Namespaces
Variants
Actions

Difference between revisions of "Routes to chaos"

From Encyclopedia of Mathematics
Jump to: navigation, search
(Importing text file)
 
m (tex encoded by computer)
 
Line 1: Line 1:
 +
<!--
 +
r0827501.png
 +
$#A+1 = 138 n = 4
 +
$#C+1 = 138 : ~/encyclopedia/old_files/data/R082/R.0802750 Routes to chaos
 +
Automatically converted into TeX, above some diagnostics.
 +
Please remove this comment and the {{TEX|auto}} line below,
 +
if TeX found to be correct.
 +
-->
 +
 +
{{TEX|auto}}
 +
{{TEX|done}}
 +
 
This phrase refers to the process by which a simple attracting set for a [[Dynamical system|dynamical system]] (like a fixed point or periodic orbit) becomes chaotic as an external parameter is varied.
 
This phrase refers to the process by which a simple attracting set for a [[Dynamical system|dynamical system]] (like a fixed point or periodic orbit) becomes chaotic as an external parameter is varied.
  
 
One considers a one-parameter family of differential equations
 
One considers a one-parameter family of differential equations
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827501.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a1)</td></tr></table>
+
$$ \tag{a1 }
 +
 
 +
\frac{d }{dt }
 +
x( t)  = F _  \lambda  ( x( t)),\  x( 0)  = x _ {0} ,
 +
$$
  
 
or difference equations (mappings)
 
or difference equations (mappings)
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827502.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a2)</td></tr></table>
+
$$ \tag{a2 }
 +
x _ {n+ 1 }  = F _  \lambda  ( x _ {n} ) ,
 +
$$
  
where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827503.png" /> is a smooth function of the real parameter <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827504.png" /> and the point <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827505.png" />, which belongs to some finite-dimensional phase space (like <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827506.png" />); <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827507.png" /> is an initial condition. (Other domains are sometimes of interest, for example, mappings of the circle <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827508.png" /> or the torus <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r0827509.png" />.) Suppose that for a fixed value of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275010.png" />, the initial conditions in some open set in the phase space approach a compact set <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275011.png" /> which exhibits sensitive dependence on initial conditions. (Such a set <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275012.png" /> is also called an attractor. However, there is no universally satisfactory definition of attractor (see e.g. [[#References|[a1]]]). In this article it means an [[Invariant set|invariant set]] for the mapping or flow that attracts almost-all initial conditions in some open neighbourhood, that cannot be split into disjoint non-trivial closed invariant subsets, and on which the motion is recurrent. See [[#References|[a2]]] for a discussion of sensitive dependence in one-dimensional mappings; cf. also [[Repelling set|Repelling set]]; [[Strange attractor|Strange attractor]].) Roughly speaking, if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275013.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275014.png" /> are two nearby initial conditions on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275015.png" />, their trajectories remain close only for a short time. Then the trajectories separate at an exponential rate until their subsequent motion appears uncorrelated. (The rate of separation eventually saturates since <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275016.png" /> is bounded.) The attractor <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275017.png" /> is chaotic whenever <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275018.png" /> exhibits sensitive dependence on initial conditions (cf. also [[Chaos|Chaos]]). This article discusses four ways in which chaotic attractors for (a1) and (a2) can arise as the parameter <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275019.png" /> is varied. These are the most common known routes to chaos, and there is ample confirmation of their existence in physical experiments. Other routes to chaos have been found (see e.g. [[#References|[a3]]]), and more undoubtedly remain undiscovered, particularly in high-dimensional systems. See [[#References|[a4]]] for a review article and bibliography.
+
where $  F $
 +
is a smooth function of the real parameter $  \lambda $
 +
and the point $  x $,  
 +
which belongs to some finite-dimensional phase space (like $  \mathbf R  ^ {n} $);  
 +
$  x _ {0} $
 +
is an initial condition. (Other domains are sometimes of interest, for example, mappings of the circle $  S  ^ {1} $
 +
or the torus $  T  ^ {2} $.)  
 +
Suppose that for a fixed value of $  \lambda $,  
 +
the initial conditions in some open set in the phase space approach a compact set $  A $
 +
which exhibits sensitive dependence on initial conditions. (Such a set $  A $
 +
is also called an attractor. However, there is no universally satisfactory definition of attractor (see e.g. [[#References|[a1]]]). In this article it means an [[Invariant set|invariant set]] for the mapping or flow that attracts almost-all initial conditions in some open neighbourhood, that cannot be split into disjoint non-trivial closed invariant subsets, and on which the motion is recurrent. See [[#References|[a2]]] for a discussion of sensitive dependence in one-dimensional mappings; cf. also [[Repelling set|Repelling set]]; [[Strange attractor|Strange attractor]].) Roughly speaking, if $  x _ {0} $
 +
and $  y _ {0} $
 +
are two nearby initial conditions on $  A $,  
 +
their trajectories remain close only for a short time. Then the trajectories separate at an exponential rate until their subsequent motion appears uncorrelated. (The rate of separation eventually saturates since $  A $
 +
is bounded.) The attractor $  A $
 +
is chaotic whenever $  A $
 +
exhibits sensitive dependence on initial conditions (cf. also [[Chaos|Chaos]]). This article discusses four ways in which chaotic attractors for (a1) and (a2) can arise as the parameter $  \lambda $
 +
is varied. These are the most common known routes to chaos, and there is ample confirmation of their existence in physical experiments. Other routes to chaos have been found (see e.g. [[#References|[a3]]]), and more undoubtedly remain undiscovered, particularly in high-dimensional systems. See [[#References|[a4]]] for a review article and bibliography.
  
 
==Period doubling route to chaos.==
 
==Period doubling route to chaos.==
In this route to chaos, a fixed point loses stability (in a pitchfork bifurcation, cf. [[Bifurcation|Bifurcation]]) to an attracting period-2 orbit as the parameter passes some critical value. At some later parameter value, the period-2 orbit loses stability (in a pitchfork bifurcation) to a period-4 orbit, etc. The parameter values at which these period doublings occur form an increasing sequence converging to some finite value <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275020.png" />, at which the original fixed point is replaced by an aperiodic attractor (which may be chaotic for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275021.png" />).
+
In this route to chaos, a fixed point loses stability (in a pitchfork bifurcation, cf. [[Bifurcation|Bifurcation]]) to an attracting period-2 orbit as the parameter passes some critical value. At some later parameter value, the period-2 orbit loses stability (in a pitchfork bifurcation) to a period-4 orbit, etc. The parameter values at which these period doublings occur form an increasing sequence converging to some finite value $  \lambda _  \infty  $,  
 +
at which the original fixed point is replaced by an aperiodic attractor (which may be chaotic for $  \lambda > \lambda _  \infty  $).
  
