s been introduced for
discriminating between real and spurious exponents.
Based on the idea that most of the deterministic dynamical systems of physical interest are invertible, we first
perform the typical analysis on the given time series, and
then we reverse the time series, relearn the reversed dynamics, and proceed to calculate exponents for this backward data.
Under time reversal, stable directions become unstable,
and vice versa, no real exponents are expected to change
sign. On the other hand, the spurious exponents, being
artifacts of the geometry imposed by our ad hoc embedding, will not, in general, behave so nicely. In fact, experience seems to show that the artificial exponents vary
strongly under time reversal, making identification easy.
H. Local exponents from known dynamics
and from observations
Global Lyapunov exponents tell us how sensitive the
dynamics is on the average over the attractor. A positive
Lyapunov exponent assures us that orbits will be exponentially sensitive to initial conditions or to perturbations or to roundoff errors. N o information is gained
from them about the local behavior on an attractor. Just
as the inhomogeneity of the attractor leads to interesting
characterization of powers of the number density, so it is
natural to ask about the local behavior of instabilities in
various parts of state space.

Abarbanel et a/.: Analysis of observed chaotic data


A question of more substantial physical interest is this:
if we make a perturbation to an orbit near the phasespace point y, how does this perturbation grow or shrink
in a finite number of time steps? This question is directly
relevant to real predictability in the system, since the local Lyapunov exponents vary significantly over the attractor, so there may be some parts of phase space where
prediction is much more possible than in other parts.
For example, if we are investigating the ability to predict
weather in a coupled ocean-atmospheric model, we are
much more interested in making predictions six weeks
ahead of a specified state-space location than 10 6 weeks
(20000 years) ahead. Global Lyapunov exponents may
suffice for the latter, but local Lyapunov exponents are
required for the former. Local Lyapunov exponents are
defined directly from the Oseledec matrix (Abarbanel
et a/., 1991),
L

OSL(y,L) = D F ( y ) . [ D F ( y ) ] ,

lim ka(y9L)

= ka

the cra(p,L)-+0
for L - > o o and p>2, again by the
Oseledec multiplicative ergodic theorem.
The behavior of the ka(y9L) can also be captured in a
probability density for values of each exponent, for a
fixed L. The probability density for the ath exponent to
lie in the interval [u,u + du] is after L steps along the orbit,

( W > L > = 4 F 2 ; Saa(y(fc),L)- -u)

(77)

the global Lyapunov exponents. It is the variation in


both phase-space location y and length of time L that is
of interest.
First of all, the ka{y,L) are strongly dependent on the
particular orbit that we examine, just as are all functions
that we evaluate on a chaotic orbit. So we must ask
about averages over orbits with the invariant density.
For this purpose we define an average local Lyapunov exponent ka(L)9

ka(L)= f ddy p(y)ka(y9L)

(81)

and the moments above follow in the usual way.


The behavior of the ka(L) in model systems is quite
striking. For the familiar test chaotic systems such as the
Henon map, the Ikeda map, the Lorenz attractor, and
others, we have for large L
-'"''

ca
T

(76)

which has eigenvalues that behave approximately as


exp[2Lka(y,L)].
The ka(y9L) are the local Lyapunov
exponents. From the multiplicative ergodic theorem of
Oseledec we know

1363

ca
a

(82)

JU

where ca and c'a are constants, va is a scaling exponent


found numerically to be in the range 0 . 6 < v f l 5-0.85 for
many systems, and the final term proportional to L " 1
comes from the lack of complete independence of the local Lyapunov exponents on the choice of coordinate system. A demonstration of this is found in Abarbanel
et ah (1992)]. This also shows the independence of the
global exponents from the coordinate system when
L > oo.
All moments vanish for large L and do so at the rate
Ora(pyL)tt

+ -

(83)

where c'J and c'a" are constants, ga is a scaling index in


approximately the same numerical range as the va> and
the last term comes from the coordinate system dependence.
A typical example can be seen in Fig. 22, where the

(78)
Lorenz 63; Local Lyapunov exponents

For large L9 since ka(y9L) becomes independent of L9


within the basin of attraction of the attractor defined by
the observations we have
lim kJL)

= kn

from equations. 10 000 locations

(79)

The variations around the mean value are also of some


importance, and we define the moments
[aa(p9L)Y=

j ddy

p{y)[ka(y9L)-ka{L)f

= T7 2
N~

[ka(y{k)9L)-ka(L)f

(80)

Two of the moments are trivial: aa(09L)=l


and
aa(\,L) = 0. The moments for p > 2 tell us how the local
exponents vary around the attractor, though they clearly
average this information over the whole attractor. The
direct local information resides in the ka(y9L).
Each of
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

FIG. 22. Average local Lyapunov exponents for Lorenz attractor calculated from the Lorenz equations (11) as a function of
the composition length L. Averages are over 10000 initial conditions.

Abarbanel et a/.: Analysis of observed chaotic data

1364

Lorenz 63; Local Lyapunov exponents

DFL(y)-[DFL(y)]r=Q1(2L)R1(2L)R1(2L - 1 ) R ^ l ) .

from equations - standard deviations. 10 000 locations

(84)
If QX(2L) were the identity matrix, the eigenvalues of
OSL(y,L) would all lie in the "upper triangular" part of
the Q R decomposition. In general, QX(2L) is not the
identity. We follow the construction in Stoer and Burlisch (1980) and make a similarity transformation to
Qf(2L)Q 1 (2L)R 1 (2L)R 1 (2L - 1 ) R 1 (1)Q 1 (2L)
= R 1 (2L)R 1 (2L - 1 ) R 1 (1)Q 1 (2L) .

(85)

Now we repeat the Q R decomposition,


DFL(y)-[DFL(y)]r
FIG. 23. Standard deviations from the average for local
Lyapunov exponents calculated from the Lorenz equations (11)
as a function of the composition length L. Averages are over
10000 initial conditions.

= Q 2 (2L)R 2 (2Z,)R 2 (2L - 1 ) R 2 ( l ) ,


which has the same eigenvalues as the
OSL(y,L). We continue this K times to reach
QK(2L)KK(2L)KK(2L

three ka(L) for the Lorenz attractor are shown, and in


Fig. 23, where the three aa(2>L) for the same system are
shown. We also present in Fig. 24 the distribution of exponents Px(u,L) for the Lorentz attractor for L = 2 and
L=50.
The scaling indices va and %a are characteristic of the
dynamics and can be used, just as are the Dq or the Kq or
the ka, for classifying the source of signals. It is important to evaluate them from data. The method is really a
straightforward extension of the techniques outlined
above for evaluating the global Lyapunov exponents
from datawith one small exception.
As before we compute the recursive Q R decomposition
of

~l)

- KK(1)

(86)

original
(87)

which still has the same eigenvalues as OSL(y,L). As K


increases, QK(2L) converges rapidly to the identity matrix (Stoer and Burlisch, 1980).
When QK(2L) is the identity to desired accuracy, the
final form of OSL(y,L) is upper triangular to desired accuracy, and one can read off the local Lyapunov exponents, ka(y,L),
from the diagonal elements of the
R(fc)*s:

K(y>LY-

l
2Z,T,s

2L

(88)

2iog[^(;U,

j= \

since the eigenvalues of a product of upper


matrices are the product of the eigenvalues of
dual matrices. The rapid rate of convergence
to the identity means only a few steps, K 3

triangular
the indiviof QK(2L)
or so, are

Lorenz 63 - data; d_E=4; d_L=3

Lorenz 63; local Lyapunov exponents

Local Lyapunov exponents; 75 000 points


20.0

Probability distribution

x.

10.0

-10.0

-20.0

-30.0 r

11
7.0
9.0
Lyapunov exponent

13.0

15.0

FIG. 24. The distribution of the largest Lyapunov exponent for


the Lorenz model for composition lengths L = 2 and L = 50.
Each distribution is normalized to unity. Note that when L = 2,
this largest exponent can be negative.
Rev. Mod. Phys.i Vol. 65, No. 4, October 1993

13

15

iog2L

FIG. 25. Average local Lyapunov exponents calculated directly


from the time series x(t) generated by the Lorenz equations
(11). 75 000 data points are used. The embedding dimension is
dE = 4 , and the order of local polynomials is taken to be dL = 3;
1500 initial conditions are used.

Abarbanel et a/.: Analysis of observed chaotic data


Ikeda Map, d_E=4; d_L=2
Local Lyapunov exponents; 60 000 points
2.0 p

1.5 i

1.0Jv
0.5 I
<<

0.0
-0.5 [
-1.0 h

-^^"^

- 1 . 5 [
-2.0 '

'

'

'

'

'

"-

11

13

'

'

15

logL

FIG. 26. Average local Lyapunov exponents for data from the
Ikeda map; 60000 points are used. dE = 2, dL=2; 1500 initial
conditions.
ever needed to assure that QK (2L) differs from the identity matrix by a part in 10~ 5 or 10~ 6 .
In Figs. 25 and 26 we show the local Lyapunov exponents determined from observations on the x(n) data
from the Lorenz equations and from Re[z(/i)] from the
Ikeda map. In the former case we used dE=4 and used
local polynomials of dimension dL=3
for the
neighborhood-to-neighborhood maps, while in the latter,
dE = 4 and dL=2.
We shall discuss below how we know
dL2 for the Ikeda map.
I. Topological invariants
An ideal strange attractor, which never has a coincidence between stable and unstable directions, has embedded within it a dense set of unstable periodic orbits
(Devany and Niteeki, 1979; Auerbach et ah, 1987; Cvitanovic, 1988; Melvin and Tufillaro, 1991). Such a hyperbolic attractor is rarely, if ever, found in physical settings, but the approximation of real strange attractors by
this idealization can be quite rewarding. Indeed, in many
examples of strange attractors, the Henon and Ikeda
maps used as examples throughout this article seem to
possess numerically many of the properties of the "ideal"
strange attractor. The topological properties of the linkings among these unstable periodic orbits characterizes
classes of dynamics, and these topological classifications
remain true even when the parameters of the vector field
governing the dynamics are changed. So when one goes
from stable periodic behavior to quasiperiodic behavior
to chaos, the classification by integers or ratios of integers characteristic of topological invariants remains the
same. This is in direct contrast to the various invariant
properties we have discussed heretofore. It is clear that,
if one were able to extract these topological invariants
from experimental observations, they could be used to exclude incorrect dynamics and to point to classes of dynamics that could be further investigated, say, by cornRev. Mod. Phys., Vol. 65, No. 4, October 1993

1365

puting dimensions or Lyapunov exponents. In that sense


such geometric quantities can be a critical tool in selecting possible vector fields that could describe the observations.
If the dynamics can be embedded in R3, then the
periodic orbits are closed curves, which can be classified
by the way they are knotted and linked with each other
(Birman and Williams, 1983a, 1983b). This is the case
that has been studied in detail in the literature (Mindlin
et ah, 1990; Flepp et ah, 1991; Mindlin et ah, 1991;
Tufillaro et ah, 1991; Papoff et ah, 1992) and applied to
experimental data. Whether one will be able, in a practical way, to extend the existing literature to more than
three dimensions (Mindlin et ah, 1991) remains to be
seen, but the applications to physical systems that happen to fall into this category have already proven quite
interesting.
The first task, clearly, is to identify within a chaotic
time series the unstable periodic orbits. To do this one
does not require an embedding (Mindlin et ah, 1991),
though to proceed further with topological classification
an embedding is required. Basically one seeks points in
the time series, embedded or not, which come within a
specified distance of one another after a fixed elapsed
timenamely, the period (Auerbach et ah, 1987; Mindlin et ah, 1990) of the unstable orbit. To succeed in this
one must be slightly lucky in that the Lyapunov number
associated with this unstable orbit had better be close to
unity or in one period the orbit will have so departed
from the unstable periodic structure in state space that
one will never be able to identify the unstable periodic orbit. If some orbits can be found by this procedure, then
one can proceed with the classification.
Using time-delay embedding may not always work well
for the purpose of classifying attractors by their topology, for in three dimensions there may be self intersections
of the flow; i.e., dE = 3 may be too low for time-delay
embedding. If the local dimension of the dynamics is
three, one may be able to follow the orbits in d = 4 or
higher, but this has not yet been done. In much of the reported work, embeddings other than time-delay embeddings are utilized, though there is no systematic procedure available. One useful embedding is to take the
scalar data s(n) and form

k= \

y2(n)=s(n)
y3(n)=s(n)

(89)
s(n

1) .

This is a perfectly fine embedding, and in the study of experiments on chemical reactions (Mindlin et ah, 1991)
and laser physics (Papoff et ah, 1992) it has served to unfold the unstable periodic orbits in three dimensions.
One may not always be able to accomplish this, but when
it is possible, tools are then available for further analysis.
In particular, by analyzing the linking of low-order

1366

Abarbanel et a/.: Analysis of observed chaotic data

periodic orbits and the torsionthe angle (in units of TT)


through which the flow twists in one period in the neighborhood of an unstable periodic orbitof such loworder orbits, one is able to establish a template or
classification scheme which fully characterizes all unstable periodic orbits and other topological properties of the
attractor (Mindlin et aL, 1991). Once the template has
been established one may verify it by examining the linkage of higher-order unstable periodic orbits, if they are
available. It is in the template that one finds possible deviation from the rigorous mathematics (Birman and Williams, 1983a, 1983b) available for hyperbolic strange attractors. In the template one may find periodic orbits
without counterpart in the actual flow. The gain for the
analysis of experimental data is such that this is a small
price t o pay. Indeed, as is often t h e case, precise
mathematics not quite applicable to a physical setting
may serve as a warning rather than forbidding attempted
application.
It appears (Gilmore, 1992) that this procedure is rather
robust against contamination, since noise first destroys
one's ability to identify high-order unstable periodic orbits. These, happily, are not needed in establishing the
template. In the presence of noise at the level of nearly
60% of the signal in one case (Mindlin et aL, 1991), the
low-order unstable periodic orbits were still visible. If
one could perform some sort of signal separation, even as
crude as local averaging on the time series, higher-order
noise levels could perhaps be handled.
This sequence of steps, from identifying the unstable
periodic orbits within a strange attractor to creating a
template, has been carried out in the analysis of a laser
physics experiment (PapofF et aL, 1992) and a chemical
reaction experiment (Mindlin et aL, 1991).
In the case of the N M R laser studied by Flepp et aL
(1991), it was established that a term in addition to the
conventional Bloch-KirchhofF equations is required by
the properties of the unstable periodic orbits in the data.
This is a powerful result and underlines the importance
of topological tools based on unstable periodic orbits
when they are available. Indeed, we would say that if
one is able to extend the existing analysis methodology to
dimensions greater than three, this topological approach
should always be among the first tools one applies to any
chaotic data set. It will then serve as a critical filter for
proposed model equations, regardless of whatever ability
one has t o evaluate invariants depending on distances
such as dimensions and Lyapunov exponents. The utility
of this approach and its robustness in the presence of
contamination serves as another item in our list of techniques that rest on geometry and may prove increasingly
important when real, noisy data are confronted. There is
no substitute, eventually, for knowing Lyapunov exponents of a system if prediction or control is one's goal,
but real data may require sharper instruments for its
study before these dynamical quantities can be reliably
extracted. The topology of the unstable periodic orbits
may well provide that instrument.

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

VI. NONLINEAR MODEL BUILDING: PREDICTION


IN CHAOS
The tasks of establishing the phase space for observations and of classifying the source of the measurements
can be enough in themselves for many purposes. Moving
on to the next step, building models and predicting future
behavior of the system, is the problem that holds the most
interest for physicists. In some sense the problem of prediction is both the simplest and t h e hardest problem
among all those we discuss in this review. The basic idea
is quite straightforward: since we have information on
the temporal evolution of orbits y(k) and these orbits lie
on a compact attractor in phase space, each orbit has
near it a whole neighborhood of points in phase space
which also evolve under the dynamics to new points. W e
can combine this knowledge of the evolution of whole
neighborhoods of phase space t o enhance our ability to
predict in time by building local or global maps
with parameters a: y>F(y,a), which evolve each
y(k)+y(k + 1 ) . Using t h e information about how
neighbors evolve, we utilize phase-space information to
construct the map, and then we can use the map either to
extend the evolution of the last points in our observations
forward in time or to interpolate any new phase-space
point near the attractor forward (or backward, if the map
is invertible) in time.
All that said, the problem reduces t o determining the
parameters a in t h e map, given a class of functional
forms which we choose in one way or another. As discussed by Rissanen (1989), there is no algorithmic way to
determine t h e functional form. This is actually good
news or all physics would be reduced to fitting data in an
algorithmic fashion without regard to the interpretation
of the data.
Suppose, from guessing or some physical reasoning, we
have chosen the functional form of our map. If we are
working with local dynamics, then local polynomials or
other natural basis functions would be a good choice,
since any global form can be approximated arbitrarily
well by the collection of such local maps, if the dynamics
is smooth enough. Choosing t h e parameters now requires some form of cost function or metric which measures the quality of the fit to the data and establishes how
well we do in matching y( k + 1 ) as observed with
F ( y ( k ) , a ) . Call this the local deterministic error, eD(k),
D(k) = y(k+l)-F(y(k),z)

(90)

The cost function for this error we shall call W(e). If the
map F ( y , a ) we are constructing is local, then for each
neighbor of y(k), yir)(k)r = 1,2, . . . ,NB,
e{\k) = y(r,k +\)-F(y{r)(kU)

y{r,k + 1 ) is the point in phase space to which the neighbor yir)(k) evolves; it is not necessarily the rth nearest
neighbor of y( k + 1 ) . For a least-squares metric the local
neighborhood-to-neighborhood cost function would be

Abarbanel et al.\ Analysis of observed chaotic data


NB

W(e,k)=

a, then

NB

\e(\k)\2/^

[y(k)-(y(k))]2

(91)
W(e)=

and the parameters determined by minimizing


W(,k)
will depend on the " t i m e " associated with y(k): a(/c).
The normalization is arbitrary, but the one shown uses
the size of the attractor to scale the deterministic error.
If the fit is to be a global least-squares determination of
;
I
/ ( y ( r , f c - M ) , F ( y u U ) , a ) ) = 2 P[y(r,k

with respect to the a(fc) at each k.


