Nonlinear Dynamical Systems, Their Stability, and Chaos
Nonlinear Dynamical Systems, Their Stability, and Chaos
Amol Marathe
This introduction to nonlinear systems is written for students of fluid mechanics, so connections are made throughout the text to familiar fluid flow systems. The aim is to present how
nonlinear systems are qualitatively different from linear and to outline some simple procedures by which an understanding of nonlinear systems may be attempted. Considerable
attention is paid to linear systems in the vicinity of fixed points, and it is discussed why this
is relevant for nonlinear systems. A detailed explanation of chaos is not given, but a flavor
of chaotic systems is presented. The focus is on physical understanding and not on mathematical rigor. [DOI: 10.1115/1.4026864]
Introduction
(1)
describing the nondimensional streamfunction f g in the incompressible boundary layer over a solid wedge placed at angle bp to
a flow, is a third-order differential equation in a nondimensional
normal distance g from the solid wall. There are three boundary
conditions, f 0 f 0 0 0, and f 0 1 1. In a linear thirdorder system with three boundary conditions, we would get either
a unique solution or no solution. The FalknerSkan equation,
however, displays at least two solutions for each b < 0; an example is shown in Fig. 1. Such multiple solutions are not too unusual
in nonlinear boundary value problems such as this one. In contrast, dynamical systems are most often posed as initial value
problems. In an initial value problem, we have only now way to
march forward and will obtain a unique solution. However, the
fact that multiple solutions exist in boundary value problems is
relevant to us. To understand which of these solutions is manifested in a real flow, we must linearize the NavierStokes equations about each of these solutions and solve the resulting
dynamical system. Usually separated flows (where the flow next
to the wall is in the direction opposite to the average flow) such as
the profile shown by the red line (dashed) are very unstable
Rama Govindarajan
(2)
(3)
Second, a system such as that given by Eq. (2), where time t does
not appear explicitly, is called an autonomous system. An
n 1th order nonautonomous system, where time appears
C 2014 by ASME
Copyright V
(4)
so the system, if initially placed at the fixed point, stays there for
all time. In its immediate neighborhood, we may write
x x0 dxt, where the magnitude of the perturbation is small,
neglect terms containing powers of dx greater than 1 and obtain a
linear system in dx. We drop the d for convenience in future discussions of linear systems. From the eigenvalues of the linear system, the fixed point may be classified, for example, as a stable or
unstable spiral point, a stable or unstable node, a saddle point or a
center. There are also some other unusual fixed points that we
shall not discuss here. Simple equations that support fixed points
of each type, and phase portraits in the vicinity of the fixed-points,
are shown in Fig. 2. Such fixed points may be obtained for simple
systems on a plane, for example, by the MATLAB software available
on the Rice University website,1 as done here. Arrows may be
drawn on the trajectories shown in the phase portraits to show the
direction of time. Trajectories approach a stable (attracting) fixed
point, while they emerge out of an unstable or repelling point. The
reader may wish to attempt the following exercise.
Exercise (i) Work out the connection between the eigenvalues
of a system and the nature of the fixed point.
A simple example of a linear system is that of Fig. 3. The density q of the fluid shown is stratified so as to be heavier at the bottom. A small blob of fluid of density q0 is displaced along the
vertical direction z by a small distance dz to a location where the
density of the blob is different from that of the surrounding fluid
by dq dq=dzdz. Putting dz x1 , the dynamics may be given
by
1
x_1 x2
x_2 N 2 x1 ;
s
g dq
where N
q0 dz
(5)
(6)
a zero real part of the eigenvalues are nonhyperbolic. Nonhyperbolic fixed points are sometimes referred to as elliptic. The
HartmanGrobman theorem assures us that the behavior of a nonlinear system close to a hyperbolic fixed point is very similar to
that of the linear system that approximates it. A saddle-point, for
example, will remain a saddle in the nonlinear system as well, and
moreover, the invariant manifolds emanating from it will look
conformally the same as their linear approximates. Thus, a hyperbolic fixed point is robust when faced with the addition of nonlinear terms. However, an elliptic point, or center, of a linear system
is a fragile fixed point. The nonlinear system that it approximates
could display a stable or unstable spiral instead or change the local
flow in other ways. Once we know all the fixed points in a given
domain, if none of them is a center, we may then obtain the qualitative features of the nonlinear phase space by knowing the linear
behavior in each neighborhood.