 
M.J. Feigenbaum
 
M.J. Feigenbaum
Line 18: Line 54:
 
originally studied period doubling in the difference equation
 
originally studied period doubling in the difference equation
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275022.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a3)</td></tr></table>
+
$$ \tag{a3 }
 +
x _ {n+ 1 }  = F _  \lambda  ( x _ {n} )  = \lambda x _ {n} ( 1 - x _ {n} ),
 +
$$
  
also called the quadratic mapping. When <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275023.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275024.png" /> has a non-zero fixed point at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275025.png" /> which is stable for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275026.png" />, since
+
also called the quadratic mapping. When $  \lambda > 1 $,  
 +
$  F _  \lambda  $
 +
has a non-zero fixed point at $  x _ {f} = 1 - 1/ \lambda $
 +
which is stable for $  1 < \lambda < 3 $,  
 +
since
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275027.png" /></td> </tr></table>
+
$$
 +
\left |
 +
\frac{dF _  \lambda  }{dx }
 +
( x _ {f} ) \right |  = | 2- \lambda |  < 1.
 +
$$
  
When <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275028.png" />, the derivative at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275029.png" /> is <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275030.png" />. For slightly larger values of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275031.png" />, the derivative is larger than <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275032.png" /> in absolute value, and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275033.png" /> is unstable: almost-all initial conditions in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275034.png" /> are attracted to a period-2 orbit, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275035.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275036.png" />. A similar derivative evaluation for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275037.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275038.png" />, shows that each <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275039.png" /> is stable for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275040.png" />. At <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275041.png" /> each <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275042.png" /> loses its stability because the derivative <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275043.png" />. As before, each <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275044.png" /> is replaced by a pair of attracting points <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275045.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275046.png" /> such that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275047.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275048.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275049.png" />. The points <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275050.png" /> correspond to an attracting period-4 orbit for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275051.png" /> when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275052.png" /> is slightly larger than <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275053.png" />. This process continues indefinitely; at each bifurcation, the periodic orbit is replaced by a new attracting periodic orbit of twice the period. The parameter values at which the bifurcations occur form an increasing sequence <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275054.png" /> that is bounded above by a number <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275055.png" /> with the property that
+
When $  \lambda = \lambda _ {1} = 3 $,  
 +
the derivative at $  x _ {f} $
 +
is $  - 1 $.  
 +
For slightly larger values of $  \lambda $,  
 +
the derivative is larger than $  1 $
 +
in absolute value, and $  x _ {f} $
 +
is unstable: almost-all initial conditions in $  [ 0, 1] $
 +
are attracted to a period-2 orbit, $  x _ {1} = F _  \lambda  ( x _ {2} ) $,  
 +
$  x _ {2} = F _  \lambda  ( x _ {1} ) $.  
 +
A similar derivative evaluation for $  F _  \lambda  ^ { 2 } ( x _ {i} ) = F _  \lambda  ( F _  \lambda  ( x _ {i} )) = x _ {i} $,
 +
$  i= 1,2 $,  
 +
shows that each $  x _ {i} $
 +
is stable for $  \lambda _ {1} < \lambda < \lambda _ {2} \approx 3.449 $.  
 +
At $  \lambda _ {2} $
 +
each $  x _ {i} $
 +
loses its stability because the derivative $  dF _ {\lambda _ {2}  } ^ { 2 } ( x _ {i} )/dx = - 1 $.  
 +
As before, each $  x _ {i} $
 +
is replaced by a pair of attracting points $  x _ {i1 }  $,  
 +
$  x _ {i2 }  $
 +
such that $  x _ {i1 }  = F _  \lambda  ^ { 2 } ( x _ {i2 }  ) $,
 +
$  x _ {i2 }  = F _  \lambda  ^ { 2 } ( x _ {i1 }  ) $,  
 +
$  i= 1, 2 $.  
 +
The points $  x _ {ij }  $
 +
correspond to an attracting period-4 orbit for $  F _  \lambda  $
 +
when $  \lambda $
 +
is slightly larger than $  \lambda _ {2} $.  
 +
This process continues indefinitely; at each bifurcation, the periodic orbit is replaced by a new attracting periodic orbit of twice the period. The parameter values at which the bifurcations occur form an increasing sequence $  \{ \lambda _ {k} \} $
 +
that is bounded above by a number $  \lambda _  \infty  \approx 3.5699 $
 +
with the property that
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275056.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a4)</td></tr></table>
+
$$ \tag{a4 }
 +
\lim\limits _ {k \rightarrow \infty } 
 +
\frac{\lambda _ {k} - \lambda _ {k- 1 }  }{\lambda _ {k+ 1 }  - \lambda _ {k} }
 +
  = \delta  \approx  4.6692.
 +
$$
  
The number <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275057.png" /> is the Feigenbaum constant. It is a remarkable fact that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275058.png" /> is independent of the details of the mapping <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275059.png" /> as long as <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275060.png" /> satisfies certain general hypotheses; see , [[#References|[a6]]] for details. Chaos occurs in the quadratic mapping for many values of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275061.png" />. In fact, M. Jakobson [[#References|[a7]]] proved that the set of such parameter values has positive measure.
+
The number $  \delta $
 +
is the Feigenbaum constant. It is a remarkable fact that $  \delta $
 +
is independent of the details of the mapping $  F _  \lambda  $
 +
as long as $  F _  \lambda  $
 +
satisfies certain general hypotheses; see , [[#References|[a6]]] for details. Chaos occurs in the quadratic mapping for many values of $  \lambda < 4 $.  
 +
In fact, M. Jakobson [[#References|[a7]]] proved that the set of such parameter values has positive measure.
  
Similar results hold in higher dimensions, i.e., for mappings <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275062.png" /> where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275063.png" /> [[#References|[a8]]]. In [[#References|[a9]]] it is shown how period-doubling cascades arise in the formation of  "horseshoes"  as the parameter varies. [[#References|[a10]]] contains a collection of papers describing the existence of period doubling in a variety of physical situations.
+
Similar results hold in higher dimensions, i.e., for mappings $  \mathbf R  ^ {n} \rightarrow \mathbf R  ^ {n} $
 +
where $  n > 1 $[[#References|[a8]]]. In [[#References|[a9]]] it is shown how period-doubling cascades arise in the formation of  "horseshoes"  as the parameter varies. [[#References|[a10]]] contains a collection of papers describing the existence of period doubling in a variety of physical situations.
  
