Vous êtes sur la page 1sur 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/231068546

Instant chaos

Article  in  Nonlinearity · January 1999


DOI: 10.1088/0951-7715/5/6/001 · Source: CiteSeer

CITATIONS READS
25 57

2 authors, including:

John Guckenheimer
Cornell University
246 PUBLICATIONS   34,911 CITATIONS   

SEE PROFILE

All content following this page was uploaded by John Guckenheimer on 08 March 2016.

The user has requested enhancement of the downloaded file.


Instant Chaos
John Guckenheimer and Patrick Worfolk
Center for Applied Mathematics
Cornell University, Ithaca, NY 14853
June 29, 1992
Abstract
One of the predominant themes of nonlinear dynamics during the
past twenty years has been the characterization of the \routes to
chaos". Within generic families of vector elds, there are ubiquitous
patterns that describe how qualitative properties of vector elds can
change with a varying parameter. Bifurcation theory classi es these
patterns, and they form the substrate for determining the routes to
chaos that one expects to see in physical systems. The presence of
symmetry typically alters the patterns that one observes, and it has
been well understood that symmetry can lead to the persistence of
new types of dynamical behavior and bifurcation. This paper gives
another, more extreme, example of how symmetry can a ect the routes
to chaos. In the systems that we describe below, there are persistent
bifurcations that lead directly from a \trivial" steady state to chaotic
attractors of small amplitude. These bifurcations are \supercritical"
in the sense that the attractors emerge from the bifurcating equi-
librium and remain con ned to arbitrarily small neighborhoods of the
equilibrium for small values of the bifurcation parameter. Our analysis
relies upon studying the global properties of a four parameter family
of four dimensional vector elds which have invariant manifolds close
to the unit sphere in their four dimensional phase spaces. A bewil-
dering variety of dynamical behavior can be found in this family of
vector elds, all of which represents possible post-bifurcation behavior
of the original problem. This analysis indicates that there are intrinsic
limits to our ability to classify bifurcations and characterize the routes
to chaos through algebraic calculations.

1 Introduction
Within generic families of vector elds, there are ubiquitous patterns that
describe how qualitative properties of vector elds can change with a varying
1
parameter. The term routes to chaos [4] describes those patterns that char-
acterize the way generic families of dynamical systems proceed from simple
attracting states to chaotic ones. The classi cation of Eckmann [4] is ap-
pealing in its simplicity, though it is far from complete. For it to apply, a
number of assumptions must be satis ed. When the assumptions fail, other
things can happen - sometimes in a context that leads to interesting alterna-
tive classi cations. The presence of symmetry in a physical problem is one
means of violating the assumptions underlying the simple routes to chaos.
An extensive theory aimed at classifying generic bifurcations in systems with
symmetry has been developed by Golubitsky et al. [7]. Techniques based
upon group theory and singularity theory lead to the formulation of \nor-
mal forms" for bifurcation problems with symmetry and to the calculation
of some branches of bifurcating solutions. Each symmetry group leads to its
associated types of generic bifurcation. In many interesting examples that
are physically important, this analysis leads to a complete classi cation of
the bifurcations into one of just a few types and one recovers a classi cation
akin to that of the asymmetric situation. At the same time, there has been a
realization that the methods based upon singularity theory and group theory
produce incomplete answers concerning the qualitative properties associated
with symmetric bifurcations. This paper presents an example that shows
substantially more complex bifurcation behavior for symmetric systems than
has been exhibited heretofore.
The symmetry group that we work with is a subgroup G of O(4) with 32
elements. It is most easily described as a subgroup of index 2 for the group
G generated by the matrices
0 ?1 0 0 0 1
BB 0 1 0 0 CC
=B
@ 0 0 1 0 CA
0 0 0 1
and 00 1 0 01
BB 0 0 1 0 CC
=B
@ 0 0 0 1 CA
1 0 0 0
The subgroup G  G consists of those matrices in G that preserve orientation
(i.e., have positive determinant). The group G has 32 elements and is given
2
by products of  and  with an even number of terms. We study families of
vector elds that are equivariant with respect to the group G. It is readily
checked that the action of G is xed point free and absolutely irreducible: the
only matrices that commute with the action of G are scalar multiples of the
identity matrix. This implies that the origin is always an equilibrium point for
a G-equivariant vector eld and that the derivative of a G-equivariant vector
eld is a multiple of the identity matrix. Thus, if a one parameter family of
vector elds bifurcates at the origin with a zero eigenvalue, the derivative of
the vector eld is identically zero there. The null space associated with the
bifurcation is all of R . 4