Figure 4 is an example of a nonlinear system with several fixed
points in the domain shown. The equations solved are
x_ y y3 0:25x x3 y
y_ x y xy xy
(7)
(8)
Bifurcations
fork bifurcation is called by that name! In the supercritical pitchfork, a stable fixed point becomes unstable at the bifurcation
point, but two other stable fixed points are generated, as seen. The
subcritical pitchfork is the same except for a sign change in all the
growth rates. The two unstable fixed points can be stabilized (and
often are, in real systems) by a fifth-order term of the opposite
sign, and such systems display hysteresis. This is an interesting
feature of nonlinear systems that is not discussed further here, but
occurs in thermoacoustic [6] and many other systems.
In a continuation of exercises from the previous section:
Exercise (iv): Plot trajectories in phase space on either side of
(a) a transcritical bifurcation, (b) a saddle-node bifurcation, (c) a
supercritical pitchfork bifurcation, (d) a subcritical pitchfork
bifurcation with a higher-order stabilization.
Exercise (v): Plot trajectories in phase space on either side
(negative and positive r) of a Hopf bifurcation, given by the
equations
a_ ra a3 ;
Fig. 8 Phase portrait for a prototypical saddle-node bifurcation. The two fixed points annihilate each other.
h_ 1 a2
(9)
wires in the wind and several other situations. Wherever one finds
a self-sustained oscillation, one immediately thinks of the van der
Pol oscillator. Other reviews in this issue discuss flows that are
globally unstable, and it may be seen that they often display selfsustained oscillations of a characteristic frequency. The frequency
and the shape of the disturbance structure are independent of the
details of the disturbance environment. This is a completely nonlinear phenomenon, more commonly known as a stable limit cycle
in phase space. A van der Pol oscillator displays a limit cycle. The
frequency of this limit cycle may be estimated from the linear
approximation, as we shall see. This will give us a basic idea of
how to obtain the periodicity of a nonlinear system by expanding
about the linear answer.
The differential equation
Fig. 10 Top view of the boundary layer on a flat plate, sketch
of Roddam Narasimha. The thick vertical line marks the transition from periodic to chaotic flow.
Fig. 11 A schematic of the Craik mode of secondary instability: an example of spatial period doubling. The lines show maxima in the disturbance amplitude, which move downstream as
well as in the spanwise direction in a sinusoidal manner.
(10)
describes the van der Pol oscillator (see, e.g., Ref. [9]). This
belongs to a class of planar systems, i.e., systems modeled as
second-order ordinary differential equations, where a special periodic solution as described above, namely a limit cycle, exists in the
x-x_ plane. (Note that in the phase plane x r cos h.) In Fig. 12, the
origin is seen to be an unstable equilibrium point of Eq. (10). With
> 0, a trajectory starting sufficiently close to the origin (let us say
_ 0:5; 0, shown by the blue dashed-dotted line) will spiat x; x
ral outward away from the origin, since for this initial condition,
the damping term has a negative coefficient, and so acts to
_ On
amplify. Note that the damping is given by the coefficient of x.
the other hand a trajectory that starts with an initial condition suffi_ 5; 0, shown by the
ciently far from the origin, (e.g., at x; x
pink solid line in the figure), spirals inwards since the damping is
positive. The two trajectories cannot intersect each other, as discussed above. Both asymptotically approach the same isolated,
simple, closed curve in the phase-plane, in other words, the limit
cycle. This curve divides the phase plane into two regions. If
instead in Eq. (10) were to be less than 0, we would have an
unstable limit cycle. In this case, the origin will be stable and trajectories from the neighborhood of the limit cycle will tend to go
away from it as t ! 1.