 
==Intermittency route to chaos.==
 
==Intermittency route to chaos.==
Y. Pomeau and P. Manneville [[#References|[a11]]] describe how an attracting periodic orbit (like a fixed point) for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275064.png" /> can disappear and be replaced by a chaotic attractor for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275065.png" />. For <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275066.png" /> slightly larger than <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275067.png" />, initial conditions near the periodic orbit remain there for a long time. This regular behaviour is interrupted by a  "burst"  in which the trajectory moves away from a neighbourhood of the periodic orbit and exhibits possibly irregular behaviour. Eventually, the trajectory re-enters the region near the periodic orbit and the process is repeated. The dynamical behaviour for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275068.png" /> is characterized by an infinite sequence of intervals of nearly periodic (laminar) motion followed by bursts. The length of the laminar regions scales as <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275069.png" /> for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275070.png" /> near <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275071.png" /> [[#References|[a11]]].
+
Y. Pomeau and P. Manneville [[#References|[a11]]] describe how an attracting periodic orbit (like a fixed point) for $  \lambda < \lambda _ {c} $
 +
can disappear and be replaced by a chaotic attractor for $  \lambda > \lambda _ {c} $.  
 +
For $  \lambda $
 +
slightly larger than $  \lambda _ {c} $,  
 +
initial conditions near the periodic orbit remain there for a long time. This regular behaviour is interrupted by a  "burst"  in which the trajectory moves away from a neighbourhood of the periodic orbit and exhibits possibly irregular behaviour. Eventually, the trajectory re-enters the region near the periodic orbit and the process is repeated. The dynamical behaviour for $  \lambda > \lambda _ {c} $
 +
is characterized by an infinite sequence of intervals of nearly periodic (laminar) motion followed by bursts. The length of the laminar regions scales as $  ( \lambda - \lambda _ {c} ) ^ {- 1/2 } $
 +
for $  \lambda $
 +
near $  \lambda _ {c} $[[#References|[a11]]].
  
Three types of intermittency are distinguished, depending on the eigenvalues of the associated [[Jacobian|Jacobian]] matrix of partial derivatives evaluated at the periodic orbit. Type-<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275073.png" /> intermittency occurs when a stable and an unstable periodic orbit that coexist for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275074.png" /> collide at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275075.png" /> (the Jacobian matrix at the resulting periodic orbit has eigenvalue <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275076.png" />) and disappear for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275077.png" /> (i.e., there is a saddle-node bifurcation at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275078.png" />). An example using the difference equation
+
Three types of intermittency are distinguished, depending on the eigenvalues of the associated [[Jacobian|Jacobian]] matrix of partial derivatives evaluated at the periodic orbit. Type- $  1 $
 +
intermittency occurs when a stable and an unstable periodic orbit that coexist for $  \lambda < \lambda _ {c} $
 +
collide at $  \lambda = \lambda _ {c} $(
 +
the Jacobian matrix at the resulting periodic orbit has eigenvalue $  1 $)  
 +
and disappear for $  \lambda > \lambda _ {c} $(
 +
i.e., there is a saddle-node bifurcation at $  \lambda _ {c} $).  
 +
An example using the difference equation
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275079.png" /></td> <td valign="top" style="width:5%;text-align:right;">(a5)</td></tr></table>
+
$$ \tag{a5 }
 +
x _ {n+ 1 }  = F _  \lambda  ( x _ {n} )  = 1 - \lambda x _ {n}  ^ {2}
 +
$$
  
is given in [[#References|[a4]]]. For <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275080.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275081.png" /> has one stable and one unstable period-3 orbit. They collide at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275082.png" />, and for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275083.png" /> slightly less than <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275084.png" />, the iterates exhibit a nearly period-3 motion interrupted by bursts. Type-<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275085.png" /> intermittency has been found in Poincaré mappings of the Lorenz equations (cf. [[Lorenz attractor|Lorenz attractor]], [[#References|[a11]]]) and in experiments on oscillating chemical reactions [[#References|[a12]]].
+
is given in [[#References|[a4]]]. For $  \lambda > 1.75 $,  
 +
$  F _  \lambda  ^ { 3 } $
 +
has one stable and one unstable period-3 orbit. They collide at $  \lambda = 1.75 $,  
 +
and for $  \lambda $
 +
slightly less than $  1.75 $,  
 +
the iterates exhibit a nearly period-3 motion interrupted by bursts. Type- $  1 $
 +
intermittency has been found in Poincaré mappings of the Lorenz equations (cf. [[Lorenz attractor|Lorenz attractor]], [[#References|[a11]]]) and in experiments on oscillating chemical reactions [[#References|[a12]]].
  
Type-<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275087.png" /> intermittency occurs when the eigenvalues of the associated Jacobian matrix are complex and cross the unit circle away from the real line. In type-<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275089.png" /> intermittency, the eigenvalues pass through <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275090.png" />. Heuristic arguments and numerical evidence suggest that the [[Lyapunov characteristic exponent|Lyapunov characteristic exponent]] of the chaotic attractor created when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275091.png" /> passes <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275092.png" /> scales as <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275093.png" />, at least in the case of type-<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275094.png" /> and type-<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275095.png" /> intermittency [[#References|[a11]]].
+
Type- $  2 $
 +
intermittency occurs when the eigenvalues of the associated Jacobian matrix are complex and cross the unit circle away from the real line. In type- $  3 $
 +
intermittency, the eigenvalues pass through $  - 1 $.  
 +
Heuristic arguments and numerical evidence suggest that the [[Lyapunov characteristic exponent|Lyapunov characteristic exponent]] of the chaotic attractor created when $  \lambda $
 +
passes $  \lambda _ {c} $
 +
scales as $  ( \lambda - \lambda _ {c} ) ^ {- 1/2 } $,  
 +
at least in the case of type- $  1 $
 +
and type- $  3 $
 +
intermittency [[#References|[a11]]].
  
 
==Ruelle–Takens–Newhouse route to chaos.==
 
==Ruelle–Takens–Newhouse route to chaos.==
Suppose that for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275096.png" /> there is an attracting fixed point for (a1) that loses its stability in a Hopf bifurcation at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275097.png" /> (i.e., the fixed point is replaced by an attracting periodic orbit). In addition, suppose that at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275098.png" /> there is another Hopf bifurcation to a quasi-periodic orbit (i.e., the attractor is now a torus <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r08275099.png" />). A subsequent Hopf bifurcation at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750100.png" /> creates a quasi-periodic <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750101.png" />-torus. However, S.E. Newhouse, D. Ruelle and F. Takens [[#References|[a13]]] showed that for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750102.png" /> every constant vector field on the torus <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750103.png" /> can be perturbed by an arbitrarily small amount to a new vector field with a chaotic attractor. Thus, in an experiment, one may see a bifurcation from a <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750104.png" />-frequency quasi-periodic flow to a chaotic attractor; see, for example, [[#References|[a14]]], . This route to turbulence is in contrast to the classical theory of L. Landau and E. Lifshitz [[#References|[a16]]], which says that turbulence arises from the successive addition of incommensurate frequences as a parameter is increased.
+
Suppose that for $  \lambda < \lambda _ {1} $
 +
there is an attracting fixed point for (a1) that loses its stability in a Hopf bifurcation at $  \lambda = \lambda _ {1} $(
 +
i.e., the fixed point is replaced by an attracting periodic orbit). In addition, suppose that at $  \lambda = \lambda _ {2} > \lambda _ {1} $
 +
there is another Hopf bifurcation to a quasi-periodic orbit (i.e., the attractor is now a torus $  T  ^ {2} $).  
 +
A subsequent Hopf bifurcation at $  \lambda = \lambda _ {3} > \lambda _ {2} $
 +
creates a quasi-periodic $  3 $-
 +
torus. However, S.E. Newhouse, D. Ruelle and F. Takens [[#References|[a13]]] showed that for $  n \geq  3 $
 +
every constant vector field on the torus $  T  ^ {n} $
 +
can be perturbed by an arbitrarily small amount to a new vector field with a chaotic attractor. Thus, in an experiment, one may see a bifurcation from a $  2 $-
 +
frequency quasi-periodic flow to a chaotic attractor; see, for example, [[#References|[a14]]], . This route to turbulence is in contrast to the classical theory of L. Landau and E. Lifshitz [[#References|[a16]]], which says that turbulence arises from the successive addition of incommensurate frequences as a parameter is increased.
  