In accord with the theory of normal forms of vector elds, we seek to


understand bifurcations at the origin (also called the trivial equilibrium),
by expanding the Taylor series of a family of vector elds at the origin and
truncating this Taylor expansion at an appropriate degree. Using results
about structural stability and persistence of dynamical structures, we seek
to prove that dynamical behavior found in the truncated system is present in
generic equivariant families undergoing bifurcation at the origin. The starting
point for our dynamical studies is therefore the Taylor series of degree 3 for a
family of equivariant vector elds bifurcating at the trivial equilibrium. This
has the form
x_ = (l + ar + bx + cx + dx )x + ex x x
1
2 2 2 2
1 2 3 4
x_ = (l + ar + bx + cx + dx )x ? ex x x
2 3 4
2 2 2 2
2 3
x_ = (l + ar + bx + cx + dx )x + ex x x
2 2
4
2
1
2
2 1
(1)3 4

3 4 1 2 3 1 2 4
x_ = (l + ar + bx + cx + dx )x ? ex x x
4
2 2
1
2
2
2
3 4 1 2 3

with
=x +x +x +x :
r2 2
1
2
2
2
3
2
4

Here (x ; x ; x ; x ) 2 R and l; a; b; c; d; e are parameters with l the parame-


1 2 3 4
4

ter which induces bifurcation at the origin. This system of di erential equa-
tions, which we call the basic system, is the object of study for the remainder
of this paper.
The study of equivariance with respect to G was motivated by the work
of Field and Swift [6]. They examine equivariance with respect to G and
the corresponding cubic truncation, which is Equations 1 with e = 0. Our
work extends theirs in continuing to characterize persistent bifurcations in
the presence of symmetry. The interested reader should refer to [6] for a
3
relevant discussion of equivariant bifurcation theory that is more detailed
than the one given here.

2 Isotropy Groups and Invariant Subspaces


This section discusses the algebraic and group theoretic structures that are
associated with the action of G on R . If x 2 R , denote the G-orbit of
4 4

x by Gx = fgxjg 2 Gg. Observe that if x is an equilibrium point of a G-


equivariant vector eld then elements of the orbit of x are also equilibrium
points, since f (gx) = gf (x). Hence the equilibrium set is a union of orbits
and we shall sometimes identify equilibrium points in an orbit by studying
the quotient space of R by the action of G. The subgroup of G which xes
4

x is called the isotropy subgroup of G at x, x = fg 2 Gjgx = xg. Since


points in the orbit of x have conjugate isotropy subgroups (gx = gxg? ), 1

we will be interested in conjugacy classes of isotropy subgroups, [x]. The


set of conjugacy classes of isotropy subgroups of G is nite and can be given
a partial ordering based on inclusion. This ordering is called the isotropy
lattice of G acting on R and plays an important role in organizing both the
4

dynamics of the system and the calculations we perform. Figure 1 is the


isotropy lattice for our group G acting on R .4

The xed-point subspace of a subgroup H of G is the subspace of points


x 2 R which are xed by the action of H , F ix(H ) = fx 2 R jhx =
4 4

x for all h 2 H g. These subspaces are important because they are invariant
under the ow induced by a G-equivariant vector eld. Conjugate subgroups
have xed-point subspaces which are the image of each other under a group
element. As in the case of equilibrium points, we shall consider xed-point
subspaces in the same \orbit" to be the same. Representative xed-point
subspaces for the isotropy subgroups of G are also given in Figure 1.

3 Dynamics on an Attracting 3-Sphere


This section discusses geometric aspects of the basic system. We set the
parameters l = 1 and a = ?1 and then investigate the behavior of the
system for small values of the remaining parameters b; c; d; e. In this regime,
the basic system has an invariant three dimensional manifold that lies close

4
to the unit sphere in R [5]. As the parameters in this four parameter family
4

are varied, a multitude of bifurcations take place on this three dimensional


sphere. In this paper, we identify two global bifurcations, a period doubling
cascade and a homoclinic bifurcation of Silnikov type, that are associated
with the presence of chaotic invariant sets and horseshoes. If one accepts the
validity of the numerical computations establishing the existence of these
bifurcations, then the theory associated with these bifurcations proves the
existence of chaotic invariant sets in the ow of the basic system for open
regions within the parameter space.
The parameter a may be set to ?1 by transforming the basic system with
a di eomorphism given by scaling the phase space variables by jaj (and,
? 1
2

possibly, a time reversal). We now state the theorem which is the foundation
of our analysis.
Theorem 1 There is a  > 0 such that if l > 0, a = ?1, jbj < , jcj < 
and jdj < , then the basic vector eld has an invariant attracting three
dimensional
p manifold S , a topological sphere, close to the sphere of radius
l.