With one exception, to be noted shortly, this general
outline is followed by all of the authors who have written
about the subject of creating maps and then predicting
the evolution of new points in the state space of the observed system. The exception is a representation of the
mapping function taking x*F(x) in an expansion in
orthonormal functions (f>a(x)
M

(94)

a=\

where the functions are chosen to be orthonormal with


respect to the observed invariant density of the data
1

(95)

as
f ddx p(x)(/)a(x)<l)b(x) = bab .

(96)

The form of p(x) allows one to determine the expansion


coefficients c(a) without any fitting (Giona et al., 1991;
Brown, 1992), as we shall see below.
All other mapping functions, local or global, are of the
form of Eq. (96) with a structure for the functions (f>a(x)
that is preassigned and not dependent on the data from
the observed system.
In what follows we give a short survey of the methods
developed for the prediction in chaos. There are many
different methods, and we shall follow the traditional
classification of them as local and global models. By
definition, local models vary from point to point (or from
neighborhood to neighborhood) in the phase space. Global models are constructed once and for all in the whole
phase space (at least, the part occupied by the data
stream). It should be noted, however, that the dividing
line is becoming harder and harder to define. Models using radial basis functions or neutral nets carry the
features of both. They are usually used as global functional forms, but they clearly demonstrate localized
behavior. In the next subsections we start with the classic local and global models in order to explain the main
ideas and then discuss the newer techniques. First we
take up the topic of choosing the appropriate dimension
for our models.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

2
k= \

ki>(fc)l2/2

[y(k)-(y(k))]2

(92)

r= l

and so forth; An interesting option for modeling is to require the a to be selected by maximizing the average mutual information between y( r, k -f 1) over the neighborhood and the F(y ( r ) (/c),a); i.e., maximize

+ l),F(y<rU),a)]log2

r= \

FJx) = 2 ca(a) , ( * ) ,

1367

P[y(r,k+l),F(y{r\kU)]
P[y(r,fc+l)]P[F(y(r)(*),a)]

(93)

A. The dimension of models


Before discussing model making, it is useful to consider how to choose the appropriate dimension of the models when we have observed only scalar data s(n). When
we reconstruct phase space using time delays, we, at best,
arrive at a necessary dimension dE for unfolding the system attractor. As the example of the Ikeda m a p shows
quite clearly, dE may be larger than the local dimension
of the dynamics, dL. For the Ikeda map, dE~4, while
the local dimension, that is, the dimension of the actual
dynamics, is dL=2.
Recall that this difference comes
from the choice of coordinate system and has nothing to
do with the dynamical content of the observations.
While we could, in the Ikeda map again, make fourdimensional models without any special theoretical
penalty, making four-dimensional local models would require unnecessary work and unwanted susceptibility to
"noise" or other contamination of various dimensions
higher than dE.
A method we can use for choosing the dimension of
models is to examine the local average Lyapunov exponents ka(L). If we compute these exponents in dimension dE, we, of course, have dE of them. dEdL are totally geometric and have nothing to do with the dynamics required to describe the observations. If we compute
the ka(L), reading the dE dimensional data forward in
time, and then do precisely the same, reading the data
backward in time, the true exponents of the system will
change sign (Eckmann and Ruelle, 1985; Bryant et al.,
1990; Brown et a/., 1991; Parlitz, 1992; Abarbanel and
Sushchik, 1993). The false dEdL exponents will not
change sign and will be clearly identified. If one performs this calculation keeping dE fixed, so distances between points on the attractor are properly evaluated, but
changing the dimension of the local model of the
neighborhood-to-neighborhood dynamics from dL = dE
to dL~dE \ to dL=dE29
etc. until there are no
remaining false local Lyapunov exponents, then the
correct value of dL will be that for which the false exponents are first absent. At that local dimension, if it is
desired, one may return to the computation of the local
exponents using more data and higher-accuracy data to
evaluate the local Lyapunov exponents with as much pre-

Abarbanel et al.: Analysis of observed chaotic data

1368

quality of this prediction can be improved in different


ways. The first thing one can do is to take a collection of
near neighbors of the point y(k) and take an averaged
value of their images (Pikovsky, 1986) as the prediction.
Further improvement of the same idea is to make a
least-squares fit of a local linear map of y(k) into y(k + 1 )
using NB > dL near neighbors for a given metric || ||,

Ikeda map, d_E=4; d_L=4; 60 000 points


Local Lyapunov exponents, forward and (-) backward

W= X \\y(ryk + \)-&{r)y{r\k)-h{r)f

(97)

r= l

10

12

14

logL

FIG. 27. Local Lyapunov exponents calculated from data generated by the Ikeda map, dE = 4, dL =4. Forward exponents are
marked by open symbols, minus backward exponents by solid
symbols. Clearly, there are two true exponents and two spurious ones.
cision as desired. The advantage to using local Lyapunov
exponents is simply that one has more points of direct
comparison between forward exponents and negative
backward exponents than just the comparison of global
exponents. Further, if the data come from a differential
equation, local exponents are able to identify true and
false versions of the zero exponent, which as a global exponent will not change sign under time reversal,
In Fig. 27 are shown the average local Lyapunov exponents for the Ikeda map with dE=4, which is larger
than dL=2 but is the dimension suggested by the global
false-nearest-neighbor algorithm. Forward and minus
backward ka(L) are displayed. Evidently only two exponents satisfy the condition of equality of forward and
minus backward average local Lyapunov exponents. Decreasing dL from four to three and then to two, we observe (but do not display) the same phenomenon: only
two exponents satisfy the required time-reversal condition.
B. Local modeling
We now assume that our data are embedded in an appropriate phase space, and we have determined the dimension of the model. The problem now is to reconstruct the deterministic rule underlying the data.
We start our discussion with the simplest and earliest
nonlinear method of local forecasting, which was suggested by E. Lorenz (1969). Let us propose to predict the
value of y(k -f 1) knowing a long time series of y(y) for
j<k.
In the "method of analogs" we find the nearest
neighbor to the current value of y(k) say, y(m) and then
assume that the y ( m + 1 ) is the predicted value for
y(k + 1 ) . This is pure persistence, and is not much of a
model. In the classification by Farmer and Sidorowich
(1988) it is called a. first-order approximation. Clearly, the
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

This represents a second-order approximation.


The sum
in Eq. (97) can be weighted in order to provide a larger
contribution from close points. More sophisticated local
prediction schemes (Farmer and Sidorowich, 1987) may
represent the map in the form of a higher-order polynomial whose coefficients must be fit using near neighbors.
In this case one can expect better results for more longterm forecasts. However, the use of high-order polynomials requires many free parameters to be determined,
namely, (m -\-dL)\/{m\dL\)^d,
where dL is the local
dimension and m is the degree of the polynomial. It
makes these models less practical for local forecasting if
the embedding dimension a n d / o r the order of the polynomial are too high.
If one wishes to predict the value of y (k -\-K), i.e., K
steps ahead, one can use a direct method, which finds the
coefficients of the polynomial map K{y) by direct
minimization of
N

{K)

= 2 l|y(>*>^ +fc)-F*(y (r) (fc))|| 2

(98)

r= \

One particularly useful property of local polynomial


forecasts is that we can estimate how the quality of the
forecasts scales as a function of prediction interval K,
size of training set N, dimension of at tract or dA, and
largest Lyapunov exponent A,j.
Assume we accumulate a data set yn =y (xn ) from observing a continuously evolving system. In other words,
we know y varies continuously, but we can only make
measurements at the xn. We might try to model approximately the intervening values of the system variable y by
interpolating a fit to the measured data. Let
F(xn)
denote the true dynamics driving the system's evolution,
and consider its Taylor expansion:
F(xn +b)=

F(xn )+DF(xn

)b + \D2F{xn

+ f D 3 F U ) 8 3 + .

)8 2
(99)

If we construct a local polynomial model of degree m,


our leading-order expectation for the forecast error is the
S m + 1 term in Eq. (99). Recall one of our basic definitions
of dimension, and rearrange it to express the distance
variable as
L=Vl/D.

(100)

If we make the admittedly faulty assumption that the


points are evenly distributed through the state space,

Abarbanel et a/.: Analysis of observed chaotic data

method. For an appropriate range of base forecasting


time steps, jF5Tstep, the prediction error scales differently,

then V =pN, which implies


-j7B=Nl/D~b

(101)

This expression for the average distance between neighboring points can be used to replace 8 in Eq. (99).
Next there are the derivative factors in each term of
Eq. (99). As we have seen in Sec. V, the Lyapunov exponents can be viewed as measuring the average derivative of the dynamics over the attractor. Discrepancies
between our forecasts and the actual points on the evolving trajectory would grow much in the same way that initially nearby points would separate under the dynamics,
and this expansion rate is measured by the largest
Lyapunov exponent:
(DF(x))^exp(X1K)

(102)

Although nontrivial to justify (Farmer and Sidorowich,


1988), the assumption that higher-order derivatives also
average in a similar manner,
(DnF(x))<xexp(nkxK)

(103)

provides a useful model that agrees well with numerical


results from ideal calculations:
<Erms)cciv"(m

+ 1)/

^exp[(m+l)^1^] .

(104)

Having a limited number of data points, we must balance


two opposing factors. Higher-order polynomials potentially promise higher accuracy, but also require more
points in the neighborhood (and, therefore, larger neighborhoods). That in turn makes the local mapping more
complex.
Frequently it is desirable to attempt long-term forecasts. One straightforward approach is to use the already
described approximation techniques to construct a
predictive model, FK{x), specifically for the long-term interval:
FK{xn)^x(n+K)

1369

c7V"(m

+ 1)/

^exp[V^step]?

(106)

and implies higher accuracy due to the absence of the


factor ( m -f-1) in the second exponent.
Experience indicates that the scaling behavior of forecasting errors strongly depends on some of the modeling
parameters. If the time series is amenable to the sort of
analysis we are prescribing, there will be a range of
values for ^ s t e p where Eqs. (104) and (106) hold. This
range typically centers on the optimal value for the time
delay used in phase-space reconstruction (Sec. IV).
If data are collected by sampling a continuous signal at
a very high rate, forecasts can be made over intervals
shorter than the delay time used in phase-space reconstruction, with performance limited by the signal-to-noise
ratio of the data. However, using the shortest possible
composition time in an iterative forecasting scheme is not
a wise strategy, since the large number of forecasts that
need to be concatenated compound the errors at each
step and can cause the recursive technique to perform
worse than the direct methods.
Different local prediction methods were tested by
Farmer and Sidorowich (1988) and Casdagli (1989) for a
variety of low-dimensional chaotic systems. In Fig. 28
we show the comparison of different orders of direct and
iterative approximations for the Ikeda map. As expected, for first-order (constant) approximations there is no
difference between direct and iterative scalings. For
linear and quadratic predictors the iterative procedure is
clearly superior, in accordance with the scaling expectations. Figure 29 illustrates the difference between direct
and iterated forecast performance of local linear models
for hysteretic circuit data of Pecora and Carroll.

(105)

In this case we construct a model that predicts over the


full time interval in a single step. Any implementation of
the supervised learning algorithm would select a sample
of domain points from the neighborhood of the point of
interest, and the range sample would be all those neighbors' post-images the full time K later.
An alternative approach is to split the interval K into n
subintervals of length KstQp=K/n
and concatenate n
short-term forecasts to build up the full prediction. The
original domain sample remains the same, but the range
sample is now found, on average, to be much more closely packed, since the prediction interval is much shorter.
The initial forecast produces a new state, which is used as
the basis for selecting a new domain sample, which in
turn supplies a new range sample for the next predictive
model fit, and so on.
This iterative, or recursive, forecasting technique can
demonstrate substantial advantages over the direct
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

FIG. 28. Prediction error as a function of prediction time for


the Ikeda map for different prediction techniques.

Abarbanel et al.: Analysis of observed chaotic data

1370

Hysteretic Circuit - 60000 6D points

The vector field F ( x ) , which evolves data points


y(k + 1 ) = F(y(/c)), is approximated in M t h order as
M

F M ( x ) = 2 c(fl)0fl(x) ,

(109)

a= l

and the coefficients c(a) are determined via


fddxF(x)(f>a(x)p(x)

c(a)=
=

^ 1 F<y(*)tyfl(y(*)>
iV

k= I

^ T F X y(k + l)(f>a(y(k)) .
0

30

60
forecast interval

90

FIG. 29. Direct and iterative linear forecast error for hysteretic
circuit data.

C. Global modeling
The collection of local polynomial (or other basis function) maps form some global model. The shortcomings
of such a global model are its discontinuities and extremely large number of adjustable parameters. The
latter is clearly a penalty for high accuracy. At the same
time, it would be nice to have a relatively simple continuous model describing the whole collection of data. A
number of solely global models has been invented which
present a closed functional representation of the dynamics in the whole phase space (or, at least, on the whole attractor).
Each method uses some expansion of the dynamical
vector field F ( x ) in a set of basis functions in Rd. The
first such global method that comes into mind is to use
polynomials again. Their advantage in local modeling
where least-squares fitting works well, is now reduced by
the extremely large number of data points and the need
to use rather high-order polynomials. There is an attractive approach to finding a polynomial representation of a
global m a p . This measure-based functional
reconstruction (Giona et al., 1991; Brown, 1992) uses orthogonal
polynomials whose weights are determined by the invariant density on the attractor.
The method eliminates the problem of multiparameter
optimization. Finding the coefficients of the polynomials
and the coefficients of the function F( y) requires only the
computation of moments of data points in phase space.
It works as follows:
We introduce polynomials (f>a(x) on R d which are orthogonal with respect to the natural invariant density on
the attractor,
fddxp(x)<f>a(x)(l>b(x)

= 8ab ,

(107)

and which are determined by a conventional GramSchmidt procedure starting from


#i(x)=l .
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

(110)

120

(108)

This demonstrates the power of the method directly.


Once we have determined the orthogonal polynomials
(f>a(x) from the data, which we can do before computations are initiated, the evaluation of the vector field at
any order we wish is quite easy: only sums over powers
of the data with themselves are required.
Furthermore, the form of the sums involved allows one
to establish the vector field from a given set of data and
adaptively improve it as new data are measured. The
best aspect of the method, however, may be robustness
against contamination of the data (Brown, 1992). There
is no least-squares parameter search involved, so no distances in state space need be evaluated. The geometric
nature of the method does not rely on accurately determining distances and is thus not so sensitive to "noise"
which spoils such distance evaluations.
It is possible that the two general forms of global modeling we have touched on can be further improved by
working with rational polynomials rather than simple
Taylor series. Rational polynomials have a greater radius of convergence, and thus one may expect them to be
more useful when the data set is small and perhaps even
allow the extrapolation of the model from the immediate
vicinity of the measured attractor out into other regions
of phase space.

D. In between local and global modeling


As we already pointed out, the present direction of
nonlinear modeling combines features of local and global
models. Consider, for example, the method of radial
basis functions (Powell, 1981, 1985, 1987), which, as Casdagli (1989) notes, "is a global interpolation technique
with good localization properties." In this method a local predictor F(y) is sought in the form
N

F ( y ) = 2 ^<*>(||y-yJD,

(in)

where 4> is some smooth function in R 1 and ||-|| is the


Euclidean norm. The coefficients kn are chosen to satisfy, as usual,
F(y(fc)) = y(fc + 1) .