 
==Crisis route to chaos.==
 
==Crisis route to chaos.==
The previous examples showed how an existing attractor can change its character as a parameter is varied. In a crisis, there is initially only a chaotic transient for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750105.png" />, i.e., initial conditions in some region of the phase space approach what appears to be a chaotic attractor and may spend a long time near it. Eventually, however, the orbit leaves for another part of the phase space, never to return. The time spent near the chaotic transient becomes longer as <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750106.png" />. When <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750107.png" />, the chaotic transient becomes a chaotic attractor: initial conditions that once escaped now remain forever.
+
The previous examples showed how an existing attractor can change its character as a parameter is varied. In a crisis, there is initially only a chaotic transient for $  \lambda > \lambda _ {c} $,  
 +
i.e., initial conditions in some region of the phase space approach what appears to be a chaotic attractor and may spend a long time near it. Eventually, however, the orbit leaves for another part of the phase space, never to return. The time spent near the chaotic transient becomes longer as $  \lambda \rightarrow \lambda _ {c} $.  
 +
When $  \lambda \leq  \lambda _ {c} $,  
 +
the chaotic transient becomes a chaotic attractor: initial conditions that once escaped now remain forever.
  
A simple example of a crisis is given by the quadratic mapping (a3). For <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750108.png" />, almost every initial condition in the interval <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750109.png" /> generates a trajectory that bounces chaotically in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750110.png" /> for a time. Eventually, some iterate falls to the left of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750111.png" />, and the orbit tends to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750112.png" />. At <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750113.png" />, the transient is converted to an attractor: almost every initial condition in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750114.png" /> approaches a chaotic attractor.
+
A simple example of a crisis is given by the quadratic mapping (a3). For $  \lambda > 4 $,  
 +
almost every initial condition in the interval $  I=[ 0, 1] $
 +
generates a trajectory that bounces chaotically in $  I $
 +
for a time. Eventually, some iterate falls to the left of 0 $,  
 +
and the orbit tends to $  - \infty $.  
 +
At $  \lambda = 4 $,  
 +
the transient is converted to an attractor: almost every initial condition in $  I $
 +
approaches a chaotic attractor.
  
In this example, the chaotic attractor is contained in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750115.png" /> when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750116.png" />. The crisis occurs at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750117.png" /> when an interior point of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750118.png" /> is mapped to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750119.png" />, which is part of the stable manifold of the unstable fixed point at <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750120.png" />. For <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750121.png" />, a portion of the interval <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750122.png" /> is mapped into the basin of attraction (cf. [[Chaos|Chaos]]) for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750123.png" />, so the chaotic attractor is replaced by a chaotic transient. The mean time that typical initial conditions spend in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750124.png" /> before escaping to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750125.png" /> scales as <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750126.png" /> with the parameter [[#References|[a17]]].
+
In this example, the chaotic attractor is contained in $  ( 0, 1) $
 +
when $  \lambda < 4 $.  
 +
The crisis occurs at $  \lambda = 4 $
 +
when an interior point of $  ( 0, 1) $
 +
is mapped to $  1 $,  
 +
which is part of the stable manifold of the unstable fixed point at 0 $.  
 +
For $  \lambda > 4 $,  
 +
a portion of the interval $  I $
 +
is mapped into the basin of attraction (cf. [[Chaos|Chaos]]) for $  - \infty $,  
 +
so the chaotic attractor is replaced by a chaotic transient. The mean time that typical initial conditions spend in $  I $
 +
before escaping to $  - \infty $
 +
scales as $  ( \lambda - 4) ^ {- 1/2 } $
 +
with the parameter [[#References|[a17]]].
  
 
Similar results hold in higher dimensions. [[#References|[a17]]] discusses the Hénon mapping
 
Similar results hold in higher dimensions. [[#References|[a17]]] discusses the Hénon mapping
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750127.png" /></td> </tr></table>
+
$$
 +
x _ {n+ 1 }  = 1 - \lambda x _ {n}  ^ {2} + y _ {n} ,
 +
$$
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750128.png" /></td> </tr></table>
+
$$
 +
y _ {n+ 1 }  = 0.3  x _ {n} ,
 +
$$
  
where a <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750129.png" />-piece and a <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750130.png" />-piece chaotic attractor coexist for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750131.png" />. As <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750132.png" />, the <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750133.png" />-piece attractor approaches the stable manifold of a period-6 saddle orbit that forms part of the boundary of the basin of attraction for the <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750134.png" />-piece attractor. When <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750135.png" />, part of the <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750136.png" />-piece attractor crosses this stable manifold. Thus, for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750137.png" /> the <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750138.png" />-piece attractor becomes a transient — eventually some iterate maps into the basin of the <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750139.png" />-piece attractor. Numerical evidence suggests that the mean time spent near the chaotic transient is proportional to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750140.png" />, where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750141.png" /> is a  "critical exponent"  that can be expressed in terms of the eigenvalues associated with the periodic orbit involved in the crisis [[#References|[a18]]]. The existence of crises has also been demonstrated for the Lorenz equations and mappings of the torus [[#References|[a17]]].
+
where a $  6 $-
 +
piece and a $  4 $-
 +
piece chaotic attractor coexist for $  1.0624 < \lambda < \lambda _ {c} \approx 1.08074 $.  
 +
As $  \lambda \rightarrow \lambda _ {c} $,  
 +
the $  6 $-
 +
piece attractor approaches the stable manifold of a period-6 saddle orbit that forms part of the boundary of the basin of attraction for the $  4 $-
 +
piece attractor. When $  \lambda > \lambda _ {c} $,  
 +
part of the $  6 $-
 +
piece attractor crosses this stable manifold. Thus, for $  \lambda > \lambda _ {c} $
 +
the $  6 $-
 +
piece attractor becomes a transient — eventually some iterate maps into the basin of the $  4 $-
 +
piece attractor. Numerical evidence suggests that the mean time spent near the chaotic transient is proportional to $  ( \lambda - \lambda _ {c} ) ^ {- \gamma } $,  
 +
where $  \gamma $
 +
is a  "critical exponent"  that can be expressed in terms of the eigenvalues associated with the periodic orbit involved in the crisis [[#References|[a18]]]. The existence of crises has also been demonstrated for the Lorenz equations and mappings of the torus [[#References|[a17]]].
  