This result follows directly from Field's Invariant Sphere Theorem [5] and
the observation that for l > 0; a = ?1; b = cp= d = 0, the basic vector eld
has an invariant attracting sphere of radius l.
p
Scaling the phase space variables by l and the time variable by l gives
a set of di eomorphisms that transform the basic system with l > 0 to the
basic system with l = 1 and the same values of the parameters a; b; c; d; e.
Thus the dynamics of the basic system for l > 0 changes only by similarity
transformations. Thus we x l = 1; a = ?1 and vary b; c; d; e, seeking ranges
of these parameters that produce chaotic behavior. Fixing the parameter
values of b; c; d; e and varying l gives bifurcation problems in which there is
instant chaos. By this we mean that chaotic behavior appears in an arbi-
trarily small neighborhood of the origin immediately after the bifurcation of
the trivial equilibrium.
The manifold S from the invariant sphere theorem is globally attracting in
the sense that every trajectory which does not contain the trivial equilibrium
is forward asymptotic to S . Consequently, we will concentrate on studying
the dynamics on S . While the geometry of the ows on S produced by the
5
basic system depends upon the parameters b; c; d; e, these ows have several
common features. These properties are immediate consequences of the basic
equations and are derived from an examination of the isotropy lattice. First,
the coordinate axes are invariant, so there are equilibria at their intersections
with S , the unit vectors. Second, the coordinate planes are invariant, so the
curves of their intersection with S are invariant. Note that the coordinate
hyperplanes are not invariant when e 6= 0. (See Field and Swift [6] for an
analysis when e = 0.) In addition to the coordinate planes, the planes de ned
by x = x , x = x are invariant resulting in more invariant curves in
1 3 2 4
S. These planes also contain invariant one-dimensional subspaces de ned by
x = x , x = x = 0 and x = x , x = x = 0. The intersection of these
1 3 2 4 2 4 1 3
lines with S give equilibria lying on the aforementioned invariant curves.
Bezout's Theorem implies that the number of equilibrium points for the
basic equations is at most 81. When b = c = d = 1 and e = 0, the basic
equations separate into equations for each xi, independent of the remaining
variables. For these parameter values, there are 81 equilibria that occur at
the points (e ; e ; e ; e ) where ei 2 f?1; 0; 1g. In the parameter regime we
1 2 3 4
study, nding parameter values with 81 equilibria requires ingenuity. Our
e orts to use Grobner basis and invariant theory techniques to nd explicit
formulas for locating all of the equilibrium points have not been completely
successful.
All equilibrium points except the trivial one lie on the invariant manifold
S . There are six di erent classes of equilibrium points distinguished by their
isotropy groups. Equilibria from three of these classes exist throughout the
parameter regime we study here. These are the trivial equilibrium and the
two sets of equilibria previously mentioned: the 8 equilibria on the coordinate
axes with isotropy group conjugate to (Z )0 and the 8 on the invariant lines
3
2
with isotropy group conjugate to Z . The existence of the solutions in the
2
2
three remaining classes depends upon the particular choice of parameter
values. These consist of a possible 16 equilibrium points with isotropy group
conjugate to (Z )0, 16 with isotropy group conjugate to Z , and 32 with trivial
2
2 2
isotropy group. The di erent classes of equilibrium points and representative
members are tabulated in Figure 2.
There are parameter regimes (bd < 0) in which there are stable het-
eroclinic cycles formed from the intersections of the manifold S with the
coordinate planes (1; 2); (2; 3); (3; 4) and (4; 1). One can picture these cycles
as being the analog of the curves cut out on the two-dimensional sphere by
6