Limit cycles are to be found in systems that are much more general than the van der Pol oscillator. How do we characterize a
limit cycle quantitatively and qualitatively? We need its shape
and size as well as its stability.
Fig. 12 Phase plane of van der Pol oscillator with 5 0.1. The
dash-dot trajectory emerges out of the origin and approaches
the limit cycle, while the pink line begins from infinity and
moves inwards towards the limit cycle. The limit cycle is where
the pink and blue trajectories meet.
(11)
(12)
(13)
x0 t A cost
_
x0
0
_
We choose x0
0, and we can do this without loss of generality,
as the van der Pol oscillator is autonomous. In other words, we can
choose the origin for time to correspond to a location anywhere on
the limit cycle, since the oscillator is autonomous. Assuming A to
be of O1; we write down ICs for x0 t; x1 t; as
x0 0 A;
x1 0 0;
x2 0 0;
x_0 0 0;
x_ 1 0 0;
x_ 2 0 0;
and
x2 1
5 2
A
12
2
Rt v
!
u
u
4
t1
1 et
R02
O2
dt
@T0
@T1
(16)
d2 @ 2
@ 2
2
O2
2
dt2
@T0 @T1
@T0
(17)
Substituting Eqs. (15) through (17) in Eq. (10) and expanding and
collecting terms, we obtain at O1 and at O
@ 2 X0
X0 0
@T02
(18)
@ 2 X1
@ 2 X0
@X0
X1 2
X02 1
@T1 @T0
@T0
@T02
(19)
2
A
sin T0
1
@T1
4
@T02
@B
B 2
2
2
B A B cos T0
@T1
4
A 2
B
A 3B2 sin 3T0 B2 3A2 cos 3T0
4
4
Removing the secular terms, i.e., equating coefficients of the resonant terms to zero, we get the slow flow
@A A A 2
A B2
@T1 2 8
@B B B 2
A B2
@T1 2 8
Transforming to polar coordinates A R cos /; B R sin / and
using
@A
A_
@T1
and
@B
B_
@T1
we get, up to O2 ,
R R3
and
R_
2
8
/_ 0
Slow flow equilibria correspond to periodic solutions of the system. For the van der Pol oscillator, these are
R 0 and
R2
(20)
For a 0, the oscillator is linear and has a unique steady state solution. For each set of initial conditions, the solution settles down
to a unique amplitude. Its frequency is the same as the forcing frequency and its amplitude depends upon the forcing amplitude c,
damping coefficient d and linear stiffness b. Nonlinear systems
(e.g., the Duffing oscillator in the a 6 0 case) differ from linear
systems as they may settle down to multiple periodic solutions,
stable or unstable, depending upon initial conditions. Figure 13
shows two such stable solutions found numerically. The frequency
of each is the same as the forcing frequency, but the solutions
differ in amplitudes, which are decided by their initial conditions,
_
_
here IC1, i.e., x0; x0
1; 0 and IC2, i.e., x0; x0
7; 0. Every stable periodic solution has its own basin of
attraction, i.e., a set of initial conditions from which the system
settles down to that particular periodic solution.
We now use the heuristic analytical but approximate method
of harmonic balance (MHB) in conjuction with a numerical
arclength-based continuation method to obtain the nonlinear harmonic response of the Duffing oscillator. MHB, a special case of
Galerkin projections, approximates periodic solutions of nonlinear
oscillators using a truncated Fourier series representation. Here,
we assume the periodic solution to be a single-term harmonic balance solution of the frequency of forcing [15],
MARCH 2014, Vol. 66 / 024802-7
Fig. 13 Multiple periodic solutions for the forced Duffing oscillator with a 5 1; d 5 0:1; b 5 1; c 5 1; x 5 2:5 with initial conditions
IC1 5 (1,0) (small amplitude) and IC2 5 (7,0) (large amplitude).