 
====References====
 
====References====
 
<table><TR><TD valign="top">[a1]</TD> <TD valign="top">  J. Milnor,  "On the concept of attractor"  ''Commun. Math. Phys.'' , '''99'''  (1985)  pp. 177–195</TD></TR><TR><TD valign="top">[a2]</TD> <TD valign="top">  J. Guckenheimer,  "Sensitive dependence to initial conditions for one-dimensional maps"  ''Commun. Math. Phys.'' , '''70'''  (1979)  pp. 133–160</TD></TR><TR><TD valign="top">[a3]</TD> <TD valign="top">  E.N. Lorenz,  ''Physica D'' , '''35'''  (1989)  pp. 299–317</TD></TR><TR><TD valign="top">[a4]</TD> <TD valign="top">  J.-P. Eckmann,  "Roads to turbulence in dissipative dynamical systems"  ''Rev. Mod. Phys.'' , '''53'''  (1981)  pp. 643–654</TD></TR><TR><TD valign="top">[a5a]</TD> <TD valign="top">  M.J. Feigenbaum,  "Qualitative universality for a class of nonlinear transformations"  ''J. Stat. Phys.'' , '''19'''  (1978)  pp. 25–52</TD></TR><TR><TD valign="top">[a5b]</TD> <TD valign="top">  M.J. Feigenbaum,  "The universal metric properties of nonlinear transformations"  ''J. Stat. Phys.'' , '''21'''  (1979)  pp. 669–706</TD></TR><TR><TD valign="top">[a6]</TD> <TD valign="top">  P. Collet,  J.-P. Eckmann,  O.E. Lanford,  "Universal properties of maps on an interval"  ''Commun. Math. Phys.'' , '''76'''  (1980)  pp. 211–254</TD></TR><TR><TD valign="top">[a7]</TD> <TD valign="top">  M. Jakobson,  "Absolutely continuous invariant measures for one-parameter families of one-dimensional maps"  ''Commun. Math. Phys.'' , '''81'''  (1981)  pp. 39–88</TD></TR><TR><TD valign="top">[a8]</TD> <TD valign="top">  P. Collet,  J.-P. Eckmann,  H. Koch,  "Period doubling bifurcations for families of maps on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750142.png" />"  ''J. Stat. Phys.'' , '''25'''  (1981)  pp. 1–14</TD></TR><TR><TD valign="top">[a9]</TD> <TD valign="top">  J.A. Yorke,  K.T. Alligood,  "Period doubling cascades of attractors: a prerequisite for horseshoes"  ''Commun. Math. Phys.'' , '''101'''  (1985)  pp. 305–321</TD></TR><TR><TD valign="top">[a10]</TD> <TD valign="top">  P. Cvitanović (ed.) , ''Universality in chaos'' , A. Hilger  (1989)</TD></TR><TR><TD valign="top">[a11]</TD> <TD valign="top">  Y. Pomeau,  P. Manneville,  "Intermittent transition to turbulence in dissipative dynamical systems"  ''Commun. Math. Phys.'' , '''74'''  (1980)  pp. 189–197</TD></TR><TR><TD valign="top">[a12]</TD> <TD valign="top">  F. Argoul,  A. Arneodo,  P. Richetti,  J.C. Roux,  H.L. Swinney,  ''Acct. Chem. Res.'' , '''20'''  (1987)  pp. 436–442</TD></TR><TR><TD valign="top">[a13]</TD> <TD valign="top">  S.E. Newhouse,  D. Ruelle,  F. Takens,  "Occurrence of strange axiom <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750143.png" /> attractors near quasiperiodic flow on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750144.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750145.png" />"  ''Commun. Math. Phys.'' , '''64'''  (1978)  pp. 35–40</TD></TR><TR><TD valign="top">[a14]</TD> <TD valign="top">  M. Giglio,  S. Musazzi,  U. Perini,  "Transition to chaotic behaviour via a reproducible sequence of period-doubling bifurcations"  ''Phys. Rev. Lett.'' , '''47'''  (1981)  pp. 243–246</TD></TR><TR><TD valign="top">[a15a]</TD> <TD valign="top">  P.R. Fenstermacher,  H.L. Swinney,  J.P. Gollub,  "Dynamical instabilities and transition to chaotic Taylor vortex flow"  ''J. Fluid Mech.'' , '''94'''  (1979)  pp. 103–128</TD></TR><TR><TD valign="top">[a15b]</TD> <TD valign="top">  A. Brandstäter,  H.L. Swinney,  ''Phys. Rev. A'' , '''35'''  (1987)  pp. 2207–2220</TD></TR><TR><TD valign="top">[a16]</TD> <TD valign="top">  L.D. Landau,  E.M. Lifshitz,  "Fluid mechanics" , Pergamon  (1959)  (Translated from Russian)</TD></TR><TR><TD valign="top">[a17]</TD> <TD valign="top">  C. Grebogi,  E. Ott,  J.A. Yorke,  "Crises, sudden changes in chaotic attractors, and transient chaos"  ''Physica D'' , '''7'''  (1983)  pp. 181–200</TD></TR><TR><TD valign="top">[a18]</TD> <TD valign="top">  C. Grebogi,  E. Ott,  J.A. Yorke,  ''Phys. Rev. Lett.'' , '''57'''  (1986)  pp. 1284–1287</TD></TR></table>
 