(112)

Abarbanel et al.: Analysis of observed chaotic data


Depending on the number of points Nc used for reconstruction, this method can be considered as local (small
Nc N) or global (Nc iV).
One can choose different 0 ( | | x | | ) as the radial basis
function. For example, <f>(r) = ( r 2 + c 2 ) _ / ? works well for
/? > 1 and /J^O. Moreover, if one adds a sum of polynomials to the sum of radial basis functions, then even increasing functions <I>(r) provide good localization properties. However, for a large number of points Nc this
method is computationally as expensive as the usual
least-squares fit. Numerical tests carried out by Casdagli
(1989) show that for a small number of data points radial
basis predictors do a better job than polynomial models,
whereas for a larger quantity of data (iV> 104) the local
polynomial models seem to be superior.
A modification of the radial-basis-function method is
the so-called kernel density estimation (Silverman, 1986).
This method allows one to estimate a smooth probability
distribution from discrete data points. Each point is associated with its kernel [a smooth function
K(\\yyf-||)],
which typically decays with distance but sometimes can
even increase. Then one can compute a probability distribution

/(y)=2*(||y-y,ID
i

or a conditional probability distribution

Pc(ylz)=2*(||y-y, + ill)*(||z-y,ll)i

Kernel density estimation can then be used for conditional forecasting by the rule
y(k + l)=fdxpc(x\y(k))x

(113)

Kernel density estimation usually provides the same accuracy as the first-order local predictors. It has the advantage of being quite stable even with noisy data (Casdagli et al.y 1991).
Again, computing the conditional probability distribution, one can impose weights in order to attach more
value to the points close to the starting point of prediction (both in time and in phase space). In fact, this leads
to a class of models that are hybrids of local and global
ones. Moreover, it allows one to construct a model that
not only possesses good predicting properties but also
preserves important invariants of the dynamics. The prediction model by Abarbanel, Brown, and Kadtke (1989)
belongs to this class. It parametrizes the mapping in the
form

1371

||y(fc + l ) -

C(X,a) =

y(* + l)g(y,y(*);a) ,

2 lly(*)||
k = l

(115)

where Xn, n = 1, . . . , L is a set of weights attached to the


sequence of points y(k n + 1 ) all of which are to be
mapped into y(k + 1 ) by maps F"(y). It is clear that then
F(y(fe),a) will be close to y{k + 1 ) , as it provides an excellent interpolation function in phase space and in time.
The free parameters a in the kernel function g and the
specific choice of the cost function allow one to predict
forward accurately a given number L > 1 of time steps
and satisfy additional constraints imposed by the
dynamicssignificant Lyapunov exponents and moments of the invariant density distribution, as determined
by the data, will be reproduced in this model. The
preservation of Lyapunov exponents does not follow automatically from the accuracy of the prediction. Indeed,
it is shown by Abarbanel et al. (1989) that models able to
predict with great accuracy can have all negative
Lyapunov exponents, even when the data are chaotic.
At first sight very different ideas are exploited in neural networks when they are used as nonlinear models for
prediction (Cowan and Sharp, 1988). Neural networks
are a kind of model utilizing interconnected elementary
units (neurons) with a specific architecture. After a training (or learning) procedure to establish the interconnections among the units, we have just a nonlinear model of
our usual sort. In fact, this is just a nonlinear functional
model composed of sums of sigmoid functions (Lapedes
and Farber, 1987) instead of sums of radial basis functions or polynomials as in the previous methods. There
are many different modifications of neural nets. Consider, for example, a typical variant of feed-forward nets. A
standard one includes four levels of units (Fig. 30): input
units, output units, and two hidden levels. Each unit collects data from a previous level or levels in the form of a
weighted sum, transforms this sum, and produces input
values for the next layer. Thus each neuron performs the
operation
X?ut=g

[ 2 TyXf

+ dt 1

(116)

where Ttj are weights, 6f are thresholds, and g(x) is a


nonlinear function typically of a sigmoidal form such as
(117)

(114)

k= \

where the function g(y,y(A:);a) is the analog of the kernel function, that is, it is near 1 for y y(k) and vanishes
rapidly for nonzero ||y y(&)||. a is a set of parameters.
The cost function to be minimized is chosen as
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

,
2

g(x) = [i+tanh(x>] .
F(y,a)= 2

XF"(y(fc-+l),a)||2

For the purpose of forecasting in dL -dimensional phase


space, the input level should consist of dL neurons, and
we feed them by the dL delayed values of the time series
{y( k), y( k 1), . . . , y( k dL + 1) ]. The predicted value
y(k +p) appears at the output level after nonlinear transformation at the two hidden levels:

Abarbanel et ai.\ Analysis of observed chaotic data

1372

-Z^**^--J**~^**z=^::^_

(X^O

OUTPUT UNIT

O O O O O^^XP

cQo o o o o o o o ox)H,Ds
J

INPUT UNITS

FIG. 30. Sketch of a feed-forward neural network.

y(k+p)
= F ( y ( * ) , y ( * - 1 ) , . . . , y(k-dL

+ l),{Tij}f{0i})

,
(118)

and the learning procedure consist of minimizing the error,

(genomes) of conditions (genes), and after each generation


some fraction of the worst sets of genes is discarded and
some mutations are produced on the rest of the conditions, The important feature of genetic algorithms is that
they allow an effective search in a very high-dimensional
phase space. Conditions learned after many generations
determine the part of the phase space (or orbit of a
dynamical system) for which prediction is possible. In
this regard we can see genetic algorithms as a global,
adaptive search method, which allows searches in both
parameter space and a predetermined, possibly large
space of functional representations of t h e dynamics
x - ^ F ( x ) . While there is really only limited development
of this approach to nonlinear forecasting at this time, it
clearly holds promise as an effective procedure for highdimensional and complex systems.
VII. SIGNAL SEPARATION "NOISE" REDUCTION

E=Z

\\y*Jk+p)-y(k+p)\\

(119)

where Ns is the number of data in a training set. D e pending on what sets of data are used for the training,
the neural net eventually can be considered as a local or
as a global model. In the first case only Nc i V near
neighbors are used for feeding the net, while in t h e
second one uses all the data. The fitted parameters are
here the weights {Ttj} and the thresholds {0 ( }. Unfortunately, this learning appears to be a kind of nonlinear
least-squares problem, which takes much longer than
linear fitting to run on regular computers. Nevertheless,
neural networks present great opportunities for parallel
implementation, and in the future that may become the
most useful technique for forecasting.
Neutral nets are quite useful when we have a simple
dynamics, and we can hope that a net with an a priori architecture will work well. Then the only thing to do is to
train this net, i.e., t o adjust the weights of connections
and thresholds. Adjustment of the architecture itself is a
secondary (and very often omitted) procedure for standard neural network modeling. The approach taken by
genetic learning algorithms (Goldberg, 1989) represent a
quite opposite point of view: This starts with finding an
architecture based on the data set. Sorting out different
architectures for a complex dynamics in a highdimensional phase space is a very difficult task. To accomplish this, genetic algorithms utilize Darwinian evolution expressed in terms of variables (data), mathematical conditions (architecture), and quality of fit (criterion
for sorting out the better architecture). Specifically, if we
have a set of pairs (xp,yp) where each xp is mapped into
yp in some dL -dimensional space, then we try to learn the
set of conditions (genes) imposed on {xp} under which
the quality of fit associated with {yp } is extremal. In particular, we can seek conditions determining the subspaces
of the whole phase space for which the distribution of
{yp} is sharpest (Packard, 1990). The search for the appropriate set of conditions is performed by a kind of natural selection. We start with a number of random sets
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Now we address some of the problems connected with


and some of the solutions to the very first task indicated
in Table I, namely, given observations contaminated by
other sources, how do we clean up the signal of interest so
we can perform an analysis for Lyapunov exponents, dimensions, model building, etc.? I n linear analysis the
problem concerns the extraction of sharp, narrowband
linear signals from broadband "noise." This is best done
in the Fourier domain, but that is not the working space
of the nonlinear analyst. Instead signals need to be
separated from one another directly in the time domain.
To succeed in signal separation, we need to characterize
one or all of the superposed signals in some fashion that
allows us to differentiate them. This, of course, is what
we do in linear problems as well. If the observed signal
s (n) is a sum of the signal we want, call it sx(n), and other signals s2(n),s3(n)..
.,
s(n)=s{(n)+s2(n)+

-"

(120)

then we must identify some distinguishing characteristic


of sx(n) which either the individual st(n), i> 1, do not
possess or perhaps the sum does not possess.
The most natural of these distinguishing properties is
that s! (n) or its reconstructed phase-space version,
yl(n) = [sl(n),sl(n

+TX),

. . . 9sl(n+T1(dl-l))]

,
(121)

satisfies a dynamical rule: y{(k + l) = Fl(yl(k))


from any dynamical rules associated with s2(n),
There are three cases we can identify:

different
... .

We know the dynamics


y\+F\(y\).
m We have observed some clean signal yR(k) from the
chaotic system, and we can use the statistics of the chaotic
signal to distinguish it.
% We know nothing about a clean signal from the dynamics or about the dynamics itself. This is the "blind"
case.

Abarbanel et a/.: Analysis of observed chaotic data


It is quite helpful as we discuss work on each of these
cases to recall what it is we are attempting to achieve by
the signal separation. Just for discussion, let us take the
observation
to be the
sum
of two
signals:
s(n)=sl(n)-\~s2(n).
If we are in the first case indicated
above, we want to know sl(n) with full knowledge of
yx>F1(y1).
The information we use about the signal
s^n) is that it satisfies this dynamics in the reconstructed
phase space, while s2(n) and its reconstruction does not.
This is not really enough, since any initial condition yx( 1)
iterated through the dynamics satisfies the dynamics by
definition. However, it will have nothing, in detail, to do
with the desired sequence s^DyS^D,
. . . which enters
the observations. This is because of the intrinsic instabilities in chaotic systems, so two different initial conditions
diverge exponentially rapidly from each other as they
move along the attractor. To the dynamics, then, we
must add some cost function or quality function or accuracy function, which tells us that we are extracting from
the observations s(n) the particular orbit that lies
"closest" to the particular sequence sl(l),sl(2),
...
which was observed in contaminated form. If we are not
interested in that particular sequence but only in properties of the dynamics, we need do nothing with the observations, since we have the dynamics to begin with and
can iterate any initial condition to find out what we want.
Similarly in the second case, we shall be interested in
the particular sequence that was contaminated or masked
during observation. In this case, too, since we have a
clean reference orbit of the system, any general or statistical question about the system is best answered by that
reference orbit rather than by attempting to extract from
contaminated data some less useful approximate orbit.
In the third instance, we are trying to learn both the
dynamics and the signal at the same time and with little a
priori knowledge. Here it may be enough to learn the dynamics by unraveling a deterministic part of the observation in whatever embedding dimension we work in. That
is, suppose the observations are composed of a chaotic
signal sx(k)y which we can capture in a five-dimensional
embedding space combined with a two-hundreddimensional contamination s2(k). If we work in five dimensions, then the second component will look nondeterministic, since with respect to that signal there will be so
many false neighbors that the direction in which the orbit moves at essentially all points of the low-dimensional
phase space will look random. Thus the distinguishing
characteristic of the first signal will be its relatively high
degree of determinism compared to the second.

1373

vations y0(k) in dx dimensions, and we wish to extract


from those measurements the best estimate yE(k) of the
particular sequence yx(k) which entered the observations.
The
error
we
make
in
this
estimation
y\(k) yE(k) = A(k) we call the absolute error. It is
some function of eA that we wish to minimize. We know
the dynamics, so we can work with the deterministic error D(k) = yE(k + 1) F1(yE(k)),
which is the amount
by which our estimate fails to satisfy the dynamics.
A natural starting point is to ask that the square of the
absolute error \eA{k)\2
be minimized subject to
D(k) = 0; that is, we seek to be close to the true orbit
yx(k) constraining corrections to our observations by the
requirement that all estimated orbits satisfy the known
dynamics as accurately as possible. Using Lagrange multipliers z(k), this means we want to minimize

\yi(k)-yE(k)\2

| i

N-l

+ X z(k)-[yE(k

-hD-F^y^k))]

(122)

with respect to the yE(m) and the z(m).


The variational problem established by this is
yE(k+l)

= F1(yE(k))

(123)
yl(k)-yE(k)

= z(k-\)-z(k)-T>Fl(yE(k))

which we must solve for the Lagrange multipliers and the


estimated orbit yE(k).
The Lagrange multipliers are not
of special interest, and the critical issue is estimating the
original orbit. If we wished, we could attempt to solve
the entire problem by linearizing in the neighborhood of
the observed orbit yo(k) and using some singular-value
method for diagonalizing the N XN matrix, where N is
the number of observation points (Farmer and
Sidorowich, 1991). Instead we seek to satisfy the dynamics recursively by defining a sequence of estimates
yE(k,p)by
yE(k,p

+ l)^yE(k,p)

+ A(k,p)

(124)

where at the zeroth step we shall choose yE(k,0) = yo(k),


since that is all we know from observations. We shall try
to let the increments A(k9p) always be "small," so we can
work to first order in them. If the Mk,p) start small and
remain small, we can approximate the solution to the dynamics as follows:
A(fc+l, j p) + y (fc + l,/?) = F 1 (y^(fc, j p) + A(fc,p)) ,
(125)

A. Knowing the dynamics: manifold decomposition


Mk+hp)
In this section we assume that the dynamics of a signal
are given to us: y 2 >F 1 (y 1 ) in J t -dimensional space, We
either observe a contaminated d x -dimensional signal or,
from a scalar measurement s(k)=sl(k)-i-s2(k),
we
reconstruct the d x -dimensional space. We call the obserRev. Mod. Phys., Vol. 65, No. 4, October 1993

= Fl(yE(k,p)

&(k,p))-yE(k+l,p)

DF 1 (y^(A:,/?))-A(^, j p) + 6 i) (fc,^) ,
where eD(kyp) = Fx(yE(k,p))
yE(k +1,/?) is the deterministic error, namely, the amount by which the estimate
at the pth step fails to satisfy the dynamics.

1374

Abarbanel et a/.: Analysis of observed chaotic data

We satisfy this linear mapping for A(k,p) by using the


known value for yE(k,p).
To find the first correction to
the starting estimate yE(ky0) = yo(k),
we iterate the
mapping for A(k,0) and then add the result to yE(k,0) to
arrive at yE(k9l) = yE(k90) + Mk,0).
The formal solution for A(kyp) can be written down in terms of A(l,p).
This approach suffers from numerical instability lying in
the positive eigenvalues of the composition of the Jacobian matrices DFj(y^(/c,/?)). These, of course, are the
same instabilities that lead to positive Lyapunov exponents and chaos itself.
To deal with these numerical instabilities, we introduce
an observation due to Hammel (1990): since we know the
map x>Fj(x), we can identify the linear stable and unstable invariant manifolds throughout the phase space
occupied by the attractor. These linear manifolds lie
along the eigendirections of the composition of Jacobian
matrices entering the recursive determination of an estimated orbit yE(k).
If we decompose the linear problem
Eq. (125) along the linear stable and unstable manifolds
and then iterate the resulting maps forward along the
stable directions and backward along the unstable directions, we shall have a numerically stable algorithm in the
sense that a small A(l,p) along a stable direction will
remain small as we move forward in time. Similarly, a
small A(N,p) at the final point will remain small as we
iterate backward in time.
Each iteration of the linear map for the A(k,p) will fail
to satisfy the condition A(k,p) = 0, which would be the
completely deterministic orbit, because the movement toward A(k,p) = 0 exponentially rapidly along both stable
and unstable directions is bumped about by the deterministic error at each stage. If that deterministic error
grows too large, then we can expect the linearized attempt to find the deterministic orbit closest to the observations to fail. This deterministic error is a measure at
each iteration of the signal-to-noise level in the estimate
relative to the original signal. If this level is too large,
the driving of the linear system by the deterministic error
will move the A(k,p) away from zero and keep them
there.
The main mode of failure of this procedure, which at
its heart is a Newton-Raphson method for solving the
nonlinear map, arises when the stable and unstable manifolds are nearly parallel. In practice they are never precisely parallel because of numerical roundoff, but if they
become so close to parallel that the forward-backward
procedure cannot correctly distinguish between stable
and unstable directions, then there will be "leakage"
from one manifold to another and this error will be
magnified exponentially rapidly until the basic stability of
the manifold decomposition iteration can catch up with
this error.
In Fig. 31 we show the result of iterating this procedure for 15 times when uniform noise is added to the
Henon map at the level of 2 . 1 % of the rms size of the
map amplitude. One can see the regions where the stable
and unstable manifolds become almost parallel, called
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Absolute Error in Cleaned Henon Map


15 Passes; 1250 Points; Original Noise in [-0.015,0.015]

40

140

240

L_^

340

440

540 640 740


Timestep in Data

840

940

1040 11.40

FIG. 31. Absolute error in use of the manifold decomposition


method for cleaning Henon map data contaminated with uniform noise at a level of 2.4% of the rms amplitude of the chaos.
Fifteen passes through the algorithm were used. Note the large
excursions due to homoclinic tangencies.

homoclinic tangencies in a precise description of the process, and then the orbit recovers. What we show in Fig.
31 is the absolute error \\yE(k,p) yi(fc)|| as a function of
k. A plot of D(k) shows no peaks and simply displays
noise at the level of machine precision.
The "glitches" caused by the homoclinic tangencies
can be cured in at least two ways (Abarbanel et al.9 1993)
that we are aware of. First, one can go back to the full
least-squares problem stated above and in a region
around the tangency do a full diagonalization of the matrices that enter. One can know where these tangencies
will occur because the Jacobians, which we know analytically since we know F ^ x ) , are evaluated at the present
best estimate yE{k,p).
The alternate solution is to scale
the step A(k,p) we take in updating the estimated trajectory by a small factor in all regions where a homoclinic
tangency might occur. One can do this automatically by
putting in a scale factor proportional to the angle between stable and unstable directions or by doing it uniformly across the orbit. The penalty one pays is a larger
number of iterations to achieve the same absolute error,
but the glitches are avoided.
In either of these cases one arrives at an estimate of the
true orbit y\(k) which is accurate to the roundoff error of
one's machine. So, if one has an orbit of a nonlinear
dynamical system that in itself is of some importance, but
it is observed to be contaminated by another signal, one
can extract that orbit extremely accurately if the dynamics are known. The limitations on the method come from
the relative amplitudes of the contamination and the
chaotic signal. In practice, when the contamination is
10% or so of the chaos, the method works unreliably.
The other important application of the method would be
using the chaotic signal as a mask for some signal of interest which has no relation to the dynamics itself. Then

Abarbanel et a/.: Analysis of observed chaotic data


one is able to estimate the chaotic signal contained in the
observations extremely accurately and by subtraction expose the signal of interest. Since this application would
be most attractive when the amplitude of the signal of interest is much less than the masking chaos, the method
should perform quite well.
B. Knowing a signal: probabilistic cleaning
Now we discuss the case in which we are ignorant of
the exact dynamics, but we are provided with a clean signal from the system of interest, which was measured at
some earlier time. Clearly, we can use any of the techniques from Sec. VI to estimate the dynamics, and then
use the resulting models in the algorithms discussed just
above. But now we focus on what can be done by using
the reference signal yR(k) to establish only the statistical
properties of the strange attractor. Later one receives
another signal from the same system yc(k) along with
some contamination z{k). T h e observed signal is

1375

s(/c)= : y c (fe) + z(fc). We have no a priori knowledge of


the contamination z(k) except that is is independent of
the clean signal yc(k).
Except for coming from the same
dynamics, y^ and y c are totally uncorrelated because of
the chaos itself.
The probabilistic cleaning procedure (Marteau and
Abarbanel, 1991) seeks a maximum conditional probability that, having observed the sequence s( 1), . . . , s(N), we
shall find the real sequence to be y c ( 1), . . . , yc(N). T o
determine the maximum over the estimated values of the
yc we write this conditional probability in the form

w(y.cM}>=

d26)

and note that only the entries in the numerator enter in


the maximization over estimated clean orbits.
The probability of the sequence of clean observations
can be estimated using the fact that yc(k + 1 ) = F(y c (A:))
so it is first-order Markov. This allows us to write

P(yc(l),yc(2), . . . ,yc(m)) = P(yc(m)|yc(m -l))P(yc(l),yc(2), . . . ,yc(m - 1 ) )


and use kernel density estimates of the Markov transition probability P(yc(m)\yc(m
2
P(yc(m)\yc(m-l))cc^l^

K(yc(m)-yR(j))K(yc(m
_
2

(127)

1)) via

-\)-yR(j'))
(128)

*(yc(m-l)-y*U))

n= l

The conditional probability P ({s] | {y c ]) is estimated by assuming that the contamination z is independent of the y c , so
m-\

P(s(j);

y = l, m\yc(k); fc = l , m ) = n

Pz(yc(k)s(k))

k = l

=P(s(j);j = l,m-l\yc(k);k=l9m-l)Pz(yc(m)s(m))

with Pz(x) the probability distribution of the contaminants.