Fig. 14 Harmonic response of the Duffing equation for different strengths of nonlinearity a
x A sinxt B cosxt
Substituting the above into Eq. (20), multiplying by sinxt and
cosxt, then integrating with respect to t from 0 to 2p=x (which
is nothing but collecting coefficients of sinxt and cosxt), we
get two simultaneous nonlinear algebraic equations
3
3
Ax2 dxB bA aA3 aAB2 c 0
4
4
(21)
3
3
Bx2 dxA bB aB3 aA2 B 0
4
4
(22)
and
We can solve Eqs. (21) and (22) simultaneously for A and B for
some fixed values of other parameters. Using a numerical technique called arclength-based continuation (also called pseudo
arclength continuation
p method) [16, 17], we find the approximate
amplitude A2 B2 as a function of x, i.e., the nonlinear harmonic response of the system. Putting d 0 and c 0 in Eqs. 21
and 22, we obtain B 0 from Eq. (22) and then substituting so in
Eq. (21), we get
r
2 x2 1
(23)
A6
3
a
The above result though approximate in nature, tells us that for
undamped, unforced periodic oscillations, the frequency of oscillations (and, hence, the period) depends on the amplitude. As we
have discussed, this is typical of nonlinear systems. Equation (23)
is the equation of a backbone curve (the continuous curve in
Fig. 16). The case of a > 0 is called a stiffening nonlinearity while
a < 0 case is a softening nonlinearity [18].
Figures 1416 show the effect of varying nonlinearity strength
a, damping d and forcing amplitude c, respectively, on the frequency response of Eq. (20). As can be seen from Fig. 14, increasing a while holding other parameters constant, the resonance
curve along with the backbone curve leans over to the right for a
stiffening nonlinearity and to the left (not shown) for a softening
nonlinearity. Figure 15 shows the effect of varying the damping d
while keeping other parameters constant. The hump in the amplitude versus frequency curve follows the backbone curve in each
case, hence, the importance of the backbone curve. Decreasing d
raises the hump, i.e., raises the maximum possible response
024802-8 / Vol. 66, MARCH 2014
plot
for
parameter
values
%Function le dufng.m
function xdotdufng(t,x)
% dene variables globally,
%i.e., across the les
global gamma omega epsilon GAMMA OMEGA
%write 2nd order forced Dufng
%oscillator as two rst order ODEs
xdot(1)-gamma*x(1)omega 2*x(2)epsilon*x(2) 3GAMMA*cos(OMEGA*t);
xdot(2)x(1);
xdotxdot;
end
%Script le Poincaresection_script.m
% dene variables globally, i.e.,
%across the les
global gamma omega epsilon GAMMA OMEGA
%numerical values of the parameters involved
gamma0.1; omega1; epsilon0.25;
OMEGA2; GAMMA1.5;
%number of cycles
N50000;
%using ode45 numerical integration routine,
%integrate forced Dufng oscillator
[t x]ode45(dufng,
linspace(0,N*2*pi/OMEGA,100*N),[0 1]);
%plot Poincare section, x(nT)
%against xdot(nT)
Psectionplot(x(1:100:end,1),x(1:100:end,2),
r.);
%label the gure &
%set x & y label font and fontsize
set(gca,fontsize,21);
set(Psection,MarkerSize,2)
xlabel(x(nT)), ylabel(x(nT));
When we talked about the Klebanoff and Craik modes, the observant reader would have wondered how to study the instability
of an oscillatory flow. We do this by Floquet theory, and a canonical example is discussed below.
6.1 Floquet Theory. Consider an n-dimensional, linear nonautonomous system
x_ Atx
(24)
where the dot denotes differentiation with respect to time t, and
A(t) is periodic in time with a fixed time period T, i.e.,
At T At for all t. Let X(t) be a fundamental solution matrix (FSM) to this system, i.e.,
Xt x1 t; x2 t; ; xn t
where x1 t; x2 t; ; xn t are n linearly independent solutions
of Eq. (24).