<table><TR><TD valign="top">[a1]</TD> <TD valign="top">  J. Milnor,  "On the concept of attractor"  ''Commun. Math. Phys.'' , '''99'''  (1985)  pp. 177–195</TD></TR><TR><TD valign="top">[a2]</TD> <TD valign="top">  J. Guckenheimer,  "Sensitive dependence to initial conditions for one-dimensional maps"  ''Commun. Math. Phys.'' , '''70'''  (1979)  pp. 133–160</TD></TR><TR><TD valign="top">[a3]</TD> <TD valign="top">  E.N. Lorenz,  ''Physica D'' , '''35'''  (1989)  pp. 299–317</TD></TR><TR><TD valign="top">[a4]</TD> <TD valign="top">  J.-P. Eckmann,  "Roads to turbulence in dissipative dynamical systems"  ''Rev. Mod. Phys.'' , '''53'''  (1981)  pp. 643–654</TD></TR><TR><TD valign="top">[a5a]</TD> <TD valign="top">  M.J. Feigenbaum,  "Qualitative universality for a class of nonlinear transformations"  ''J. Stat. Phys.'' , '''19'''  (1978)  pp. 25–52</TD></TR><TR><TD valign="top">[a5b]</TD> <TD valign="top">  M.J. Feigenbaum,  "The universal metric properties of nonlinear transformations"  ''J. Stat. Phys.'' , '''21'''  (1979)  pp. 669–706</TD></TR><TR><TD valign="top">[a6]</TD> <TD valign="top">  P. Collet,  J.-P. Eckmann,  O.E. Lanford,  "Universal properties of maps on an interval"  ''Commun. Math. Phys.'' , '''76'''  (1980)  pp. 211–254</TD></TR><TR><TD valign="top">[a7]</TD> <TD valign="top">  M. Jakobson,  "Absolutely continuous invariant measures for one-parameter families of one-dimensional maps"  ''Commun. Math. Phys.'' , '''81'''  (1981)  pp. 39–88</TD></TR><TR><TD valign="top">[a8]</TD> <TD valign="top">  P. Collet,  J.-P. Eckmann,  H. Koch,  "Period doubling bifurcations for families of maps on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750142.png" />"  ''J. Stat. Phys.'' , '''25'''  (1981)  pp. 1–14</TD></TR><TR><TD valign="top">[a9]</TD> <TD valign="top">  J.A. Yorke,  K.T. Alligood,  "Period doubling cascades of attractors: a prerequisite for horseshoes"  ''Commun. Math. Phys.'' , '''101'''  (1985)  pp. 305–321</TD></TR><TR><TD valign="top">[a10]</TD> <TD valign="top">  P. Cvitanović (ed.) , ''Universality in chaos'' , A. Hilger  (1989)</TD></TR><TR><TD valign="top">[a11]</TD> <TD valign="top">  Y. Pomeau,  P. Manneville,  "Intermittent transition to turbulence in dissipative dynamical systems"  ''Commun. Math. Phys.'' , '''74'''  (1980)  pp. 189–197</TD></TR><TR><TD valign="top">[a12]</TD> <TD valign="top">  F. Argoul,  A. Arneodo,  P. Richetti,  J.C. Roux,  H.L. Swinney,  ''Acct. Chem. Res.'' , '''20'''  (1987)  pp. 436–442</TD></TR><TR><TD valign="top">[a13]</TD> <TD valign="top">  S.E. Newhouse,  D. Ruelle,  F. Takens,  "Occurrence of strange axiom <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750143.png" /> attractors near quasiperiodic flow on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750144.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/r/r082/r082750/r082750145.png" />"  ''Commun. Math. Phys.'' , '''64'''  (1978)  pp. 35–40</TD></TR><TR><TD valign="top">[a14]</TD> <TD valign="top">  M. Giglio,  S. Musazzi,  U. Perini,  "Transition to chaotic behaviour via a reproducible sequence of period-doubling bifurcations"  ''Phys. Rev. Lett.'' , '''47'''  (1981)  pp. 243–246</TD></TR><TR><TD valign="top">[a15a]</TD> <TD valign="top">  P.R. Fenstermacher,  H.L. Swinney,  J.P. Gollub,  "Dynamical instabilities and transition to chaotic Taylor vortex flow"  ''J. Fluid Mech.'' , '''94'''  (1979)  pp. 103–128</TD></TR><TR><TD valign="top">[a15b]</TD> <TD valign="top">  A. Brandstäter,  H.L. Swinney,  ''Phys. Rev. A'' , '''35'''  (1987)  pp. 2207–2220</TD></TR><TR><TD valign="top">[a16]</TD> <TD valign="top">  L.D. Landau,  E.M. Lifshitz,  "Fluid mechanics" , Pergamon  (1959)  (Translated from Russian)</TD></TR><TR><TD valign="top">[a17]</TD> <TD valign="top">  C. Grebogi,  E. Ott,  J.A. Yorke,  "Crises, sudden changes in chaotic attractors, and transient chaos"  ''Physica D'' , '''7'''  (1983)  pp. 181–200</TD></TR><TR><TD valign="top">[a18]</TD> <TD valign="top">  C. Grebogi,  E. Ott,  J.A. Yorke,  ''Phys. Rev. Lett.'' , '''57'''  (1986)  pp. 1284–1287</TD></TR></table>

Latest revision as of 08:12, 6 June 2020


This phrase refers to the process by which a simple attracting set for a dynamical system (like a fixed point or periodic orbit) becomes chaotic as an external parameter is varied.

One considers a one-parameter family of differential equations

$$ \tag{a1 } \frac{d }{dt } x( t) = F _ \lambda ( x( t)),\ x( 0) = x _ {0} , $$

or difference equations (mappings)

$$ \tag{a2 } x _ {n+ 1 } = F _ \lambda ( x _ {n} ) , $$

where $ F $ is a smooth function of the real parameter $ \lambda $ and the point $ x $, which belongs to some finite-dimensional phase space (like $ \mathbf R ^ {n} $); $ x _ {0} $ is an initial condition. (Other domains are sometimes of interest, for example, mappings of the circle $ S ^ {1} $ or the torus $ T ^ {2} $.) Suppose that for a fixed value of $ \lambda $, the initial conditions in some open set in the phase space approach a compact set $ A $ which exhibits sensitive dependence on initial conditions. (Such a set $ A $ is also called an attractor. However, there is no universally satisfactory definition of attractor (see e.g. [a1]). In this article it means an invariant set for the mapping or flow that attracts almost-all initial conditions in some open neighbourhood, that cannot be split into disjoint non-trivial closed invariant subsets, and on which the motion is recurrent. See [a2] for a discussion of sensitive dependence in one-dimensional mappings; cf. also Repelling set; Strange attractor.) Roughly speaking, if $ x _ {0} $ and $ y _ {0} $ are two nearby initial conditions on $ A $, their trajectories remain close only for a short time. Then the trajectories separate at an exponential rate until their subsequent motion appears uncorrelated. (The rate of separation eventually saturates since $ A $ is bounded.) The attractor $ A $ is chaotic whenever $ A $ exhibits sensitive dependence on initial conditions (cf. also Chaos). This article discusses four ways in which chaotic attractors for (a1) and (a2) can arise as the parameter $ \lambda $ is varied. These are the most common known routes to chaos, and there is ample confirmation of their existence in physical experiments. Other routes to chaos have been found (see e.g. [a3]), and more undoubtedly remain undiscovered, particularly in high-dimensional systems. See [a4] for a review article and bibliography.

Period doubling route to chaos.

In this route to chaos, a fixed point loses stability (in a pitchfork bifurcation, cf. Bifurcation) to an attracting period-2 orbit as the parameter passes some critical value. At some later parameter value, the period-2 orbit loses stability (in a pitchfork bifurcation) to a period-4 orbit, etc. The parameter values at which these period doublings occur form an increasing sequence converging to some finite value $ \lambda _ \infty $, at which the original fixed point is replaced by an aperiodic attractor (which may be chaotic for $ \lambda > \lambda _ \infty $).