These forms for the probabilities lead t o a recursive
procedure for searching phase space in the vicinity of the
observations s. One searches for adjustments to the observations s which maximize the conditional probability
^ ( ( y d I l s P - The procedure is performed recursively
along the observed orbit, with adjustments made at the
end of each pass through the sequence. Then the observations are replaced by the estimated orbit, the size of the
search space is shrunk down toward the orbit by a constant scale factor, and the procedure is repeated. This
search procedure is similar t o a standard maximum a
posteriori (Van Trees, 1968) technique with two alterations: (1) the probabilities required in the search are not
assumed beforehand, b u t rather are estimated from
knowledge of the reference orbit {y^(fc.))., and (2) in the
search procedure the initial volume in state space, in
which a search is made for adjustments to the observations, is systematically shrunk.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

(129)

The nonlinear dynamics enters through the reference


orbit and the use of statistical quantities of a deterministic system. The method is cruder than the manifold
decomposition algorithm, but it also requires much less
knowledge of the dynamical system. It works when the
signal-to-noise ratio is as low as 0 dB for extracting an initial chaotic signal contaminated by iid (independently
and identically distributed) noise. For a signal of interest
buried in chaos, it has proven effective when the signalto-chaos ratio is as low as 20 dB, and even smaller signals will work.
In Fig. 32 we show first a sample of the Lorenz attractor contaminated by uniform noise with a signal (Lorenz
chaos)-to-noise ratio of 0.4 dB, and then in Fig. 33 we
show the signal recovered by the probabilistic cleaning
method after 35 passes through the algorithm. T h e
signal-to-noise ratio in the final signal, relative t o the
known input clean signal, is 12.6 dB for a gain against
noise of 12.2 dB. When the initial signal-to-noise ratio is
higher, gains of about 30 to 40 dB with this method are

Abarbanel et al.: Analysis of observed chaotic data

1376

Lorenz Attractor
Uniform Noise in [-12.57, 12.57]

-40.0

Pulse of Amplitude 0.076

in Henon Data

Clean Pulse (Solid); Cleaned Pulse (Solid + Stars)

-30.0

FIG. 32. Noisy Lorenz attractor data. Contamination is uni


form noise in the interval [ 12.57,12.57]; this gives a signal
to-noise ratio of 0.4 dB.
Lorenz Attractor; Cleaned Data
Uniform Noise in [-12.57. 12.57]

-30.0

30.0

-20.0

40.0

FIG. 33. The cleaned version of the data in Fig. 32. Thirty-five
passes of the probabilistic cleaning algorithm are used. The
final signal-to-noise ratio, comparing with the known clean
data, is 12.6 dB.

Henon Data (Solid); Henon Data with Pulse (Solid +

Stars)

Pulse neor Step 75; Amplitude 0.076

FIG. 35. Pulse recovered from contamination by data from the


Henon map. Signal-to-noise ratio is now +20.8 dB.
easily achieved.
In Fig. 34 we show a pulse added to the Henon m a p
with an initial pulse-to-chaos ratio of 20 dB. This
pulse was then extracted from the Henon chaos to give
the result shown in Fig. 35, which has a signal-to-noise
ratio of 20.8 dB relative to the known initial pulse. This
method can clearly extract small signals out of chaos
when a good reference orbit is available.
The probabilistic cleaning method and the manifold
decomposition method both use features of the chaos
which are simply not seen in a traditional Fourier treatment of signal-separation problems. In conventional approaches to noise reduction the "noise" is considered unstructured stuff in state space with some presumed
Fourier spectrum. Here we have seen the power of
knowing that chaotic signals possess structure in state
space, even though their Fourier spectra are continuous,
broadband, and appear as noise to the linear observer.
We are confident that this general observation will prove
the basis for numerous applications of nonlinear dynamics t o signal-based problems in data analysis and signal
synthesis or communications.
C. Knowing very little

40

45

50

55

FIG. 34. A pulse added to data from the Henon map. Signalto-noise ratio is 20 dB. Pulse is in the neighborhood of time
step 75.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

When we have no foreknowledge of the signal, the dynamics, or the contaminant, we must proceed with a
number of assumptions, explicit or implicit. We cannot
expect any method to be perfectly general, of course, but
within some broad domain of problems we can expect to
succeed in separating signals with some success.
In the work that has been done in this area there have
been two general strategies:
Make local polynomial maps using neighborhood-toneighborhood information. Then adjust each point to
conform to the map determined by a collection of points;
i.e., the whole neighborhood. In some broad sense a local
map is determined with some kind of averaging over
domains of state space, and after that this averaged dy-

Abarbanel et a/.: Analysis of observed chaotic data


namics is used to realign individual points to a better
deterministic map. The work of Kostelich and Yorke
(1991) and Farmer and Sidorowich (1991, 1988) is of this
kind.
Use linear filters locally or globally and then declare
the filtered data to be a better "clean" orbit. Only
moving-average or finite-impulse-response filters will be
allowed, or we will have altered the state-space structure
of the dynamics itself. Many papers have treated this issue. Those of Pikovsky (1986), Landa and Rosenblum
(1989), Sauer (1991), Schrieber and Grassberger (1991),
and Cawley and Hsu (1992) all fall in this class.
The blind cleaning proceeds in the following fashion.
Select neighborhoods around every data point y(n) consisting of NB points y (r) (ft), r = \,2, . . . ,NB. Using the
y{r\n) and the y(r,n + 1) into which they map, form a local polynomial map
y ( r , / t + l ) = g(y ( r > <n),a)'
= A + B-,y (r kn) + C-y ( r ) (n)y ( r ) (n)+ . ,
(130)
where the coefficients are determined by a local leastsquares minimization of the residuals in the local map.
Now, focusing on successive points along the orbit y(n)
and y(n + 1 ) , find small adjustments 8y(n) and by(n + 1 )
which minimize
| | y ( r c + l ) + 8 y ( t t + l ) - g ( y U ) + 8y(n))|| 2

(131)

at each point on the orbit. The adjusted points


y(k) + 8y(k) are taken as the cleaned orbit.
In the published work one finds the use only of local
linear maps, but the polynomial generalization seems
clear. Indeed, it seems to us that use of local-measurebased orthogonal polynomials </>a(x) would be precisely
in order here as a way of capturing, in a relatively
contamination-robust fashion, a good representation of
the local dynamics, which would then be used to adjust
orbit points step by step. Using only local linear maps is
in effect a local linear filter of the data, which does not
explicitly use any features of the underlying dynamics,
such as details of its manifold structure or its statistics, to
separate the signal from the contamination. We suspect
that the order of signal separation is limited to
O (1 /\/NB)
as a kind of central limit result. This is actually quite sufficient for some purposes, though it is unlikely to allow for separation of signals that have
significant overlap in their Fourier spectrum or for separation of signals to recover a useful version of the original, uncontaminated signal as one would desire for a
communications application. The method of Schrieber
and Grassberger (1991) injects an implicit knowledge of
dynamics into their choice of linear filters. They utilize
an embedding with information both forward and backward in time from the observation point. Thus, as they
argue, the exponential contraction backward and forward along the invariant manifolds in the dynamics can
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1377

significantly improve the removal of the contamination.


The second approach to separating signals using no
special knowledge of the dynamics rests on the intuition
that a separation in the singular values of the local or
global sample covariance matrix of the data will occur
between the signal, which is presumed to dominate the
larger singular values, and the "noise," which is
presumed to dominate the smaller singular values. The
reasoning is more or less that suggested in an earlier section: the "noise" contributes equally to all singular
values, since the singular values of the sample covariance
matrix measures how many data, in a least-squares sense,
lie in the eigendirection associated with that singular
value. If the data are embedded in a rather large dimension >dE, the singular values with the largest index are
populated only by the "noise," while the lower-index
singular values are governed by the data plus some contamination. While this may be true of white noise, it is
quite unlikely for " r e d " noise, which has most of its
power at low frequencies. Nonetheless, given this limitation, the method may work well.
The idea is to form the sample covariance matrix in dimension d>dE and then project the data (locally or globally) onto the first ds singular values. This is then taken
as new data, and a new time-delay embedding is made.
The procedure is continued until the final version of the
data is "clean enough." Landa and Rosenblum (1989)
project onto the direction corresponding to the largest
singular value and argue that the amount of "noise
reduction" is proportional to dE/d multiplied together
the number of times the procedure is repeated. Cawley
and Hsu (1992) do much the same thing, with interesting
twists on how the "new" scalar data are achieved, but locally. The projection onto the singular directions is a
linear filter of the data. A combination of local filters,
each different, makes for a global nonlinear filter, and in
that sense the method of Cawley and Hsu represents a
step forward from the earlier work of Landa and Rosenblum. Pikovsky (1986) also does a form of local averaging, but does not use the local singular-value structure in
it.
All in all one can say that there is much exploration
yet to be done on the properties and limitations of blind
signal separation. The hope is certainly that a method,
perhaps based on invariant-measure orthogonal polynomials, that utilizes the local nonlinear structure of the
data will have some general applicability and validity.
VIII. LINEARLY FILTERED SIGNALS
An important footnote to the discussion of signal separation is to be found in the issue of the effect of filtering
by linear filters by convolution on a chaotic signal. All
signals are observed through the filter of some measuring
instrument. The instrument has its own characteristic
response time and bandpass properties. Many filters can
be represented as a convolution of the chaotic signal s (n)
with some kernel h (n) so that the actual measurement is

Abarbanel et al.: Analysis of observed chaotic data

1378

S{n) = ^h(j)s{n-j)

(132)

The question of what happens to the properties of the


signal under this filtering has been addressed in many
places (Sakai and Tokumaru, 1980; Badii et al., 1988;
Mitschke et al., 1988; Chennaoui et al., 1990; Isabelle
et al., 1991) with the following basic results. If the filter
is a so-called moving-average or finite-impulse-response
(FIR) filter, that is, a kernel of the form
M

5 ( n ) = 2 h(j)s(n-j)

(133)

then the dimension and other characteristics of the attractor are unchanged. This is more or less obvious,
since this filter is simply a linear change of coordinates
from t h e original s(k), to S(k). By t h e embedding
theorem, invariant properties of the system seen in the
reconstructed phase space should be unaltered.
In t h e case of an autoregressive or infinite-impulseresponse (IIR) filter, the story is different. A model of
this kind of filter is
s(n+l)

= F(s(n))

(134)
S(n+l)

= aS(n)+g(s{n))

The first equation is just the dynamics of the system. The


second produces from the signal sin) an observation S(n)
that has its own (linear) dynamics [S(n + l) = aS(n)]
plus a nonlinear version of s(n). By adding another degree of freedom to the original dynamics, it is no wonder
that one can change the properties of the observed chaotic attractor. One practical implication of this, pointed
out by Badii et al. (1988), is that low-pass filtering one's
data, a quite common practice in experimental setups,
can cause an anomalous increase in estimated dimension.
It is rather succinctly pointed out by Isabelle et al.
(1992) that the criterion for a change in the Lyapunov dimension is the nonconvergence of the sum
2

\h(k)\exp[kkN] ,

(135)

k=o

where kN is the smallest Lyapunov exponent of the total


system. A F I R filter always converges, while an I I R
filter may fail to converge and thus change properties of
the apparent attractor.
The nonlinear dynamics reason for this is that, if the
additional dynamics characterizing t h e filter results in
Lyapunov exponents that are large and negative, their
effect on the original system is essentially unnoticed because small deviations from the original system are rapidly damped out. If the additional exponent decreases in
magnitude, eventually it gets within the range of the original exponents and affects the properties of the original
attractor because its effects on orbits becomes both
measurable and numerically substantial.
In essence then, the issue of linear filters on a chaotic
signal is rather straightforward, since linear systems are
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

rather well understood. The results stated should act as a


sharp warning, however, since numerical issues associated with t h e clearly stated theory may lead to severe
misunderstandings. Nonetheless, from the theoretical
point of view, this is a well understood area.
IX. CONTROL AND SYNCHRONIZATION
OF CHAOTIC SYSTEMS
A. Controlling chaos
1. Using nonlinear dynamics
In this part of our review we present a method that has
been demonstrated in experiments for moving a chaotic
system from irregular to periodic, regular behavior. T h e
method has two elements: (1) an application of relatively
standard control-theory methods, beginning with a
linearization around the orbits, and (2) an innovative idea
on how to utilize properties of the nonlinear dynamics to
achieve linearized control in an exponentially rapid
fashion. One of t h e keys to t h e method is detailed
knowledge of the phase space and dynamics of the system. This can be gleaned from observations on the system in the manner we have outlined throughout this review. An attractive feature of the method is its reliance
on quite small changes in parameter values to achieve
control once the system has been maneuvered into the
appropriate linearized regime by the nonlinear aspects of
the problem.
We assume that the experiment is driven by some outside force and the data are in the form of a time series
taken once every period of the drive. We call such a data
collection technique stroboscopic. An equally useful assumption is that there is some type of fundamental frequency in the nondriven dynamics. We would then use
one period of this frequency for our stroboscopic time.
Next we shall assume an adjustable parameter which we
call p. Let the data vectors on the stroboscopic surface of
a section be given by v(n) for n = 1,2, . . . ,N.
Embedded within a hyperbolic strange attractor is a
dense set of unstable periodic orbits (Devany and Nitecki, 1979). The presence of these unstable periodic orbits and the absence of any stable ones is a characteristic
sign of chaotic motion. The goal of the control algorithm is to move the system from chaotic behavior to
motion along a selected periodic orbit. The orbit is unstable in the original, uncontrolled system but clearly
stable in the controlled system. The reason for this apparent paradox is that the controlled system has a larger
phase space, since the parameter that was fixed in the
original dynamics is now changed as a function of time.
In the new dynamics, the periodic orbit of the original
system becomes stable.
The first step in solving this problem is obvious. We
must find the locations of the periodic orbits. We discuss
how this can be accomplished for the two cases intro-

Abarbanel et al.: Analysis of observed chaotic data


duced above (Auerbach et al., 1987). In the first case the
system is not driven, but there is some fundamental frequency. We choose some, typically small, positive number 6. For each data vector y(n), we find the smallest
time index k > n such that
\\y(n)-y(k)\\<e

We define the period of this point by M ~k n. As an


example, experimental data collected from the BelousovZhabotinski (BZ) chemical reaction (Roux et al., 1983;
CofFman et al., 1987) indicate that there is a fundamental
periodic orbit yM( 1 )h^y M (2)n^ *->y M (M) = y M ( 1)
near
M X25.
Our
stroboscopic
vectors
are
v(m) = y(mM) for m = 1 , 2 , . . . .
When the system is driven, the procedure is even
easier. We again look for recurrent points. Typically for
driven systems with data acquired once every period of
the drive we will find M = \. There is only one major
difference between this system and the previous one. For
the previous system the data vectors between y M ( l ) and
y M ( M ) = y M ( 1) give the trajectory covered by the periodic orbit. For data that are taken only once every period
of the drive we usually do not have the details of trajectory of the periodic orbit with the lowest period.
The evolution of orbits is taken to be governed by a
map v(n -\~l) = F(v(n),p) with some parameters that are
at our disposal. The exact phase-space location of the
periodic orbit v M H->VM is a function of the parameter p.
We choose the origin of values of p to be on the orbit we
have found.
We want to vary p in the neighborhood of p = 0 , when
the orbit of the undriven system comes "close" to the
v M ( 0 ) = v M (/? = 0 ) . Now, we must either wait until the
experimental chaotic orbit comes near this or use some
technique to force it into this region (Shinbrot et al.,
1990, 1992). Once v comes near v M ( 0 ) the control is
given by the following steps, illustrated in Fig. 36 (Ditto
et al., 1990).
The part of the figure on the left indicates the situation
just prior to the control mechanism's being activated.
The unstable fixed point is indicated by v M ( 0 ) . Using the
techniques we described in the previous section, we
evaluate D F [ v M ( 0 ) ] , the Jacobian matrix of the mapping

at the point v M ( 0 ) . We also determine the directions of


the stable and unstable eigenvectors of the Jacobian. The
location of the phase-space trajectory at time n is indicated in this figure by the vector v(n).
The control mechanism is an instantaneous change in
the parameter from p =0 to p =p. For p p the cross
shown in solid lines on the left figure becomes the cross
shown in dashed lines in the center figure (Ditto et al.,
1990). This change is nothing more than the small
change in the location of the periodic orbit after a change
in the parameter, i.e., v M ( 0 ) moves to v M ( p ) . We have
indicated the location of the cross prior to the change by
a solid line. What is important is that the change in the
parameter has created a situation that will force the trajectory of v{n) in the direction of the stable manifold of
v M ( 0 ) . In fact, one chooses the magnitude of the parameter p so that on the next time around the orbit v( n + 1 )
is on the stable manifold of v M ( 0 ) . One then instantaneously moves p from p back to p = 0 . From now on the
trajectory will march along the stable manifold (of the
uncontrolled system) until it reaches the periodic point
*M (0).
The hard part of the control process is determining the
value of p. When p =p the dynamics near vM{p) can be
approximated by the linear map
v U + l ) = vM(^) + DF_-[vU)-vM(/>)] .