It is straightforward to see that if x(t) is a solution to Eq. (24),
then so is x(t T). Since x(t) is a solution, we have
section for the forced Duffing oscillator for
Fig. 17 Poincare
d 0:1; b 1; a 0:25; c 1:38; x 2
d
xt Atxt
dt
MARCH 2014, Vol. 66 / 024802-9
y ut l cos h
where u(t) is the imposed displacement of the pivot. Differentiating with respect to time, we have
_
x_ l cos hh;
y_ u_ l sin hh_
Xt T XtC
~ we denote a special
Setting t 0, we get XT X0C. By Xt
~ I , where I denotes the n n
FSM to Eq. (24) such that X0
~
identity matrix. Then we have C XT
[9] and, therefore,
~
Xt T XtXT
(25)
i 1; ; n
(26)
~
ki is ith eigenvalue of XT.
Each zi t is a linear combination of
x1 t; x2 t; ; xn t. Equation (26) is a functional equation. We
assume its solution to be of the form
zi t kkt
i pi t
h0
h
l
l
zi t ki pi t
At the end of each time period, zi attains the values shown below:
t 0;
t T;
..
.
t mT;
zi 0 pi 0
zi T ki pi 0 ki zi 0
m
zi mT km
i pi 0 ki zi 0
~
Eigenvalues of XT,
i.e., ki s, determine the long-term behavior
of solutions to Eq. (24). If jki j > 1 for some i, then the corresponding solution grows unboundedly and the system is unstable.
If jki j < 1 for all i 1; ; n, then all solutions remain bounded
as t ! 1 and the system is stable. The matrix T is called the
Floquet matrix and its eigenvalues ki s are called Floquet
multipliers.
~ is diagonalizable.
We assume that Xt
(28)
with f t T f t for all t. Equation (28) is called Hills equation, which appeared in a memoir on the motion of the lunar perigee by Hill in 1886 [21].
We study the asymptotic stability of solutions to Eq. (28) with
the help of Floquet theory. Writing Eq. (28) as a system of two
first-order ordinary differential equations, we get
( ) "
#( )
0
1
x1
d x1
dt x2
x2
f t 0
_ By
where x1 x and x2 x.
"
Xt x1 t x2 t
x11 t
x12 t
x21 t
x22 t
dt x21 t x22 t
f t 0 x21 t x22 t
i.e., x_11 t x21 t;
f tx12 t
We have
x_ 12 t x22 t;
(29)
det Xt constant
%parameter values
delta 1;
epsilon 1;
%Mathieu equation written as
%two rst order ODEs
xdot[x(2);-(deltaepsilon*cos(t))*x(1)];
%Script le FloquetMultipliers_script.m
%set absolute and relative tolerances
optionsodeset(AbsTol,1e-7,RelTol,1e-8);
% Numerical integration of Mathieu equation
%for IC (1,0)
[t,x]ode45(mathieu,[0 2*pi],[1 0],options);
x1x(end,:);
~ det X0
~ 1 det XT
~
det Xt
According to Floquet theory, the asymptotic stability of solutions
to the Hills equation is governed by the Floquet multipliers, i.e.,
~
the eigenvalues of XT.
These satisfy the quadratic equation
~
10
k2 trXTk
~
where 1 det XT.