M.J. Feigenbaum

originally studied period doubling in the difference equation

$$ \tag{a3 } x _ {n+ 1 } = F _ \lambda ( x _ {n} ) = \lambda x _ {n} ( 1 - x _ {n} ), $$

also called the quadratic mapping. When $ \lambda > 1 $, $ F _ \lambda $ has a non-zero fixed point at $ x _ {f} = 1 - 1/ \lambda $ which is stable for $ 1 < \lambda < 3 $, since

$$ \left | \frac{dF _ \lambda }{dx } ( x _ {f} ) \right | = | 2- \lambda | < 1. $$

When $ \lambda = \lambda _ {1} = 3 $, the derivative at $ x _ {f} $ is $ - 1 $. For slightly larger values of $ \lambda $, the derivative is larger than $ 1 $ in absolute value, and $ x _ {f} $ is unstable: almost-all initial conditions in $ [ 0, 1] $ are attracted to a period-2 orbit, $ x _ {1} = F _ \lambda ( x _ {2} ) $, $ x _ {2} = F _ \lambda ( x _ {1} ) $. A similar derivative evaluation for $ F _ \lambda ^ { 2 } ( x _ {i} ) = F _ \lambda ( F _ \lambda ( x _ {i} )) = x _ {i} $, $ i= 1,2 $, shows that each $ x _ {i} $ is stable for $ \lambda _ {1} < \lambda < \lambda _ {2} \approx 3.449 $. At $ \lambda _ {2} $ each $ x _ {i} $ loses its stability because the derivative $ dF _ {\lambda _ {2} } ^ { 2 } ( x _ {i} )/dx = - 1 $. As before, each $ x _ {i} $ is replaced by a pair of attracting points $ x _ {i1 } $, $ x _ {i2 } $ such that $ x _ {i1 } = F _ \lambda ^ { 2 } ( x _ {i2 } ) $, $ x _ {i2 } = F _ \lambda ^ { 2 } ( x _ {i1 } ) $, $ i= 1, 2 $. The points $ x _ {ij } $ correspond to an attracting period-4 orbit for $ F _ \lambda $ when $ \lambda $ is slightly larger than $ \lambda _ {2} $. This process continues indefinitely; at each bifurcation, the periodic orbit is replaced by a new attracting periodic orbit of twice the period. The parameter values at which the bifurcations occur form an increasing sequence $ \{ \lambda _ {k} \} $ that is bounded above by a number $ \lambda _ \infty \approx 3.5699 $ with the property that

$$ \tag{a4 } \lim\limits _ {k \rightarrow \infty } \frac{\lambda _ {k} - \lambda _ {k- 1 } }{\lambda _ {k+ 1 } - \lambda _ {k} } = \delta \approx 4.6692. $$

The number $ \delta $ is the Feigenbaum constant. It is a remarkable fact that $ \delta $ is independent of the details of the mapping $ F _ \lambda $ as long as $ F _ \lambda $ satisfies certain general hypotheses; see , [a6] for details. Chaos occurs in the quadratic mapping for many values of $ \lambda < 4 $. In fact, M. Jakobson [a7] proved that the set of such parameter values has positive measure.

Similar results hold in higher dimensions, i.e., for mappings $ \mathbf R ^ {n} \rightarrow \mathbf R ^ {n} $ where $ n > 1 $[a8]. In [a9] it is shown how period-doubling cascades arise in the formation of "horseshoes" as the parameter varies. [a10] contains a collection of papers describing the existence of period doubling in a variety of physical situations.

Intermittency route to chaos.

Y. Pomeau and P. Manneville [a11] describe how an attracting periodic orbit (like a fixed point) for $ \lambda < \lambda _ {c} $ can disappear and be replaced by a chaotic attractor for $ \lambda > \lambda _ {c} $. For $ \lambda $ slightly larger than $ \lambda _ {c} $, initial conditions near the periodic orbit remain there for a long time. This regular behaviour is interrupted by a "burst" in which the trajectory moves away from a neighbourhood of the periodic orbit and exhibits possibly irregular behaviour. Eventually, the trajectory re-enters the region near the periodic orbit and the process is repeated. The dynamical behaviour for $ \lambda > \lambda _ {c} $ is characterized by an infinite sequence of intervals of nearly periodic (laminar) motion followed by bursts. The length of the laminar regions scales as $ ( \lambda - \lambda _ {c} ) ^ {- 1/2 } $ for $ \lambda $ near $ \lambda _ {c} $[a11].

Three types of intermittency are distinguished, depending on the eigenvalues of the associated Jacobian matrix of partial derivatives evaluated at the periodic orbit. Type- $ 1 $ intermittency occurs when a stable and an unstable periodic orbit that coexist for $ \lambda < \lambda _ {c} $ collide at $ \lambda = \lambda _ {c} $( the Jacobian matrix at the resulting periodic orbit has eigenvalue $ 1 $) and disappear for $ \lambda > \lambda _ {c} $( i.e., there is a saddle-node bifurcation at $ \lambda _ {c} $). An example using the difference equation

$$ \tag{a5 } x _ {n+ 1 } = F _ \lambda ( x _ {n} ) = 1 - \lambda x _ {n} ^ {2} $$

is given in [a4]. For $ \lambda > 1.75 $, $ F _ \lambda ^ { 3 } $ has one stable and one unstable period-3 orbit. They collide at $ \lambda = 1.75 $, and for $ \lambda $ slightly less than $ 1.75 $, the iterates exhibit a nearly period-3 motion interrupted by bursts. Type- $ 1 $ intermittency has been found in Poincaré mappings of the Lorenz equations (cf. Lorenz attractor, [a11]) and in experiments on oscillating chemical reactions [a12].

Type- $ 2 $ intermittency occurs when the eigenvalues of the associated Jacobian matrix are complex and cross the unit circle away from the real line. In type- $ 3 $ intermittency, the eigenvalues pass through $ - 1 $. Heuristic arguments and numerical evidence suggest that the Lyapunov characteristic exponent of the chaotic attractor created when $ \lambda $ passes $ \lambda _ {c} $ scales as $ ( \lambda - \lambda _ {c} ) ^ {- 1/2 } $, at least in the case of type- $ 1 $ and type- $ 3 $ intermittency [a11].

Ruelle–Takens–Newhouse route to chaos.

Suppose that for $ \lambda < \lambda _ {1} $ there is an attracting fixed point for (a1) that loses its stability in a Hopf bifurcation at $ \lambda = \lambda _ {1} $( i.e., the fixed point is replaced by an attracting periodic orbit). In addition, suppose that at $ \lambda = \lambda _ {2} > \lambda _ {1} $ there is another Hopf bifurcation to a quasi-periodic orbit (i.e., the attractor is now a torus $ T ^ {2} $). A subsequent Hopf bifurcation at $ \lambda = \lambda _ {3} > \lambda _ {2} $ creates a quasi-periodic $ 3 $- torus. However, S.E. Newhouse, D. Ruelle and F. Takens [a13] showed that for $ n \geq 3 $ every constant vector field on the torus $ T ^ {n} $ can be perturbed by an arbitrarily small amount to a new vector field with a chaotic attractor. Thus, in an experiment, one may see a bifurcation from a $ 2 $- frequency quasi-periodic flow to a chaotic attractor; see, for example, [a14], . This route to turbulence is in contrast to the classical theory of L. Landau and E. Lifshitz [a16], which says that turbulence arises from the successive addition of incommensurate frequences as a parameter is increased.