the equator and two mutually orthogonal lines of longitude. Parameter val-
ues with b > 0; c; d < 0 can be found so that these cycles are attracting in
the sense that any trajectory which enters a neighborhood of the cycle will
be forward asymptotic to the cycle. Each of the equilibrium points on the
coordinate axes is a saddle with three stable eigenvalues and one unstable
eigenvalue. Thus, as one approaches one of the saddle points in the cycle,
there is no choice as to the axis along which one moves away from the sad-
dle, but there is a choice of orientation. The chaotic behavior that we shall
describe can be regarded as lying in the neighborhood of these heteroclinic
cycles, and one can think of them in terms of the bifurcations that occur
along a homotopy that begins with parameter values for which there is a
stable heteroclinic cycle of the type described above.
The rst bifurcation from a state with stable heteroclinic cycles that we
describe involves bifurcations that take place in the invariant planes. The
equilibrium at (1; 0; 0; 0) has eigenvalues ?2; b; c; d. When d changes sign, a
pitchfork bifurcation that leads to the creation of equilibrium points in the
conjugacy class of (x ; x ; 0; 0) occurs in each invariant coordinate plane. The
1 2
Jacobian derivative at these points has a block structure, with two eigenvec-
tors lying in the invariant plane and a 2  2 block describing the stability
in the normal directions. As the parameter c is varied, bifurcations can take
place in which the real parts of eigenvalues associated with the behavior
normal to the invariant plane change sign. This can occur as a pitchfork
bifurcation, as a Hopf bifurcation, or (by varying a second parameter, e) as
a Bogdanov-Takens bifurcation with Z symmetry.
2
We are particularly interested in states in which the pitchfork bifurca-
tion has occurred and the eigenvalues for the 2  2 block of the Jacobian
representing behavior transversal to the invariant plane are a complex pair
with positive real parts. Label one equilibrium point in the invariant plane
p and the two equilibria which bifurcated from it q ; q . When we discuss
1 2
properties common to both q and q and do not wish to di erentiate be-
1 2
tween them, we will refer simply to the point q. The general pattern of the
ow, beginning at an equilibrium point on a coordinate axis, is to follow the
invariant curve in the direction of the unstable eigenvector with eigenvalue
b. The ow then reaches the vicinity of the equilibrium point p and follows
the two dimensional unstable manifold of p approaching the vicinity of q ; q .
1 2
There are many possibilities for the details of the ow at this point. Two of
these will be described here.
7
In the rst scenario, there is an attracting periodic orbit \surrounding"
the invariant curve containing p and the equilibria q ; q . This periodic orbit
1 2
has a Z symmetry and, as the parameter e is varied, undergoes a pitchfork
2
bifurcation creating two new attracting, asymmetric periodic orbits. These
asymmetric orbits are observed to undergo a period doubling cascade. De-
tails of this cascade are described in Section 4. In the second scenario, most
of the points in the unstable manifold of p follow branches of the unstable
manifolds of q ; q to another equilibrium point in the orbit of p. In this
1 2
scenario, parameter values are located with the property that q ; q have ho-
1 2
moclinic orbits when we look at the quotient of the ow with the action of
the symmetry group G. The points q ; q have a pair of complex stable eigen-
1 2
values and a real unstable eigenvalue in S . The unstable eigenvalue is larger
than the magnitude of the real part of the complex stable eigenvalues. These
properties characterize the homoclinic orbits as being of Silnikov type [8].
We describe in Section 5 some of the geometric consequences and numerical
arguments that lead to the conclusion that there is indeed a homoclinic orbit
of Silnikov type for some parameter values in this system.