^F(^O)

FIG. 36. Sketch of control algorithm invented by Ott et al.


(1990): (a) if the rcth iterate n is near the fixed point |>(Po)
then (b) one changes the parameter p from p0 to p0 + dp to move
the fixed point so as to (c) force the next iterate onto the stable
manifold of the undisturbed fixed point (from Ditto et al.,
1990).
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

(136)

We must choose p so that in one period of the drive the


phase-space trajectory from v( n) to v( n -f 1) lands on the
stable manifold of v M ( 0 ) . To choose this value of/?
define the vector g via the following rule:
3v M
dp

P=P*

VM(P)

To simplify the calculations we shall confine ourselves


to d =2 dimensions, though the method is readily generalized to higher dimensions. Since p is small, we assume DF_ can be approximated by D F p = = 0 = D F . Using
the techniques described above one can find D F from the
data vectors v(n). Furthermore, one can write D F as
D F = A tl fif n + A,3f,

?,F(PQ + SP)

1379

(137)

where the As and Au are the stable and unstable eigenvalues, and t and u are the stable and unstable eigenvectors. The vectors fs and fu are the stable and unstable
contravariant eigenvectors defined by f u - u = l , fu-/s = 0,
fg-'s^l, and f s -u = 0.
Equations (136) and (137) and the vector g are all that
is necessary to define the control mechanism. The
motion along the stable manifold of vM brings the original orbit very rapidly to the vicinity of the unstable
periodic orbit vM. Once the orbit is in this vicinity, small
changes in p proportional to v v M are made. This
linearization defines a new linear map near vM, and we

Abarbanel et al.: Analysis of observed chaotic data

1380

require the eigenvalues of this new linear problem to be


within the unit circle. The requirement on the change in
parameter necessary to control the oscillations is given
by
v(/i)-f n
A-l

(138)

g-f

This type of control mechanism has been experimentally verified in an experiment on magnetoelastic ribbons
(Ditto et al., 1990). The device involves a ribbon whose
Young's modulus changes under the influence of small
magnetic fields. An inverted ribbon is placed in the
Earth's gravitational field so that initially it buckles due
to gravity. By adding a vertical oscillating magnetic
field, one creates a driven chaotic oscillator. For this experiment the force driving the oscillation is gravity, while
the control parameter is the magnetic field strength. The
results of this experiment are shown in Figs. 37 and 38.
In Fig. 37 the iteration number represents passes
through the stroboscopic surface of a section, which is
one period of the drive. The gray portion for iterates less
than approximately 2350 represents the chaotic motion
of the ribbon before the control is initiated. For an
iterate near 2350, the motion of the ribbon passes near an
unstable period-one orbit, and the control is initiated.
After initiating control the motion of the ribbon remains
near the periodic orbit. In the figure this is indicated by
the fact that the location of the orbit on the surface of
the section is constant near 3.4 for all subsequent intersections. The period of the driving of this magnetic field
is typically 0.85 Hz; thus the period-one orbit has period
1/0.85 Hz1.2 sec. The experimenters have been able
to confine the motion of the ribbon to the unstable orbit
for as long as 64 hours. Clearly the control mechanisms
works.
Figure 38 is an admirable piece of showing off on the
part of the researchers. First they control near a periodone orbit. Then they change their minds and control
near period-two orbits. Then they change their minds
again and control back on the original period-one orbit.
Control has also been demonstrated near period-four orbits.

1000

2000

3000

4000

Iteration Number
FIG. 37. Time series of the voltages stroboscopically sampled
from the sensor attached to magnetoelastic ribbon (Ditto et al.,
1990). Control was initiated after iteration 2350.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

2000

4000
6000
Iteration Number

10000

FIG. 38. Time series of voltages from the magnetoelastic ribbon. System is switched from no control to controlling the (unstable) period-1 orbit (fixed point) at iteration 2360, to controlling period-2 orbit at n =4800, and back to the fixed point at
=7100.

2. More traditional methods


It is also possible to influence the dynamics of a chaotic system using traditional techniques from control
theory (Cremers and Hubler, 1986; Eisenhammer et ah,
1991; Singer and Bau, 1991; Singer et al, 1991). To illustrate this approach we imagine that the uncontrolled dynamics is given by the equation
dy_
dt

(139)

F[y,f]

and that y*(t) represents the trajectory we are interested


in controlling (stabilizing or, possibly, destabilizing).
Traditional control techniques implement a feedback
mechanism to manipulate the dynamics. The feedback
is an externally applied force we denote as
H[y(t) y*(t),t;p].
The vector p contains adjustable parameters, which are often called the feedback gains.
When the control mechanism is implemented the dynamics changes from Eq. (139) to

^=F[y(t),t]
at

H[y{t)-y*(t),t;p]

(140)

The purpose of H is clear from Eq. (140). H is a force


that gives nudges to orbits. If y* is unstable without the
control, and the trajectory y moves away from y*, then
H can push it back towards y* and thus stabilize an undriven, unstable orbit.
This type of control has been implemented in a laboratory thermofluid experiment. The experimental device
consisted of a torus heated from below and cooled from
above. This thermosyphon implements convection,
behaving in a similar fashion to the original 1963 Lorenz
equations. The feedback was implemented by controlling
the current flow into the heater as applied to the lower
part of the torus. The control part of the heating was
K(AT AT0) where AT is the actual temperature
difference between two selected points and AT0 is the
desired temperature difference. The value of K was much
smaller than the constant heat component applied to the
bottom part of the torus (Bau and Singer, 1992).
Some of the results of this experiment are shown in

Abarbanel et al.\ Analysis of observed chaotic data

1381

role of external forces is to effect changes in the parameters of the system rather than the form of the system.
The main distinctions are the use of the nonlinear dynamics of the system in the first method and the important fact that only infinitesimal external forces are required to change parameters. This latter point comes
from the use in the method of the exponential approach
along stable manifolds.

fl|^^

3. Targeting
32

*0

48

64

72

80

Time (minutes)

FIG. 39. Experimentally measured temperature oscillations as


a function of time. The controller was activated 33 min after
the start of the experiment (Bau and Singer, 1992).
Figs. 39 and 40. Figure 39 shows the temperature profile
A T as a function of time. The control is implemented at
time equal to 33 minutes. The control mechanism causes
the dynamics of the temperature to stabilize about a particular value that indicates steady laminar flow, corresponding to counterclockwise rotation of the fluid. Figure 40 is showing off in a manner similar to the magnetoelastic ribbon example. In this case the control is used
to change from steady counterclockwise motion to steady
clockwise motion.
It is important to note that the two main types of control we have discussed are not equivalent. The first type
involves changing parameters of the system to achieve
control. It also used a knowledge of the location and
function of the stable and unstable manifolds of the dynamics. The dynamics in this control technique is given
by
dy _ ,
i , 3F
- ^ - F T[ yr ; p ] + . 8E p .
In this equation there is no need for a new class of external forcings as in the one found in Eq. (140). Instead the
18

-15
-19
20

25

30

Minutes

FIG. 40. Switching the direction of the flow in a thermoconvection experiment from clockwise to counterclockwise using the
controller (Bau and Singer, 1992).
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

The final type of control that we shall discuss differs


from both of the two previous cases. As with many
things, the control mechanisms work well when they
work. After the system moves into a region of the phase
space that we wish to control, we are able to implement
the control with great success. However, we have to wait
until the dynamics approaches the desired region before
control can be initiated. This can often take a long time.
What would be very useful is some technique to persuade
the dynamics to approach the control region quickly. In
other words, we need a net that can be thrown over the
phase space in order to capture the trajectory. Once it is
in the net we can drag it back to our control cage and
lock up its behavior.
There are two different techniques for directing trajectories towards specific regions of phase space. They
differ in detail and implementation, but the targeting goal
is the same. The first technique is useful when the target
point is on the attractor given by the data. When the target is not on the attractor given by the data, the second
technique is required.
The first targeting technique involves both forward
and backward iterations of small regions of the phase
space. Assume that the data are given on a surface of a
section by v(), n = 1,2, . . . ,N. For purposes of illustration also assume that the dynamics is in two phase-space
dimensions and is invertible. Let the initial point be
given by v, and the target point by v r . Without altering
the dynamics, the time necessary for a typical point on
the attractor to enter a ball of size et centered on v r
scales like r ~ l//x(e,), where JJL is the natural measure on
the attractor. This measure scales with dl9 the information dimension of the attractor (Badii and Politi, 1985),
s o r ^ ( l / 6 f ) l. Obviously for small et the time it takes
for a trajectory to wander into the target region can be
quite long. This phenomenon is evident in Fig. 41, where
it took thousands of iterations before the trajectory was
near enough to the period-one target to implement control. Other examples of even longer duration exist (Shinbrot et aL, 1990).
Clearly this is not acceptable, and happily it can be improved upon. Assume that the dynamics on the strobosc o p e surface is given by
v(n+l)

= F[v(n);p]

where we have explicitly indicated the parameter depen-

Abarbanel et al.: Analysis of observed chaotic data

1382

1000
1/E,

FIG. 41. Average time required to reach a typical target neighborhood from a typical initial position as a function of the inverse neighborhood size: dashed line, no control; circles and
solid line, with control (from Shinbrot et al., 1990).
dence of F by p. For now we shall use only one parameter, but that is not necessary. The change in one iteration
on the surface due to a change in the parameter is
8v(n + l ) = v(w + l ) - v U + 1 )
=

3F[v(n)]
dp
dp

If we choose 8p small, then \{n + 1) v(n + 1 ) sweeps


out a small line segment through the point v ( + l ) .
After m steps through the unperturbed map F[v;p],
the line segment will have grown to a length
v(/i +m) v(n + M ) ~ e x p [ m ^ 1 ] 8 v ( r c + 1 ) .
Eventually
the length will grow to the same order as the size of the
attractor, which we shall conveniently call 1. The number of iterations necessary for this to happen is
ml = log[8v(n + l)]/X1. We now imagine iterating the
target region eT backwards in time. This process will
transform the target region into an elongated ellipse.
The length of the ellipse will grow like exp[A, 2 ra]. After a
total of m2 iterations the ellipse will have been stretched
to the point where its length is the same order as the size
of the attractor.
Generically the line formed by the forward iterations
of 8v(n + 1) will intersect the ellipse formed by the backward iterations of the region T. For some value of the
parameter, call it px, in the range p +8/? the end of the
line segment will be inside the backward images of eT.
By changing the parameter at time n from the value p to
the value px, we can guarantee that the trajectory will
land inside the target region. The number of iterations
required for this to happen is
rmx

+ra.

ln[6v(/t+D]

ln[er]

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

The targeting strategy above utilizes only one change


in the parameter. After the initial perturbation, the parameter value is returned to p from p x. It also assumes
that one has complete knowledge of the dynamics. In the
presence of noise, or incomplete knowledge of, the dynamics the above procedure will not work without
modifications. The modifications are very small. Instead
of only perturbing the dynamics once at time n, one
should perturb the dynamics at each iteration through
the surface. Therefore the entire forward-backward calculation should be done at \(n), v(n + 1 ) , v(n + 2 ) , etc.
until the target region is reached.
The forward-backward technique has been demonstrated to be effective in computer experiments (Shinbrot
et al., 1990). More impressive is experimental implementation of these techniques. For the experimental situation only the forward portion of the targeting method
was used. This is equivalent to setting m 2 = 0 and waiting for the stretched line segment to contact eT. The results indicate that for the magnetoelastic ribbon experiment without targeting it required about 500 iterations to
reach a typically chosen T. With control the number of
iterations typically was reduced to 20 (Shinbrot et al.,
1992). The power of using the nonlinear dynamics of the
process is quite apparent.
The second type of targeting is much more ambitious.
The basic problem is still the same. Assume that initially
the trajectory is located at yt. How can we control the
system so that it quickly arrives at the location yfl Notice that the locations y, and y^ are legitimate states of
the dynamics for some value of the parameters. Other
than that they are arbitrary. The type of control that accomplishes this broad category of behavior has the drawback of requiring a complete model of the dynamics for
all values of the parameter.
The first step in the control technique involves using
cell maps to determine the dynamics of the entire phase
space for a range of parameter values, pij) where
j.= 1,2, . . . , Np (Hsu, 1987). Each different value of the
control parameter p{j) implies a different dynamical system for Eq. (139) and hence different trajectories in its
corresponding cell maps. The dynamics may even experience bifurcations without a conceptual change in the
control technique. Once the cell maps have been established, one searches through the possible trajectories. A
final path is chosen by using pieces of the possible trajectories provided by the cell maps. Thus the entire controlled trajectory can be given by the sequence of cells
S =(S{0),Sil),
. . . ,S{1)). Each of the individual S{jhs is a
true trajectory of the dynamical system y =
F[y,t;p{j)].
iJh
The criteria for determining the S s is that S be as
short as possible.
The details of the construction of the cell maps and the
search algorithm for determining the shortest 5 from the
large number of possible values can be found in the references (Bradley, 1991, 1992). Once the final phase-space
destination is achieved, standard control techniques, as
above, can be employed to ensure that y stays near yy.

Abarbanel et al.: Analysis of observed chaotic data


B. Synchronization and chaotic driving

wx w2 >

Many physical phenomena can be modeled as an underlying dynamical system driven by external forces.
The most often studied types of driving are periodic and
steady forcing. A down-to-earth example is given by a
system that is driven with a period half that of the natural oscillation. This models pumping a playground
swing.
A type of driving that is new to our field of inquiry is
chaotic driving (Kowalski et al., 1990; Pecora and Carroll, 1990, 1991a, 1991b). Under this type of scenario one
has a dynamical system that is driven by the output of
another dynamical system. If the second dynamical system, the one that is providing the drive, is exhibiting
chaotic motion, then we say that the first system is undergoing chaotic driving. In Fig. 42 we schematically illustrate chaotic driving. The first dynamical system, the
one that is doing the driving, we call the driving system.
The second dynamical system, the one that is being
driven we call the response system.
Imagine dividing the two dynamical systems into three
parts. The degrees of freedom of the driving system are
uU) and v(t) while those of the response system are
called wit).
The vector fields for the driving and
response are g, f, and h, and the equations of motion are
du
= g[u;v] ,
dt
dv
= f[u;v],
dt

(141)

dw
= h[w;v] .
dt
The first and second equations of motion involve the
driving system. We have split it into two parts. The first
part represents the variables in the driving system that
are not involved in driving the response. The second part
represents the variables that are actually involved in the
driving of the second dynamical system. Finally, the
third equation represents the response system. We shall
assume that the combined system has n equations of
motion, while partitions g, f, and h have k, m, and /
equations of motion, respectively.
A simple familiar example is provided by a damped
driven pendulum: 0-\-y6^-sm(0)
= cos(cot). This system
can be written in the form of Eqs. (141) through the
identification of variables: B = w1, 6 = w2, (f> v, and
(f> = u. The result is written

FIG. 42. Sketch of chaotic driving: chaotic generator 1 drives


another chaotic system 2.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1383

ib2::=yw2sm(ivl)-\-v
v=v

In this example m = 1, k = 1, and / = 2 . The driving system is a simple harmonic oscillator and the response is
the damped pendulum.
In chaotic driving we replace the simple harmonic oscillator with a chaotic system. When the chaotic system
is identical to the response system the most dramatic
effects occur. Under these conditions the chaotic response
system may synchronize with the chaotic driving system.
Synchronization means here that the two systems follow
the same trajectory in phase space (Afraimovich et al.9
1986). This occurs despite the fact that the response system has a different initial condition from the drive system. Under normal chaotic circumstances a different initial condition would imply that the two trajectories
would diverge exponentially on the attractor. Thus the
fact that chaotic driving can cause two systems to follow
the same trajectory for a long time is quite an attention
getter.
When does synchronization occur? That is, when will
two nearby initial conditions in w converge onto the
same orbit? The condition that determines whether or
not a driven dynamical system will synchronize is obtained by examining the values of the
conditional
Lyapunov
exponents.
To determine the conditional
Lyapunov exponents, as well as to define them, we must
simultaneously solve Eqs. (141) and the variational equations of the response system,
d(Aw)
dt