Crisis route to chaos.

The previous examples showed how an existing attractor can change its character as a parameter is varied. In a crisis, there is initially only a chaotic transient for $ \lambda > \lambda _ {c} $, i.e., initial conditions in some region of the phase space approach what appears to be a chaotic attractor and may spend a long time near it. Eventually, however, the orbit leaves for another part of the phase space, never to return. The time spent near the chaotic transient becomes longer as $ \lambda \rightarrow \lambda _ {c} $. When $ \lambda \leq \lambda _ {c} $, the chaotic transient becomes a chaotic attractor: initial conditions that once escaped now remain forever.

A simple example of a crisis is given by the quadratic mapping (a3). For $ \lambda > 4 $, almost every initial condition in the interval $ I=[ 0, 1] $ generates a trajectory that bounces chaotically in $ I $ for a time. Eventually, some iterate falls to the left of $ 0 $, and the orbit tends to $ - \infty $. At $ \lambda = 4 $, the transient is converted to an attractor: almost every initial condition in $ I $ approaches a chaotic attractor.

In this example, the chaotic attractor is contained in $ ( 0, 1) $ when $ \lambda < 4 $. The crisis occurs at $ \lambda = 4 $ when an interior point of $ ( 0, 1) $ is mapped to $ 1 $, which is part of the stable manifold of the unstable fixed point at $ 0 $. For $ \lambda > 4 $, a portion of the interval $ I $ is mapped into the basin of attraction (cf. Chaos) for $ - \infty $, so the chaotic attractor is replaced by a chaotic transient. The mean time that typical initial conditions spend in $ I $ before escaping to $ - \infty $ scales as $ ( \lambda - 4) ^ {- 1/2 } $ with the parameter [a17].

Similar results hold in higher dimensions. [a17] discusses the Hénon mapping

$$ x _ {n+ 1 } = 1 - \lambda x _ {n} ^ {2} + y _ {n} , $$

$$ y _ {n+ 1 } = 0.3 x _ {n} , $$

where a $ 6 $- piece and a $ 4 $- piece chaotic attractor coexist for $ 1.0624 < \lambda < \lambda _ {c} \approx 1.08074 $. As $ \lambda \rightarrow \lambda _ {c} $, the $ 6 $- piece attractor approaches the stable manifold of a period-6 saddle orbit that forms part of the boundary of the basin of attraction for the $ 4 $- piece attractor. When $ \lambda > \lambda _ {c} $, part of the $ 6 $- piece attractor crosses this stable manifold. Thus, for $ \lambda > \lambda _ {c} $ the $ 6 $- piece attractor becomes a transient — eventually some iterate maps into the basin of the $ 4 $- piece attractor. Numerical evidence suggests that the mean time spent near the chaotic transient is proportional to $ ( \lambda - \lambda _ {c} ) ^ {- \gamma } $, where $ \gamma $ is a "critical exponent" that can be expressed in terms of the eigenvalues associated with the periodic orbit involved in the crisis [a18]. The existence of crises has also been demonstrated for the Lorenz equations and mappings of the torus [a17].

References

[a1] J. Milnor, "On the concept of attractor" Commun. Math. Phys. , 99 (1985) pp. 177–195
[a2] J. Guckenheimer, "Sensitive dependence to initial conditions for one-dimensional maps" Commun. Math. Phys. , 70 (1979) pp. 133–160
[a3] E.N. Lorenz, Physica D , 35 (1989) pp. 299–317
[a4] J.-P. Eckmann, "Roads to turbulence in dissipative dynamical systems" Rev. Mod. Phys. , 53 (1981) pp. 643–654
[a5a] M.J. Feigenbaum, "Qualitative universality for a class of nonlinear transformations" J. Stat. Phys. , 19 (1978) pp. 25–52
[a5b] M.J. Feigenbaum, "The universal metric properties of nonlinear transformations" J. Stat. Phys. , 21 (1979) pp. 669–706
[a6] P. Collet, J.-P. Eckmann, O.E. Lanford, "Universal properties of maps on an interval" Commun. Math. Phys. , 76 (1980) pp. 211–254
[a7] M. Jakobson, "Absolutely continuous invariant measures for one-parameter families of one-dimensional maps" Commun. Math. Phys. , 81 (1981) pp. 39–88
[a8] P. Collet, J.-P. Eckmann, H. Koch, "Period doubling bifurcations for families of maps on " J. Stat. Phys. , 25 (1981) pp. 1–14
[a9] J.A. Yorke, K.T. Alligood, "Period doubling cascades of attractors: a prerequisite for horseshoes" Commun. Math. Phys. , 101 (1985) pp. 305–321
[a10] P. Cvitanović (ed.) , Universality in chaos , A. Hilger (1989)
[a11] Y. Pomeau, P. Manneville, "Intermittent transition to turbulence in dissipative dynamical systems" Commun. Math. Phys. , 74 (1980) pp. 189–197
[a12] F. Argoul, A. Arneodo, P. Richetti, J.C. Roux, H.L. Swinney, Acct. Chem. Res. , 20 (1987) pp. 436–442
[a13] S.E. Newhouse, D. Ruelle, F. Takens, "Occurrence of strange axiom attractors near quasiperiodic flow on , " Commun. Math. Phys. , 64 (1978) pp. 35–40
[a14] M. Giglio, S. Musazzi, U. Perini, "Transition to chaotic behaviour via a reproducible sequence of period-doubling bifurcations" Phys. Rev. Lett. , 47 (1981) pp. 243–246
[a15a] P.R. Fenstermacher, H.L. Swinney, J.P. Gollub, "Dynamical instabilities and transition to chaotic Taylor vortex flow" J. Fluid Mech. , 94 (1979) pp. 103–128
[a15b] A. Brandstäter, H.L. Swinney, Phys. Rev. A , 35 (1987) pp. 2207–2220
[a16] L.D. Landau, E.M. Lifshitz, "Fluid mechanics" , Pergamon (1959) (Translated from Russian)
[a17] C. Grebogi, E. Ott, J.A. Yorke, "Crises, sudden changes in chaotic attractors, and transient chaos" Physica D , 7 (1983) pp. 181–200
[a18] C. Grebogi, E. Ott, J.A. Yorke, Phys. Rev. Lett. , 57 (1986) pp. 1284–1287
How to Cite This Entry:
Routes to chaos. Encyclopedia of Mathematics. URL: http://encyclopediaofmath.org/index.php?title=Routes_to_chaos&oldid=12583
This article was adapted from an original article by E.J. Kostelich (originator), which appeared in Encyclopedia of Mathematics - ISBN 1402006098. See original article