4 Period Doubling Cascade


This section describes the numerical analysis which demonstrates the exis-
tence of a cascade of period doubling bifurcations in the basic system. The
starting point is a set of parameter values (b = 0:1; c = ?0:053; d = 0:02; e =
0:198) for which the ow contains a pair of asymmetric periodic orbits enclos-
ing the invariant curve which is the intersection of the (1; 2) coordinate plane
with S . These periodic orbits are the image of each other under rotation by
 in the (3; 4) coordinate plane and are only one pair of 32 periodic orbits
equivalent by the symmetry group. These orbits all undergo the same bifur-
cations simultaneously. We will study the bifurcations of just one periodic
orbit from this original pair by computing a Poincare map for the ow.
Consider the Poincare map in S de ned by the section x = 0. This
3
is a two dimensional di eomorphism and the periodic orbit being studied is
represented by a xed point. As the parameter e is varied, we can numerically
study the bifurcations of this xed point and follow it undergoing several
period doubling bifurcations. Further bifurcations are revealed by studying
the attracting set that bifurcates from the original xed point. Figure 3

8
shows the results of the numerical computation and is a one-dimensional
projection of part of the attracting set plotted for a range of values for e.
Several features of Figure 3 are signi cant. First, as e increases from the
left we observe a sequence of period doubling bifurcations. The parameter
values at which the bifurcations occur are accumulating at a rate proportional
to Feigenbaum's universal constant for one-dimensional period doubling [12].
To the right of the accumulation of period doubling bifurcations, there are
windows in the parameter values of stable periodic orbits. It is particularly
easy to observe those of periods 5 and 6. The windows of periodic behavior
are created in the order predicted by the one-dimensional theory. The win-
dows appear to be separated by regions where orbits either have long period
or are aperiodic. This is all consistent with the theory of one-dimensional
period doubling. It is easy to verify numerically that the Poincare map being
studied is essentially one-dimensional when the parameter e is close to the
regime of period doubling.
For parameter values corresponding to the windows of stable periodic
orbits, we note that the ow is structurally stable, since the structures we
are observing are hyperbolic. The universal theory tells us to expect co-
existing unstable periodic orbits of all periods lower in the Sarkovskii ordering
[3]. Consequently, we infer that the chaotic behavior found in the period
doubling cascade is robust. Note that the theory developed by Benedicks
and Carleson [2] and extended by Mora and Viana [9] also predicts that
there are positive measure sets of parameter values for which the system has
chaotic, nonhyperbolic attractors.
Finally, we conclude this discussion of the period doubling cascade with
two additional comments about the behavior of this system beyond the period
doubling regime. These observations are derived from our numerical analysis.
As the parameter e increases the attractor undergoes several \crises" which
are analogous to bifurcation in one-dimensional unimodal maps where the
critical point is mapped outside of the interval. This occurs before the full
one-dimensional picture has been realized; for example, the stable period 3
orbit is not produced. Recall that the attractors formed after the period
doubling cascade are asymmetric. As e increases further, there is a nal
symmetry increasing bifurcation as the two attractors merge and become
symmetric with respect to rotation by  in a coordinate plane.

9
5 Homoclinic Orbit of Silnikov type
This section describes how to observe that for suitable parameter values the
asymmetric equilibria, previously denoted q ; q , are part of a Silnikov con-
1 2
nection. By this, we mean that there are heteroclinic cycles containing q ; q 1 2
and their images which, when reduced by the action of the symmetry group
G, are equivalent to homoclinic orbits of S ilnikov type. To this end, we will
rst discuss further details of the geometry of the ow for this particular pa-
rameter regime. Then we will discuss our observations based on a numerical
analysis.
The ow is essentially characterized by the geometry of the invariant
manifolds associated to the equilibriumpoints in the orbits of p and q. Within
the invariant sphere S , the equilibrium point p has a one-dimensional stable
manifold lying in the (1; 2)-coordinate plane. The two unstable eigenvalues
are a complex conjugate pair and have eigenvectors forced by the symmetry
to lie in the (3; 4)-coordinate plane of the tangent space at p. The symmetry
also forces the unstable manifold to be symmetric with respect to rotation
by  in the (3; 4)-coordinate plane.
In contrast, the equilibrium point q has a two-dimensional stable manifold
with complex conjugate eigenvalues and a one-dimensional stable manifold.
The stable manifold of q intersects the unstable manifold of p. One branch
of the unstable manifold of q simply ows away approaching what used to
be the heteroclinic cycle containing the equilibrium points on the coordinate
axes until it enters the neighborhood of a point in the orbit of p. We shall call
this the escaping branch. The other branch turns towards p and approaches
the unstable manifold of p, reinjecting itself into the local ow. We will call
this the captured branch.
Consequently, a typical point in a neighborhood of p will undergo the
following behavior. First, it follows the unstable manifold of p and will
approach either q or q . If it is on one side of the stable manifold of q it
1 2
will follow the escaping branch and move onwards towards a new point in
the orbit of p. But if it is on the other side of the stable manifold of q, it will
follow the captured branch and be trapped and sent back towards p again
to try to escape the next time it ows out along p's unstable manifold. See
Figure 4.
Once again consider the escaping branch of the point q. This will ap-