= DFw[v,w]-Aw ,

(142)

where Aw = w ' ( f ) - w ( f ) and


( E > F w W = dwr.
The eigenvalues e x p [ L ^ ] of compositions of D F W are the
conditional Lyapunov exponents.
For periodic trajectories in w, understanding the
behavior of Eqs. (141) and (142) and its influence on Aw
is a straightforward application of Floquet theory. But
we are interested in chaotic instead of periodic trajectories. A moment's thought convinces us that determining the behavior of Aw is equivalent to calculating the
Lyapunov exponents associated with Eqs. (141) and (142).
Numerical experiments have determined that the
Lyapunov exponents associated with Eqs. (141) and (142)
are not the same as the Lyapunov exponents calculated
only from Eq. (141) alone. N o r are they a subset of those
calculated from Eq. (141) alone. The first of these facts is
obvious. The variational equations associated with Eq.
(141) have an nXn
Jacobian and will produce n
Lyapunov exponents. On the other hand, Eq. (142) has
an IXI
Jacobian and will produce / conditional

Abarbanel et al.: Analysis of observed chaotic data

1384

TABLE III. Conditional Lyapunov exponents for synchronized


Lorenz attractors. With z as the drive, one of the conditional
exponents is positive, so this will not synchronize.
System

Drive Response Conditional Lyapunov exponents

Lorenz
a = 16
6=4
r =45.92

(y,z)

( 2.50,-2.5)

(x,z)

(-2.95,-16.0)

(x,y)

(0.007 89,-17.0)

Lyapunov exponents.
The set of / Lyapunov exponents associated with Eqs.
(141) and (142) are the conditional Lyapunov exponents.
If all / conditional exponents are less than zero, then the
subsystem, w, will synchronize [i.e., w'(t)>w(t) as
f>oo]. If we denote the conditional Lyapunov exponents as 0>%1<X2* -A,/ then the convergence
rate is given by expfA^]. Numerical experiments on the
Lorenz system are given in Table III. As the table
shows, for the parameter values used one can implement
either x or y as the drive. The corresponding subsystems
[(y,z) in the case of x drive and (x,z) in the case of y
drive] have two negative conditional Lyapunov exponents. The third possible subsystem has z as the driver
and possesses a small positive conditional Lyapunov exponent. The positive conditional Lyapunov exponent implies that this type of drive/response system will not synchronize.
One can cascade several dynamical systems together
(Carroll and Pecora, 1992; Oppenheim et a/., 1992) and
achieve synchronization in all of them by using only one
drive. An example of this type of system is given in Fig.
43. Here we show a system of three cascaded Lorenz systems. The x coordinate of the first is used to drive the
second. Similarly, the y coordinate of the second is used
to drive the third. One could extend this cascade
indefinitely without changing the results. Moreover,
since the second and third systems respond to the drive
of the first system there is really only one drive, namely,
the x coordinate of the first system.

difference

4x
I

Drive

-L-^x

Response

FIG. 43. Sketch of cascaded chaotic systems.


Rev. Mod. Phys., Vol. 65, No. 4, October 1993

X. SPATIO-TEMPORAL CHAOS
As we remarked in the introduction, the study of
spatio-temporal chaos is much further from a full understanding than solely temporal chaotic behavior. The situation is changing quite quickly, and while we shall
merely touch upon our own expectation of where this
area will go, it seems certain that the developments will
bring the analysis of spatio-temporal observations to the
same stage as that of temporal chaos, reviewed in this article. So what we include now is less a review than, we
hope, a preview to indicate how the subject might develop from the results reviewed throughout this article.
In the last few years increasing attention has naturally
been focused on spatio-temporal chaos in various experimental setups such as Rayleigh-Benard convection
(Manneville, 1990), Taylor-Couette flow (Anderek and
Hayot, 1992), and surface excitations of the Faraday sort
(Gollub and Ramshankar, 1991). On the theoretical side
there have been extensive studies of cellular automata
which provide coarse graining in the dynamical variables
as well as in the independent variables of space and time,
coupled map lattices (Keeler and Farmer, 1986), which
study the interaction between known chaotic maps coupled together on spatial lattices, and model physical
equations such as the Kuramoto-Sivashinsky and various
Ginzburg-Landau equations (Kuramoto, 1984). Discussion of those studies as well as relevant mechanisms of
transition to chaos go far beyond our scope. In that regard, however, we recommend the review by Cross and
Hohenberg (1993), which reports on studies in which the
instabilities of linearized dynamics can be traced as the
origin of observed space-time patterns in a variety of experiments.
Our present aim is to outline the approach to analysis
of data from spatio-temporal systems following the philosophy and methods developed for the pure temporal
cases in previous sections. In other words, we need to develop machinery that allows us to calculate meaningful
characteristics identifying the spatio-temporal behavior
and to build models for prediction of behavior of the
spatio-temporal systemand to do all of this working
with data.
The main difference (and difficulty) which comes with
spatial degrees of freedom is the very high (or even, typically, infinite) dimension of the system. Of course, one
may take only one observable from one space location
and try to reconstruct the dynamics in the usual way.
The idea would be that since the system is totally coupled, by the embedding theorem, any measurement
should reveal the full structure of all degrees of freedom
in interaction. This standard approach of treating a
spatio-temporal system as though it were purely temporal, while possible in principle, is practically infeasible
with existing or projected computational capability. It
seems much more promising to account for the spatial
extension explicitly at both the phase-space reconstruction and model-making stages of the problem. We have

Abarbanel et a/.: Analysis of observed chaotic data


seen in the preceding sections that a scalar time series
s(n) represents the projection of a multidimensional
geometric object onto the observation axis. In the case of
spatially extended systems, we deal with similar scalar
measurements labeled by space as well as time.
To be concrete, we restrict our discussion to one space
dimension, call it x, and time. The amplitude which is
measured we call a(x,t).
The first step towards the
reconstruction of the spatio-temporal phase space is to
consider just space series rather than temporal ones
(Afraimovich et al., 1992). This approach is quite legitimate if we study the "frozen" spatial disorder resulting,
for example, from the evolution of a gradient spatiotemporal system of the form

J ^ = _JM
dt

(1 43)

oa

where F(a) is some free-energy functional. For this kind


of dynamics, it is easy to show that any initial state

a(x,0) evolves to a steady state a^ix).


For this pure
space series all the machinery developed for time series
can be directly reformulated by substituting x for t. In
this way one can calculate fractal dimension of the
"space" attractor, evaluate spatial Lyapunov exponents,
make models, etc.
A richer situation arises when we have spatio-temporal
dynamics and measure the spatio-temporal series a(x,t).
The first thing we want to deal with is the correct phase
space for this form of measurement. Recall that we
reconstruct the phase space in a temporal case using
time-delay coordinates. Determination of the time delay
T in the temporal case requires that the measurements
x(t) and x(t+T)
were independent in an informationtheoretic sense. In the case of space as well as time, the
natural modification of this procedure is to seek a space
delay at which observations become slightly uncorrelated
(in an information-theoretic sense again). In this way we
arrive at the idea of a "matrix" phase space composed of
Mx (spatial) by Mt (temporal) matrices formed from
space and time delays of the observations a ix,t):

a(i,n+(Mt-l)T)

a(i,n)

a(i+D,n+(Mt-l)T)

a(i+D,n)
a(i+(Mx-l)D9n)

1385

a{i +(Mx-l)D,n

+{Mt-l)T)

where a(i,n) = a(x,t) at a(x0+iD,t0-\~nT),


and the time
delay T is an integer multiple of the sampling time r 5 ,
while the space delay is an integer multiple of the spatial
sampling A 5 . If we have Nt data points in the temporal
direction n = 1 , 2 , . . . , Nt and Nx points in the space
direction / = 1,2, . . . , Nx, then the total number of matrices
available
to
populate
the
space
is
(Nt ~Mt) X (Nx -Mx).
It will be usual to have Nx MX
and
NtMt.
To determine the time delay we use average mutual information in the familiar fashion. To determine the
space delay Z>, we do the same, but now we must compute the average mutual information for measurements
a(n,i) and measurements a(n,i-\-D)
averaged over all /
and n to cover the attractor. Thus, it requires just a
slight generalization of existing algorithms for temporal
chaos.
To determine the dimension of our "matrix" phase
space MtMx9
we need to establish values for the
minimum number of coordinates needed to unfold the attractor observed through the projection onto our measured a(i,n).
For this we can use the method of false
nearest neighbors (Kennel et al.9 1992), which works
efficiently in the case of temporal chaos alone.
The result of this analysis will be the Mt XM Z dimensional state space within which we can work to determine
the same kind of quantity of interest as in temporal
chaos: dimensions, Lyapunov exponents, models, etc.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

We denote the members of our state space by Aaf3{i,n)>


where the ranges of a and fi are a = 1 , 2 , . . . ,Mt and
/ 3 = 1 , 2 , . . . ,MX while the "labels" (i,n) run over
/ = 1,2, . . . ,NX-MX
a n d / i = 1 , 2 , . . . ,NtMt.
Essentially the same approach can be employed if the
object of interest is a "snapshot" of a dynamical process
with two spatial dimensions, namely, an image. We need
only change the "labels" from xyt to x,y and the variables
to a(x,y) = a(xQ + ikx, y0+jky)
with space delays
Ax,Ay.
We would establish two embedding dimensions
Mx,My, and the remainder of the general discussion
would go as before. The resulting modeling of images as
nonlinear processes in the MxMy-dimensional
space of
matrices will then replace the present "state of the a r t "
of linear modeling (Biemond et ah, 1990).
To determine quantities such as dimensions of the
state-space a t t r a c t o r t h a t is, the dimension of the collection of matrices located in the Mt X Mx -dimensional
spacewe must address the question of the distance between two rectangular matrices. This question has been
studied in some detail by mathematicians. The most interesting work is reported in the monograph by Horn and
Johnson (1985), where the idea of matrix norms is explored. This is summarized in the book by Golub and
Van Loan (1989), where the needed computations are
also discussed. The idea is that one can always read a
matrix along its rows and then along its columns to consider it a vector of length L =MtMx, then define the usu-

Abarbanel et al.: Analysis of observed chaotic data

1386

al distances between vectors in that L-dimensional space.


This approach has been used by Afraimovich et al.
(1992) for the analysis of ID snapshots of Faraday ripples. They calculated the correlation dimension of the
patterns at different stages of disorder corresponding to
different pumping magnitude. It should be noted, however, that matrices have more structure than just long vectors; in particular, they can be multiplied in a natural
way, and thus there are new alternatives for the norms of
such objects.
The norm that seems much more relevant for the
present purposes is the so-called spectral norm. This
norm follows from the idea of a singular-value decomposition of a rectangular matrix. Suppose we have a matrix
A which i s m X n ; this can be decomposed as
A = U-2-Vr,
where the mXm
orthogonal:
U r - U = I,

(144)
matrix U and the nXn

matrix V are

Vr V= I ,

and, taking m 5: n, 2 is the nXn

(145)
diagonal matrix

2 = diag[o- 1 ,a 2 , . . . ,an ] .

(146)

The ai are known as the singular values of the matrix A.


They are also the square root of the eigenvalues of the
nXn matrix AT- A, We order them as is conventional
b y c r 1 > c r 2 > >an.
The relevance of the singular values to our considerations is that the norm of a matrix defined by
II A l l - m a x | | A - x | |

(147)

||X|| = l"

is just the largest singular value, sx( A ) . Here x is an n


vector and

IWI=(*i+*2+ * +*') 1 / 2 >


the Euclidean distance in our n space.
matrix is A = (a1,a2, . . . ,am) and
sl{A)*Va\+al+'-+a2m

(148)
If = 1 , the

which is a familiar norm.


Once one has a way of finding the appropriate space
for the discussion of spatio-temporal data, the various aspects of the analysis we have discussed in previous sections become ready for use in this enlarged context. We
consider these comments as suggestive of how the general
subject will unfold as spatio-temporal data are studied
from the nonlinear dynamics viewpoint. It remains to do
the hard work of carrying out such analyses on models
and measured data to see what general classification
schemes and tools for physics can be extracted.
XI. CONCLUSIONS
A. Summary
Over the past decade we find an enormous literature on
the analysis of temporal chaos and correspondingly

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

significant progress for purposes of studying chaotic laboratory experiments and chaotic field observations of interest to physicists. In this review we have tried to emphasize tools that are of direct utility in understanding
the physical mechanisms at work in these measurements.
We have placed less emphasis on mathematical results,
using them as a guide to what one can learn about fundamental physical processes. Numerical results are critical
to the study of nonlinear systems that have chaotic
behavior. Indeed, computation plays a larger role in
such studies than is traditional in many parts of the physics literature. Progress such as that reported throughout
this review rests heavily on the ability to compute accurately and rapidly, and as such would often not have been
possible a decade ago. The subject reviewed here is almost "experimental" in that sense through its reliance on
computers as an instrument for its study. This bodes
well for further analysis of chaos and its physical manifestations, since one can expect even more powerful computing tools to be widely available on a continuing basis.
The point of view taken throughout this review, often
implicitly, is that chaotic behavior is a fundamental and
interesting property of nonlinear physical systems. This
point of view suggests that as a basic property of the nonlinear equations governing real physical systems there is
much new to be learned by the study of chaos. While at
times we have reported on efforts to " t a m e " that chaos,
as in the section on control of periodic behavior, our
main goal has been to bring to physicists a set of tools for
extracting important information on the underlying
physics of chaotic observations. Absent many of the
tools created by those whose contributions we have reviewed here, much of the physics of chaos has been
passed by as "noise." We expect the development of
these tools to continue and the exploitation of the physics
thus exposed to blossom in applications across the spectrum of phenomena where chaos is found. Indeed, this
spectrum is so vast that our review has been inevitably
unable to cite fully even a fraction of the literature on
laboratory and field observations where chaos is known
to occur. In any case, such a full recitation might be perceived as a litany of all of classical and semiclassical
physics, and thus slightly useless.
The critical aspect of chaotic behavior from the
physicist's view is that it is broadband spectrally, yet
structured in an appropriate phase space, which itself can
be reconstructed from data. This distinguishes it from
the conventional view of "noise" and opens up for review
and reexamination a wide variety of work which
presumes unstructured stochastic behavior to be a driver
or major influence on the physical processes being studied. That reexamination has not been undertaken in this
review, nor is it a finished project to be found in the
literature. Our review has focused on the tools one will
use or the ideas underlying the tools, which will evolve
for the extraction of new physics from observations of
chaotic physics. We have no doubt whatsoever that
where physics has been hidden by the assumption of un-

Abarbanel et a/.: Analysis of observed chaotic data


structured stochastic forces, an understanding that there
may well be structure there after all will provide exciting
new discoveries.
In a broad sense, what we have covered in this review
might be cast as "signal processing" in another context.
It seems to us substantially appropriate for physicists to
have a major role in defining what constitutes signal processing for signals arising from physical systems. We expect that many of the developments over the coming decade in this field will come from the more traditional
signal-processing community. This is borne out by a
careful look at our references and a glance at the issues
now developing in the analysis of temporal chaos. Further, while we have touched upon it only slightly, we also
expect that from the same community we shall see an
effort in "signal synthesis" in addition to the kind of "signal analysis" that has been our emphasis.
This review is in many ways an update of the earlier
review by Eckmann and Ruelle (1985) about eight years
ago. Clearly much has been achieved since then, and
much rests on ideas carefully exposed in that previous review. It is impossible that we will have covered all the
topics seen as important and interesting by all the physicists, mathematicians, chemists, engineers, and others
who have contributed in the past eight years to the subject covered here. We offer our apologies to those whose
work has only indirectly been cited. We do not deem
that as a signal of its lack of importance, but eventually
the finiteness of our task prevailed. We have appended a
set of additional references in the hopes of identifying
some of the omissions in the main text. In any case, we
have made no attempt to survey the literature outside
physics and what is now called "nonlinear science" for
this review. This has excluded interesting work in
economics, statistics, chemistry, engineering, biological
sciences, and medicineto name just a few important
areas. This is a choice we have made, and we acknowledge the many interesting contributions that we have
chosen not to assess or include.
B. Cloudy crystal ball gazing
The study of temporal chaosthe analysis of physical
measurements at a point in spaceis by now rather
thoroughly explored. The various items reviewed try to
bring into focus the methods now available in a practical
sense for the study of basic properties of nonlinear physical systems in purely temporal measurements. We are
certain that new developments in many of the areas we
have touched upon and in areas we have not yet even imagined will continue to occur before the next review of
this general area. We expect numerous algorithmic improvements to be created in each of the areas we have
touched upon. We hope there will be a significant push
in the formal and rigorous statistical analysis of many of
the algorithms already available. It would be enormously
useful, for example, in the analysis of data to have some
estimate of the error in all conclusions arising from finite