10
proach an image of p, follow another unstable manifold and approach an
image of q, call it q~. It will either escape again or be trapped depend-
ing on the side of q~'s stable manifold from which it approached. If we
can vary a parameter in the system so that sometimes it approaches on
one side and other times approaches on the other side, then we can expect
that for some intermediate value it will lie exactly on the stable manifold.
This would form a heteroclinic connection between the points q and q~. It
is now a matter of computing the eigenvalues at q in order to verify that
we do indeed have a Silnikov connection. We search numerically for exactly
this crossing phenomenon. Figure 5 shows one branch of the stable mani-
fold of q corresponding to two slightly di erent values of the parameter e,
(b = 0:1; c = ?0:085; d = 0:005; e = 0:53440 and e = 0:53441). We observe
that the manifolds do in fact arrive on opposite sides of the stable manifold
of q~ since they follow di erent branches of the unstable manifold at q~. The
consequence is that for some intermediate value of e, there must be a connec-
tion. The eigenvalues of q are approximately 0:0422 and ?0:0196  0:0343i.
These satisfy the Silnikov condition that the magnitude of the real eigen-
value is larger than the real parts of the complex eigenvalues and hence we
conclude that our system does in fact contain a Silnikov connection.
We can now defer to Silnikov theory [8, 11] to make observations about the
ow in a neighborhood of this connection. We conclude that there exist Smale
horseshoes embedded in the ow for an open neighborhood of parameter
values containing the point of Silnikov connection. Hence, there is an open
region of parameter space where the ow is chaotic.

6 Remarks on Structural Stability


The numerical observations that are discussed in the two previous sections
form the basis of a \strong" conjecture that the phenomenon of \instant
chaos" occurs in the class of vector elds with the symmetry group G. We
formulate two theorems that are based upon the numerical computations
supporting these conjectures:
Theorem 2 Let G  O(4) be the group generated by p =  and s =  . 2

Assume that there are parameter values l = 1, a = ?1 and b; c; d; e for which


Equations 1 have a homoclinic orbit that satis es the Silnikov conditions [8].

11
Then there is an open region U in the space of one parameter families of
vector elds on R4 equivariant with respect to G, such that X 2 U implies
that a bifurcation of the trivial equilibrium point occurs in X at  = 0
that produces \instant chaos". This means that if U is a neighborhood of the
origin in R4, then there is an  > 0 such that X has a chaotic hyperbolic
invariant set contained in U for either 0 <  ? 0 <  or 0 < 0 ?  < .
Theorem 3 Let G  O(4) be the group generated by p =  and s = 2.
Assume that there are parameter values l = 1, a = ?1 and b; c; d; e for which
Equations 1 have a period doubling cascade of bifurcations. Then there is
an open region V in the space of two parameter families of vector elds on
R4 equivariant with respect to G with the following properties. If X 2 V
and U is a neighborhood of the origin, then there is a positive measure set of
parameters for which X has a chaotic, nonhyperbolic attractor in U .
These theorems are straightforward corollaries of the analysis of the Sil-
nikov bifurcation [11], the theory of period doubling bifurcations for di eo-
morphisms of the plane [12], and the analysis of the Henon map and its gen-
eralizations [2, 9]. To apply these theories, one need note only that rescaled
versions of a generic bifurcation problem with symmetry grouppG converge
to Equations 1 when the phase space variables are scaled by l and time
is rescaled by l. Hyperbolic invariant sets persist under perturbation [10].
Therefore, for parameter values with l = 1 for which the basic equations
have a hyperbolic invariant set, perturbations of the system will also have
a hyperbolic invariant set. We conclude that a generic bifurcation problem
with values of the coecients yielding a hyperbolic invariant set for the ba-
sic system will also have small hyperbolic invariant sets for small values of
j ? 0 j. Since the presence of a Silnikov bifurcation implies the existence
of hyperbolic invariant sets in a ow, Theorem 2 is proved. The proof of
Theorem 3 is similar. In place of the stability results for hyperbolic invariant
sets, we use the results of [2, 9] concerning the existence of positive measure
sets of parameter values yielding chaotic attractors in one parameter fami-
lies of transformations passing through a homoclinic tangency of a periodic
orbit. These results are applied here to two parameter families in which one
parameter creates the bifurcation of the trivial equilibrium and the second
parameter varies coecients of the basic system to create a transversal pas-
sage through a homoclinic tangency close to the limit of a period doubling
cascade.
12
7 Acknowledgements
All of the numerical computations and gures necessary for this work were
generated with the aid of the software package dstool [1]. We would like to
thank Mike Field for suggesting that vector elds with the symmetry studied
here might have interesting dynamical properties. Phil Holmes suggested the
name \instant chaos". This research was partially supported by the National
Science Foundation.