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1387

data sets as well as from numerical, experimental, or instrumental uncertainty. It would be extremely important
to extend the full power of the existing topological
analysis (Mindlin et ah, 1991) in three dimensions to
higher dimensions. We anticipate extensive work on the
analysis of geometric properties of experimental strange
attractors. These may also be more robust against contamination by unwanted signals than the more familiar
derived or metric quantities such as dimensions and
Lyapunov exponents.
We look forward to the application of chaotic analysis
tools, both as covered here and in their further developments, to the detailed study of numerous physical systems. The few examples we have presented here are
clearly the "tip of the iceberg." It is in the actual making
of these applications where the push for further progress
is sure to come most persuasively. The part of the task
for the physicist which has been little touched on in this
review, since there is little to review on it as a matter of
fact, is the building and verification of models that arise
from considerations of chaotic physical experiments. We
took the liberty to say a few general words about this in
the Introduction, but the "proof of the pudding" is in
performing the task itself. It is clear, and even
mathematically good sense (Rissanen, 1989), that we cannot algorithmically discover fundamental physics. For
the physicist the road ahead with the subject of chaotic
data analysis will differ from signal processing and return
to the analysis of interesting physical systems with the
methods reviewed. The point of view in the creation of
physical models will change, of course. It will not be
necessary, we imagine, always to start with the NavierStokes equations to build models of chaos in fluids,
though, perhaps, we will learn to project those infinitedimensional equations down to the few relevant physical
degrees of freedom, which will appear in the models.
The discovery of temporal chaos in physical systems is
a "finished" topic of research, we believe. Many physical
systems have been shown to exhibit chaotic orbits, and
we are certain many more will be found. It is no longer
enough, however, to have good chaos, one must move on
and extract good physics out of it. The methods and
techniques in this review are directed towards that in the
realm of temporal chaos. An important challenge for the
area of research we have reviewed here is to extend both
the existing methods and the classification of widely expected and universal behaviors to spatio-temporal chaos.
This is to say that the study of chaotic fields remains
rather untouched from the point of view taken
throughout this review. We certainly do not mean that
early and productive steps have not been taken; they
have! We simply desire to underline the clear fact that
the developments in the analysis of spatio-temporal
chaotic data and the connection with model building, signal separation, and the other goals we have outlined are
not yet as far along as for temporal chaos. The recent
thorough review of Cross and Hohenberg (1993) makes it
quite clear that the predominant body of work in this

1388

Abarbanel et al.: Analysis of observed chaotic data

area has concentrated on the patterns and behaviors


coming from identifiable linear instabilities in spatiotemporal dynamical systems. The challenge of moving
the study of experimentally observed chaotic fields forward is a large one. Just as much of physics is nonlinear
and has postured behind easier problems posed by local
linearization for many years, so have we focused on point
measurements or "lumped parameter" models when
studying a remarkable phenomenon like chaos in physics.
The physics of chaotic fields will bring us from this too
narrow focus to substantial problems encountered in the
laboratory and in geophysical field observations. We expect the next review of this sort to be able to report major accomplishments in the analysis of experimental
chaotic fields, and we expect the lessons we have tried to
report here to pervade the way we proceed to those accomplishments.
ACKNOWLEDGMENTS
The material we have reviewed here is unquestionably
the result of conversations with colleagues too numerous
to identify individually. We have read their papers, enjoyed their talks at many meetings, and argued with them
about the meaning of each item in the text. Eventually
our own opinions are what we put into writing, and we
must thank all these colleagues for the generosity of their
time and for their critical interaction. Among those who
have been most helpful in at least making us think over
what we have gathered into this review, even if they were
unable to change our minds at all times, we wish to thank
directly J. D. Farmer, R. Gilmore, C. Grebogi, M. B.
Kennel, D . R. Mook, C. Myers, A. V. Oppenheim, E.
Ott, L. M. Pecora, M; I. Rabinovich, J. Rissanen, and D .
Ruelle. This work was supported in part by the U.S.
Department of Energy, Office of Basic Energy Sciences,
Division of Engineering and Geosciences, under Contract
DE-FG03-90ER14138, and in part by the Army
Research Office and the Office of Naval Research under
subcontracts to the Lockheed/Sanders Corporation,
Number DAAL03-91-C-052 and NQ0014-91-C-0125. In
particular, the support of the Office of Naval Research
allowed the three American authors (H.D.I.A., R.B., and
J.J.S.) to spend very productive time at the Institute for
Applied Physics in Nizhni Novgorod, Russia and has allowed the Russian author (L.Sh.T.) to join the scientific
group of the Institute for Nonlinear Science at UCSD for
an extended working period. The Russian Academy of
Science was also very helpful in making this review possible.

REFERENCES
Abarbanel, H. D. I., R. Brown, and J. B. Kadtke, 1989, Phys.
Rev. A 41, 1782.
Abarbanel, H. D. I., R. Brown, and M. B. Kennel, 1991, J.
Nonlinear Sci. 1, 175.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Abarbanel, H. D. L, R. Brown, and M. B. Kennel, 1992, J.


Nonlinear Sci. 2, 343.
Abarbanel, H. D. I., S. M. Hammel, P.-F. Marteau, and J. J.
("Sid") Sidorowich, 1992, "Scaled Linear Cleaning of Contaminated Chaotic Orbits," preprint, University of California, San
Diego, INLS.
Abarbanel, H. D. L, and M. M. Sushchik, 1993, Int. J. Bif.
Chaos 3, 543.
Abraham, N. D., A. M. Albano, B. Das, G. DeGuzman, S.
Yong, K. S. Gioggia, G. P. Puccioni, and J. R. Tredicce, 1986,
Phys. Lett. A 114,217.
Afraimovich, V. S., A. B. Ezersky, M. I. Rabinovich, M. A.
Shereshevsky, and A. L. Zheleznyak, 1992, Physica D 15, 331.
Afraimovich, V. S., N. N. Verichev, and M. I. Rabinovich,
1986, Izv. Vyssh. Uchebn. Zaved. Radiofiz. 29, 1050.
Andereck, C. D., and F. Hayot, 1992, Eds., Ordered and Turbulent Patterns in Taylor-Couette Flow(Plenum, New York).
ArnoPd, V. I., 1978, Mathematical Methods of Classical
Mechanics, translated by K. Vogtmann and A. Weinstein
(Springer, New York).
Auerbach, D., P. Cvitanovic, J.-P. Eckmann, G. Gunaratne,
and I. Procaccia, 1987, Phys. Rev. Lett. 58, 2387.
Badii, R., G. Broggi, B. Derighetti, M. Ravani, S. Ciliberto, A.
Politi, and M. A. Rubio, 1988, Phys. Rev. Lett. 60, 979.
Badii, R., and A. Politi, 1985, J. Stat. Phys. 40, 725.
Bau, H. H., and J. Singer, 1992, in Proceedings of the 1st Experimental Chaos Conference, edited by S. Vohra, M. Spano, M.
Shlesinger, L. Pecora, and W. Ditto (World Scientific, Singapore), p. 145.
Beck, C , 1990, Physica D 41, 67.
Benettin, G., C. Froeschle, and J.-P. Scheidecker, 1979, Phys.
Rev. A 19, 2454.
Benettin, G., L. Galgani, A. Giorgilli, and J.-M. Strelcyn, 1980,
Meccanica 15, 9 (Part I); 15, 21 (Part II).
Benettin, G., L. Galgani, and J.-M. Strelcyn, 1976, Phys. Rev. A
14,2338.
Biemond, J., R. L. Lagendijk, and R. M. Mersereau, 1990, Proc.
IEEE 78, 856.
Birman, J. S., and R. F. Williams, 1983a, Topology 22, 47.
Birman, J. S., and R. F. Williams, 1983b, Contemp. Math. 20, 1.
Bradley, E., 1991, in Algebraic Computation in Control, Lecture
Notes in Control and Information Sciences, edited by G. Jacob
and F. Lamnabhi-Lagarrigue (Springer, Berlin), p. 307.
Bradley, E., 1992, in Proceedings of the 1st Experimental Chaos
Conference, edited by S. Vohra, M. Spano, M. Shlesinger, L.
Pecora, and W. Ditto (World Scientific, Singapore), p. 152.
Briggs, K., 1990, Phys. Lett. A 151, 27.
Broomhead, D. S., and R. Jones, 1989, Proc. R. Soc. London,
Series A 423, 103.
Broomhead, D. S., and G. P. King, 1986, Physica D 20, 217.
Brown, R., 1992, "Orthonormal Polynomials as Prediction
Functions in Arbitrary Phase Space Dimensions," preprint,
University of California, San Diego, INLS.
Brown, R., P. Bryant, and H. D. I. Abarbanel, 1991, Phys. Rev.
A 43, 2787.
Bryant, P., R. Brown, and H. D. I. Abarbanel, 1990, Phys. Rev.
Lett. 65, 1523.
Carroll, T. L., and L. M. Pecora, 1991, IEEE Trans. Circuits
Syst. 38, 453.
Carroll, T. L., and L. M. Pecora, 1992, "Cascading and Controlling Synchronized Chaotic Systems" preprint, Naval
Research Laboratory, Washington, D.C.
Casdagli, M., 1989, Physica D 35, 335.
Casdagli, M., D. Des Jardins, S. Eubank, J. D. Farmer, J. Gib-

Abarbanel et al.: Analysis of observed chaotic data


son, N . Hunter, and J. Theiler, 1991, "Nonlinear Modeling of
Chaotic Time Series: Theory and Application," Preprint, Los
Alamos National Laboratory N o . LA-UR-91-1637.
Cawley, R., and G.-H. Hsu, 1992, "Local-Geometric-Projection
Method for Noise Reduction in Chaotic Maps and Flows,"
Phys. Rev. A 46, 3057.
Chandrasekhar, S., 1961, Hydrodynamic
and
Hydromagnetic
Stability (Clarendon, Oxford).
Chatterjee, S., and A. S. Hadi, 1988, Sensitivity Analysis in
Linear Regression (Wiley, New York).
Chennaoui, A., K. Pawelzik, W. Liebert, H. G. Schuster, and
G. Pfister, 1990, Phys. Rev. A 41, 4151.
CofFman, K. G., W. D . McCormick, Z. Noszticzius, R. H.
Simoyi, and H. L. Swinney, 1987, J. Chem. Phys. 86, 119.
Cowan, J. D., and D . H . Sharp, 1988, Q. Rev. Biophysics 2, 365.
Crawford, J. D., 1991, Rev. Mod. Phys. 63, 991.
Crawford, J. D., and E. Knobloch, 1991, Annu. Rev. Fluid Dynamics 23, 341.
Cross, M . C , and P. C. Hohenberg, 1993, Rev. Mod. Phys. 65,
851.
Cremers, A. J., and A. Hiibler, 1986, Z. Naturforsch A 42, 797.
Cvitanovic, P., 1988, Phys. Rev. Lett. 61, 2729.
Devany, R., and Z. Nitecki, 1979, Commun. Math. Phys. 67,
137.
Ditto, W. L., S. N . Rauseo, and M. L. Spano, 1990, Phys. Rev.
Lett. 65, 3211.
Eckmann, J.-P., S. O. Kamphorst, D . Ruelle, and S. Ciliberto,
1986, Phys. Rev. A 34, 4971.
Eckmann, J.-P., and D . Ruelle, 1985, Rev. Mod. Phys. 57, 617.
Eisenhammer, T., A. Hiibler, N . Packard, J. A. S. Kelso, 1991,
Bio. Cybernetics 65, 107.
Elmer, S., A. R. Gallant, D . McGafFrey, and D. Nychka, 1991,
Phys. Lett. A 153, 357.
Essex, C , and Nerenberg, M. A. H., 1991, Proc. R. Soc. London, Series A 435, 287.
Farmer, J. D., E. Ott, and J. A. Yorke, 1983, Physica D 7, 153.
Farmer, J. D., and J. J. ("Sid") Sidorowich, 1987, Phys. Rev.
Lett. 59, 845.
Farmer, J. D., and J. J. ("Sid") Sidorowich, 1988, in Evolution,
Learning and Cognition, edited by Y.-C. Lee (World Scientific,
Singapore), p. 277.
Farmer, J. D., and J. J. ("Sid") Sidorwich, 1991, Physica D 47,
373.
Flepp, L., R. Holzner, E. Brun, M. Finardi, and R. Badii, 1996,
Phys. Rev. Lett. 67, 2244.
Ford, J. R., and W. R. Borland, W. R., 1988, Common Los
Alamos Mathematical Software Compendium, Document C I C
N o . 148 (Los Alamos National Laboratory, Los Alamos, New
Mexico).
Fraser, A. M., 1988, Ph.D. thesis (University of Texas, Austin).
Fraser, A. M., 1989a, I E E E Trans. Inf. Theory 35, 245.
Fraser, A. M., 1989b, Physica D 34, 391.
Fraser, A. M., and H. L. Swinney, 1986, Phys. Rev. A 33, 1134.
Frederickson, P., J. L. Kaplan, E. D . Yorke, and J. A. Yorke,
1983, J. DhT. Eq. 49, 185.
Gallager, R. G., 1968, Information Theory and Reliable Communication (Wiley, New York).
Gilmore, C , 1992, J. Econ. Behav. Organization (in press).
Giona, M., F . Lentini, and V. Cimagalli, 1991, Phys. Rev. A 44,
3496.
Goldberg, D . E., 1989, Genetic Algorithms in Search, Optimization, and Machine Learning (Addison-Wesley, Reading, MA).
Golub, G. H., and C. F . Van Loan, 1989, Matrix
Computations,
2nd ed. (Johns Hopkins, Baltimore, MD).
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1389

Gollub, J. P., and R. Ramshankar, 1991, in New Perspectives in


Turbulence, edited by L. Sirovich (Springer, New York), p.
165.
Grassberger, P., and I. Procaccia, 1983a, Phys. Rev. Lett. 50,
346.
Grassberger, P., and I. Procaccia, 1983b, Physica D 9, 189.
Grassberger, P., T. Schrieber, and C. SchafFrath, 1991, Int. J.
Bif. Chaos 1,521.
Greene, J. M., and J.-S. Kim, 1986, Physica D 24, 213.
Guckenheimer, J., and P. Holmes, 1983, Nonlinear
Oscillations,
Dynamical
Systems,
and Bifurcations
of Vector
Fields
(Springer, New York).
Hammel, S. M., 1990, Phys. Lett. A 148, 421.
Hammel, S. M., C. K. R. T. Jones, and J. V. Moloney, 1985, J.
Opt. Soc. Am. B 2 , 552.
Hao, Bai-Lin, 1984, Chaos (World Scientific, Singapore).
Hao, Bai-Lin, 1990, Chaos II (World Scientific, Singapore).
Henon, M., 1976, Commun. Math. Phys. 50, 69.
Horn, R. A., and C. R. Johnson, 1985, Matrix Analysis (Cambridge University Press, New York).
Hsu, C. S., 1987, Cell to Cell Mapping, Applied Mathematical
Sciences No. 64 (Springer, New York).
Ikeda, K., 1979, Opt. Commun. 30, 257.
Isabelle, S. H., A. V. Oppenheim, and G. W. Wornell, 1992, in
7992 International Conference on Acoustics, Speech, and Signal
Processing, I E E E Signal Processing Society Volume 4 (IEEE,
New Jersey), p. IV-133.
Kaplan, D. T., and L. Glass, 1992, Phys. Rev. Lett. 68, 427.
Kaplan, J. L., and J. A. Yorke, 1979, in Functional
Differential
Equations and Approximations of Fixed Points, Lecture Notes
in Mathematics N o . 730, edited by H.-O. Peitgen and H-O.
Walther (Springer, Berlin), p. 228.
Keeler, J. D., and J. D . Farmer, 1986, Physica D 23, 413.
Kennel, M. B., R. Brown, and H. D . I. Abarbanel, 1992, Phys.
Rev. A 45, 3403.
Kolmogorov, A. N., 1958, Dokl. Akad. Nauk. SSSR 119, 861
[Math. Rev. 21, 386 (I960)].
Kostelich, E. J., and J. A. Yorke, 1991, Physica D 41, 183.
Kowalski, J. M., G. L. Albert, and G. W. Gross, 1990, Phys.
Rev. A 42, 6260.
Kuramoto, Y., 1984, Chemical Oscillations, Waves and Turbulence, Springer Series in Synergetics Vol. 19 (Springer, New
York).
Landa, P. S., and M. G. Rozenblyum (Rosenblum), 1989, Zh.
Tekh. Fiz. 59, 13 [Sov. Phys. Tech. Phys. 34, 6].
Lapedes, A., and R. Farber, 1987, "Nonlinear signal processing
using neutral networks: Prediction and signal modeling," Preprint, Los Alamos National Laboratory N o . LA-UR-87-2662.
Liebert, W., and H. G. Schuster, 1989, Phys. Lett. A 142, 107.
Lorenz, E. N., 1963, J. Atmos. Sci. 20, 130.
Lorenz, E. N., 1969, J. Atmos. Sci. 26, 636.
MacKay, R. S., 1992, "Hyperbolic Structure in Classical
Chaos," University of Warwick Scientific Preprint, 4/1993.
This paper was presented as a review Lecture at the International Summer School on Quantum Chaos, Enrico Fermi Institute, Varenna, Italy, 1992.
Mane, R., 1981, in Dynamical
Systems and
Turbulence,
Warwick, 1980, edited by D . Rand and L.-S. Young, Lecture
Notes in Mathematics N o . 898 (Springer, Berlin), p. 230.
Manneville, P., 1990, Dissipative Structures and Weak Turbulence (Academic, Boston).
Marteau, P.-F., and H. D . I. Abarbanel, 1991, J. Nonlinear Sci.
1,313.
McGafFrey, D . F., S. Ellner, A. R. Gallant, and D . W. Nychka,

1390

Abarbanel et aL: Analysis of observed chaotic data

1992, J. Am. Stat. Assoc. 87, 682.