References
[1] A. Back, J. Guckenheimer, M. Myers, F. Wicklin and P. Worfolk. dstool:
Computer assisted exploration of dynamical systems, Notices of the Am.
Math. Soc., 39(4), 303-309.
[2] M. Benedicks and L. Carleson, The dynamics of the Henon map, Annals
of Mathematics, 133 (1991), 73-169.
[3] P. Collet and J.P. Eckmann. Iterated Maps on the Interval as Dynamical
Systems, Birkhauser, 1980.
[4] J.P. Eckmann. Roads to turbulence in dissipative dynamical systems,
Rev. Mod. Phys., 53 (1981), 643-654.
[5] M. Field. Equivariant bifurcation theory and symmetry breaking, J. Dy-
namics and Di . Eqns., 1(4) (1989), 369-421.
[6] M. Field and J. Swift. Stationary bifurcation to limit cycles and hetero-
clinic cycles, Nonlinearity, 4 (1991), 1001-1043.
[7] M. Golubitsky, I. Stewart, and D. Shae er. Singularities and Groups in
Bifurcation Theory, Vol.II., Springer-Verlag, 1988.
[8] J. Guckenheimer and P. Holmes. Nonlinear Oscillations, Dynamical Sys-
tems, and Bifurcations of Vector Fields, Springer-Verlag, 1983.
[9] L. Mora and M. Viana, The abundance of strange attractors, preprint,
IMPA.

13
[10] S. Smale. Di erentiable dynamical systems, Bull. Am. Math. Soc., 73
(1967), 747-817.
[11] C. Tresser. About some theorems by L.P. Silnikov, Ann. Inst. Henri
Poincare, 40(4) (1984), 441-461.
[12] E.B. Vul, G. Sinai and K.M. Khanin. Feigenbaum universality and the
thermodynamic formalism, Russian Math. Surveys, 39(3) (1984), 1-40.

14
[G] (0;0;0;0)

  QQQ
0 QQ
[(Z ) ] x1; ; ;
3
2 ( 0 0 0) [Z ] x1; ;x1;
2
2 ( 0 0)

?? @@ ?? @@
? @ ~ ? @
[(Z )0] x1;x2; ;
2
2 ( 0 0) [Z ] x ; ;x ;
2 ( 1 0 3 0) [Z ] x1;x2;x1;x2
2 ( )

a aa ! !!
a aaa !
!!!
[1] x1;x2;x3;x4
( )

h 0i h i
(Z ) = [f1; sp ; sp? sp; sp? sp? sg] Z = [f1; s; p? sp; sp? spg]
3
2
2 1 1 1 2
2
1 1

h 0i h i
(Z ) = [f1; sp? sp? sg] Z~ = [f1; sp? spg] [Z ] = [f1; sg]
2
2
1 1
2
1
2

Figure 1: Isotropy lattice and representative xed-point subspaces, where


.
p =  , s =  are generators of G  G
2

15
isotropy group representative equilibria n
[G] x =x =x =x =0
1 2 3 4 1
[(Z )0]
3
2
2
1 x = ?a ; x = x = x = 0
2 3 4 8
[Z ]
2
2 2
1 x =x = ?
2
3 a c; x = x = 0

2 + 2 4 8
[(Z )0  Z ] ?
b = d = a b d bd ; x = x = 0
2 2
2
2 2
x1 x2
( + )+ 3 4 16
[Z~ ]
2 | 2 0
[Z ] x21
= b?c d?e = b?c d e = b?cx4d?e = b?c d
x22 x23 ? 16
2 b?c+d+e + + + + ( ?e2
+ )(4a+b+c+d)
[1] ? 32

Figure 2: Zeros of isotropy conjugacy classes.

16
Captions for Figures 3,4,5:

Figure 3: Cascade of period doubling bifurcations.


Figure 4: Two possibilities for a typical orbit in a neighborhood of p.
Figure 5: One branch of the unstable manifold of q for two nearby values
of e.

17

View publication stats

Vous aimerez peut-être aussi