Melvin, P., and N . B. Tufillaro, 1991, Phys. Rev. A 44, R3419.
Mindlin, G. B., X.-J. Hou, H. G. Solari, R. Gilmore, and N . B.
Tufillaro, 1990, Phys. Rev. Lett. 64, 2350.
Mindlin, G. B., H . G. Solari, M. A. Natiello, R. Gilmore, and
X.-J. Hou, 1991, J. Nonlinear Sci. 1, 147.
Mitschke, F., M. Moller, and W. Lange, 1988, Phys. Rev. A 37,
4518.
Oppenheim, A. V., and R. W. Schafer, 1989, Discrete-Time Signal Processing (Prentice Hall, Englewood Cliffs, NJ).
Oppenheim, A. V., G. W. Wornell, S. H. Isabelle, and K. M.
Cuomo, 1992, in International Conference on Acoustics, Speech,
and Signal Processing, I E E E Signal Processing Society Vol. 4
(IEEE, New Jersey), p. IV-117.
Orayevsky, , et aL, 1965, Molecular Oscillators (Nauka, Moscow).
Osborne, A. R., and A. Provenzale, 1989, Physica D 35, 357.
Oseledec, V. I., 1968, Trudy. Mosk. Mat. Obsc. 19, 197 [Moscow Math. Soc. 19, 197 (1968)].
Ott, E., C. Grebogi, and J. A. Yorke, 1990, Phys. Rev. Lett. 64,
1196.
Packard, N . H., 1990, Complex Systems 4, 543.
Packard, N . H., J. P. Crutchfield, J. D . Farmer, and R. S. Shaw,
1980, Phys. Rev. Lett. 45, 712.
Paladin, G., and A. Vulpiani, 1987, Phys. Rep. 156, 147.
Papoff, F., A. Fioretti, E. Arimondo, G. B. Mindlin, H. Solari,
and R. Gilmore, 1992, Phys. Rev. Lett. 68, 1128.
Parlitz, U., 1992, Int. J. Bif. Chaos 2, 155.
Pecora, L. M., and T. L. Carroll, 1990, Phys. Rev. Lett. 64, 821.
Pecora, L. M., and T. L. Carroll, 1991a, Phys. Rev. Lett. 67,
945.
Pecora, L. M., and T. L. Carroll, 1991b, Phys. Rev. A 44, 2374.
Pesin, Y. B., 1977, Usp. Mat. Nauk. 32, 55 [Russ. Math. Sur. 32,
55 (1977)].
Pikovsky, A. S., 1986, Radio Eng. Electron. Phys. 9, 81.
Powell, M . J. D., 1981, Approximation
Theory and Methods
(Cambridge University Press, Cambridge).
Powell, M. J. D., 1985, "Radial basis functions for multivariate
interpolation: a review" preprint, University of Cambridge.
Powell, M. J. D., 1987, "Radial basis function approximation to
polynomials" preprint, Univ. of Cambridge.
Rabinovich, M. I., 1978, Usp. Fiz. Nauk. 125, 123 [Sov. Phys.Usp. 21,443(1979)].
Rissanen, J., 1989, Stochastic Complexity in Statistical
Inquiry
(World Scientific, Singapore).
Roux, J . - C , R. H . Simoyi, and H. L. Swinney, 1983, Physica D
8, 257.
Ruelle, D., "Ergodic Theory of Differentiate Dynamical Systems," 1979, Publications Mathematiques of the Institut des
Hautes Etudes Scientifiques, N o . 50, p. 27.
Ruelle, D., 1990, Proc. R. Soc. London, Series A 427, 241.
Ruelle, J.-P., 1978, Bol. Soc. Bras. Mat. 9, 83.
Sakai, H., and H. Tokumaru, 1980, I E E E Trans. AASP 28, 588.
Salzmann, B., 1962, J. Atmos. Sci. 19, 329.
Sano, M., and Y. Sawada, 1985, Phys. Rev. Lett. 55, 1082.
Sauer, T., 1992, Physica D 58, 193.
Sauer, T., J. A. Yorke, and M . Casdagli, 1991, J. Stat. Phys. 65,
579.
Schreiber, T., and P. Grassberger, 1991, Phys. Lett. A 160, 411.
Shaw, R. Z., 1981, Naturforsch A 36, 80.
Shaw, R., 1984, The Dripping Faucet as a Model Dynamical System (Aerial Press, Santa Cruz).
Shimada, I., and T. Nagashima, 1979, Prog. Theor. Phys. 61,
1605.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

Shinbrot, T., W. Ditto, C. Grebogi, E. Ott, M . Spano, and J. A.


Yorke, 1992, Phys. Rev. Lett. 68, 2863.
Shinbrot, T., E. Ott, C. Grebogi, and J. A. Yorke, 1990, Phys.
Rev. Lett. 65, 3215.
Silverman, B. W., 1986, Density Estimation for Statistics and
Data Analysis (Chapman and Hall, London).
Sinai, Y., 1959, Dokl. Akad. Nauk. SSSR 124, 768 [Math. Rev.
21,386(1960)].
Singer, J., and H. H. Bau, 1991, Phys. Fluids A 3, 2859.
Singer, J., Y.-Z. Wang, and H. H. Bau, 1991, Phys. Rev. Lett.
66,1123.
Smith, L. A., 1988, Phys. Lett. A 133, 283.
Stoer, J., and R. Burlisch, 1980, Introduction
to Numerical
Analysis, translated by R. Bartels, W. Gautschi, and C.
Witzgall (Springer, New York).
Takens, F., 1981, in Dynamical
Systems and
Turbulence,
Warwick, 1980, edited by D . Rand and L.-S. Young, Lecture
Notes in Mathematics N o . 898 (Springer, Berlin), p. 366.
Theiler, J., 1990, J. Opt. Soc. Am. A 7, 1055.
Theiler, J., 1991, Phys. Lett. A 155, 480.
Tufillaro, N . B., R. Holzner, L. Flepp, E. Brun, M. Finardi, and
R. Badii, 1991, Phys. Rev. A 44, R4786.
Van Trees, H . L., 1968, Detection, Estimation, and Modulation
Theory, Part I (Wiley, New York).
Whitney, H., 1936, Ann. Math. 37, 645.
Wolf, A., J. B. Swift, H . L. Swinney, and J. A. Vastano, 1985,
Physica D 16, 285.

ADDITIONAL REFERENCES
One of the challenges in a review such as this one is making
it finite. In rising to this challenge we have had to make a
selection of topics and to identify the appropriate literature
that bears quite directly on those topics. We have, of course,
done this, and thus risked ignoring vast parts of the vast literature on the analysis of time series from chaotic observations.
The items cited in the main body of the text by no means
represent the same selection many of our colleagues would
have made, but they do represent the literature we consulted
and learned from in putting together this review. Many other
papers and books have contributed to our understanding of the
material we are covering in this review, and we attempt to
identify below the main body of these. As ever, we are pleased
to hear from authors whose work we have inadvertently failed
to cite. We shall be happy to read these papers and we shall
happily cite that work as well.
Abarbanel, H. D . I., R. Brown, and M. B. Kennel, 1991,
"Lyapunov Exponents in Chaotic Systems: Their Importance
and Their Evaluation using Observed Data," Int. J. Mod.
Phys. B 5, 1347.
Abarbanel, H. D . I., S. Koonin, H. Levine, G. J. F . MacDonald,
and O. Rothaus, 1990a, "Statistics of Extreme Events with Application to Climate," J A S O N / M I T R E Report JSR-90-305.
Abarbanel, H. D . I., S. Koonin, H. Levine, G. J. F . MacDonald,
and O. Rothaus,
1990b, "Issues in
Predictability,"
J A S O N / M I T R E Report JSR-90-320.
Afraimovich, V. S., I. S. Aranson, and M . I. Rabinovich, 1989,
"Multidimensional Strange Attractors and Turbulence," Sov.
Sci. C Math. Phys. 8, 3.
Afraimovich, V. S., M. I. Rabinovich, and V. I. Sbitnev, 1985,
"Attractor dimension in a chain of coupled oscillators," Sov.
Tech. Phys. Lett. 11, 139.
Albano, A. M., J. Muench, C. Schwartz, A. Mees, and P. R a p p ,

Abarbanel et al.: Analysis of observed chaotic data


1988, "Singular Value Decomposition and the GrassbergerProcaccia Algorithm," Phys. Rev. A 38, 3017.
Aranson, I. S., M . I. Rabinovich, A. M. Reyman, V. G. Shekhov, 1988, "Measurement of Correlation Dimension in Experimental Devices,'' Preprint No. 215, Institute of Applied
Physics, Gorky.
Baake, E., M. Baake, H. G. Bock, and K. M. Briggs, 1992, "Fitting Ordinary Differential Equations to Chaotic D a t a , " Phys.
Rev. A 45, 5524.
Badii, R., G. Broggi, B. Derighetti, M. Ravani, S. Ciliberto, A.
Politi, and M. A. Rubio, 1988, "Dimension Increase in Filtered Chaotic Signals," Phys. Rev. Lett. 60, 979.
Baker, G. L., and J. P. Gollub, 1990, Chaotic Dynamics: An Introduction (Cambridge University Press, New York).
Bayly, B. J., I. Goldhirsch, and S. A. Orzag, 1987, "Independent
degrees of freedom of dynamical systems," J. Scientific Computing 2, 111.
Benzi, R., and G. F. Carnevale, 1989, " A Possible Measure of
Local Predictability," J. Atmos. Sci. 46, 3595.
Bingham, S., and M. K o t , 1989, "Multidimensional Trees,
Range Searching, and a Correlation Dimension Algorithm of
Reduced Complexity," Phys. Lett. A 140, 327.
Buzug, Th., T. Reimers, and G. Pfister, 1990, "Optimal Reconstruction of Strange Attractors from Purely Geometrical Arguments," Europhys. Lett. 13, 605.
Cenys, A., and K. Pyragas, 1988, "Estimation of the N u m b e r of
Degrees of Freedom from Chaotic Time Series," Phys. Lett. A,
129, 227.
Cenys, A., G. Lasiene, and K. Pyragas, 1991, "Estimation of Interrelations between Chaotic Observables," Physica D 52, 332.
Ciliberto, S. and B. Nicolaenko, 1991, "Estimating the Number
of Degrees of Freedom in Spatially Extended Systems," Europhys. Lett. 14, 303.
Crutchfield, J. P., and B. S. M a c N a m a r a , 1987, Equations of
motion from a data series. Complex Systems 1, 417.
Ellner, S., 1988, "Estimating Attractor Dimensions from Limited Data: A New Method, with Error Estimates," Phys. Lett.
A 133, 128.
Farmer, J. D., E. Ott, and J. A. Yorke, 1983, "The Dimension
of Chaotic Attractors," Physica D 7, 153.
Farmer, J. D., 1982, "Chaotic Attractors of an InfiniteDimensional Dynamical System," Physica D 4, 366.
Farmer, J. D., 1990, "Rosetta stone for connectionism," Physica D 42, 153.
Fraedrich, K., 1986, "Estimating the Dimensions of Weather
and Climate Attractors," J. Atmos. Sci. 43, 419.
Frederickson, P., J. L. Kaplan, E. D . Yorke, and J. A. Yorke,
1983, "The Liapunov Dimension of Strange Attractors," J.
Diff. Eq. 49, 185.
Friedman, J. H., J. L. Bentley, and R. A. Finkel, 1977, " A n Algorithm for Finding Best Matches in Logarithmic Expected
Time," A C M Trans. Math. Software 3, 209.
Fujisaka, H., 1983, "Statistical Dynamics Generated by Fluctuations of Local Lyapunov Exponents," Prog. Theor. Phys.
70, 1274.
Geist, K., U. Parlitz, and W. Lauterborn, 1990, "Comparison of
Different Methods for Computing Lyapunov Exponents,"
Prog. Theor. Phys. 83, 875.
Goldhirsch, I., P-L. Sulem, and S. A. Orzag, 1986, "Stability
and Lyapunov Stability of Dynamical Systems: a Differential
Approach and Numerical Method," Physica D 27, 311.
Grassberger, P., R. Badii, and A. Politi, 1988, "Scaling Laws for
Invariant Measures on Hyperbolic and Nonhyperbolic Attractors," J. Stat. Phys. 51, 135.
Rev. Mod. Phys., Vol. 65, No. 4, October 1993

1391

Grebogi, C , E. Ott, and J. A. Yorke, 1988, "Unstable Periodic


Orbits and the Dimensions of Multifractal Chaotic Attractors," Phys. Rev. A 37, 1711.
Hou, X. J., R. Gilmore, G. B. Mindlin, and H. G. Solari, 1990,
" A n Efficient Algorithm for Fast 0(Nlog(N))
Box Counting,"
Phys. Lett. A 151, 43.
Kaplan, J. L., and J. A. Yorke, 1979, "Chaotic Behavior in
Multidimensional Difference Equations," Lect. Notes M a t h .
730,204.
Kostelich, E. J., and J. A. Yorke, 1988, "Noise reduction in
dynamical systems," Phys. Rev. A 38, 1649.
Landa, P. S., and M. Rosemblum, 1991, "Time Series Analysis
for System Identification and Diagnostics," Physica D 48, 232.
Liebert, W., K. Pawelzik, and H. G. Schuster, 1991, "Optimal
Embeddings of Chaotic Attractors from Topological Considerations," Europhys. Lett. 14, 521.
Lindenberg, K., and B. J. West, 1986, "The First, the Biggest,
and Other Such Considerations," J. Stat. Phys. 42, 201.
Lorenz, E. N., 1984, "Irregularity: A Fundamental Property of
the Atmosphere," Tellus A 36, 98.
Lukashchuk, S. N., G. E. Falkovich, and A. I. Chernykh, 1989,
"Calculation of the dimensions of attractors from experimental data," J. Appl. Mech. Tech. Phys. 30, 95.
Mackey, M., and L. Glass, 1977, "Oscillation and Chaos in
Physiological Control Systems," Science 197, 287.
Mayer-Kress, G., 1986, Ed., Dimensions and Entropies in Chaotic Systems (Springer, Berlin).
Mees, A. I., P. E. R a p p , and L. S. Jennings, 1987, "Singularvalue Decomposition and Embedding Dimension," Phys. Rev.
A 37, 340.
Moon, Francis C , 1987, Chaotic Vibrations (Wiley, New York).
Myers, C , S. Kay, and M. D . Richard, 1992, "Signal Separation
for Nonlinear Dynamical Systems," in International
Conference on Acoustics, Speech, and Signal Processing, I E E E Signal
Processing Society, Vol. 4 (IEEE, New Jersey), p. 129.
Nese, Jon M., 1989, "Quantifying Local Predictability in Phase
Space," Physica D 35, 237. Also see his P h D thesis entitled
"Predictability of Weather and Climate in a Coupled-Ocean
Atmosphere Model: A Dynamical Systems A p p r o a c h , " The
Pennsylvania State University, Department of Meteorology,
1989, and references therein.
Osborne, A. R., and A. Provenzale, 1989, "Finite Correlation
Dimension for Stochastic Systems with Power-Law Spectra,"
Physica D 35, 357.
Parker, T. S., and L. O. Chua, 1990, Practical Numerical Algorithms for Chaotic Systems (Springer, New York).
Pawelzik, K., and H. G. Schuster, 1987, "Generalized Dimensions and Entropies from a Measured Time Series," Phys. Rev.
A 35, 481.
Pyragas, K., and A. Cenys, 1987, " M e t h o d for the Rapid Determination of the Number of Active Degrees of Freedom of Dynamic Systems with Chaotic Behavior," Litovskii Fizicheskii
Sbornik 27, 437.
Reyl, C , L. Flepp, R. Badii, and E. Brun, 1993, "Control of
NMR-Laser Chaos in High-Dimensional Embedding Space,"
Phys. Rev. E 47, 267.
Rossler, O. E., 1976, " A n Equation for Continuous Chaos,"
Phys. Lett. A 57, 397.
Russell, D . A., J. D . Hanson, and E. Ott, 1980, "Dimension of
Strange Attractors," Phys. Rev. Lett. 45, 1175.
Sato, S., M . Sano, and Y. Sawada, 1987, "Practical Methods of
Measuring the Generalized Dimension and Largest Lyapunov
Exponents in High-Dimensional Chaotic Systems," Prog.
Theor. Phys. 77.

1392

Abarbanel et a/.: Analysis of observed chaotic data

Smith, L. A., 1988, "Intrinsic Limits on Dimension Calculations," Phys. Lett. A 133, 283.
Stoop, R., and J. Parisi, 1991, "Calculation of Lyapunov Exponents Avoiding Spurious Elements," Physica D 50, 89.
Theiler, J., 1986, "Spurious Dimension from Correlation Algorithms Applied to Limited Time Series Data," Phys. Rev. A
34, 2427.
Theiler, J., 1987, "Efficient Algorithm for Estimating the Correlation Dimension from a Set of Discrete Points," Phys. Rev. A
36, 4456.
Theiler, J., 1988, "Lacunarity in a Best Estimator of Fractal Dimension," Phys. Lett. A 133, 195.
Theiler, J., 1990, "Statistical Precision of Dimension Estimators," Phys. Rev. A 41, 3038.
Thompson, J. M. T., and H. B. Stewart, 1986, Nonlinear Dy-

Rev. Mod. Phys., Vol. 65, No. 4, October 1993

namics and Chaos (Wiley, New York).


Tufillaro, N. B., H. G. Solari, and R. Gilmore, 1990, "Relative
Rotation RatesFingerprints for Strange Attractors," Phys.
Rev. A 41, 5717.
Wales, D. J., 1991, "Calculating the Rate of Loss of Information from Chaotic Time Series by Forecasting," Nature 350,
485.
Wright, Jon, 1984, "Method for Calculating a Lyapunov exponent," Phys. Rev. A 29, 2924.
Young, L-S., 1983, "Entropy, Lyapunov Exponents, and Hausdorff Dimension in Differentiable Dynamical Systems," IEEE
Trans. Circuits Systems C AS-30, 599.
Zeng, X., R. Eykholt, and R. A. Pielke, 1991, "Estimating the
Lyapunov Exponent Spectrum from Short Time Series of Low
Precision," Phys. Rev. Lett. 66, 3229.

You might also like