So the Floquet multipliers are
k1;2
q
~
~ 24
trXT6
trXT
2
~
Based on numerical value of trXT,
we have the following cases:
~
(1) jtrXTj
> 2: Both k1 and k2 are real. If k1 > 1, then
0 < k2 < 1 since k1 k2 1. The second possibility is
k1 < 1, so 0 > k2 > 1. Out of the two linearly independent solutions x1 t and x2 t, one grows exponentially
Applied Mechanics Reviews
Chaotic Systems
When we obtained a Poincare section with nonrepeating intersection points for the forced Duffing oscillator, we termed it a chaotic system but did not say why. When do we call the dynamics of
a system chaotic? The most often quoted property of a chaotic
system is that it is sensitive to initial conditions. By this term,
we mean that trajectories beginning from two initial conditions
that are arbitrarily close to each other will diverge exponentially,
so that, prescribing an initial condition to any degree of accuracy
is insufficient to predict the complete dynamics of the system. For
example, let us say Eq. (2) describes chaotic dynamics that we
wish to compute the value of x at all times and that we know the
initial condition x(0) correct to 10 decimal places. Even if we
have a completely accurate solver, our predictions will go very
wrong after some time t. If we know x(0) correct to 20 decimal
places, our predictions would be good for longer, but again at
some long time we would go very wrong. Chaotic dynamics thus
cannot be predicted completely. Such systems by definition will
display trajectories that are not periodic, because periodic dynamics, however long the period, is completely predictable once we
obtain the dynamics within one period. All chaotic systems, however, do contain many orbits that are periodic and have various
periods. In fact, one feature of a chaotic system is that it is dense
in periodic orbits. We note that these periodic orbits are unstable,
unlike limit cycles.
It is pertinent to mention here that not all dynamical systems
that are not periodic are chaotic. An Hamiltonian system of n
degrees of freedom and n integrals or constant of motion, will in
general display quasi-periodic motion on the surface of an n-torus.
Such a system is termed as integrable. Quasi-periodicity is created
by the existence of two or more time scales that are incommensurate (do not have a rational ratio). Trajectories winding around the
torus will cover its surface densely but will not return to the point
they started from. If on the other hand such a system has fewer
than n integrals, it is likely to display chaotic or nonintegrable
behavior. A convincing explanation of this is beyond the scope of
this article, but the interested reader is directed towards any good
text book on classical mechanics, such as Ref. [23].
The discussion above brings to light a fundamental difference
between a system that is deterministic but chaotic, and one that is
noisy. In the former, we know the equations governing the system
exactly but since the smallest changes in initial conditions take us
ultimately on completely different trajectories, we cannot predict
the dynamics completely. In a noisy system, we do not know the
equations of motion exactly, i.e., there are small external forces,
usually random, which produce changes in the dynamics. Of biggest interest to fluid dynamicists is the feature of topological mixing that chaotic systems possess. The effect is the same as when
024802-12 / Vol. 66, MARCH 2014
we mix sugar into tea. By this, every region of phase space will
overlap at some time or other with every other region.
When we talk ofunpredictability, we are usually comfortable
with the idea that a system of many variables, such as our lives,
shows unpredictable turns. On the other hand, when a system that
contains only a few independent unknowns (degrees of freedom)
displays chaotic behavior (three is the minimum number required
for chaos), it is very surprising. In turbulence, although there are
an infinite number of variables (the velocity, temperature, etc., at
each point in space is a variable), it seems that some aspects of
the physics can be reduced to that dominated by a much smaller
set of variables. We should, however, warn the reader of the following: at high Reynolds numbers, it is not clear what the dimensionality of turbulence is. Even at moderate Reynolds numbers,
we often need quite a high dimensionality in our models to get
reasonable answers. In other words, we have not clearly understood even what we need in order to understand turbulence as the
Reynolds number goes higher and higher. In spite of this, lowdimensional models are instructive to study and make a good
starting point toward the understanding of more complicated problems such as turbulence, and life. The section on the Lorenz system will describe the most famous example of low-dimensional
chaos, written for a toy weather problem.
Systems often go through a series of bifurcations to go from ordered to chaotic. The routes to chaos can most often be classified
under the period-doubling route, the RuelleTakensNewhouse
quasi-periodic route, the crisis route, and the intermittency route.
We will talk only about the first here. We had seen that period
doubling occurs in a Craik mode of disturbance growth in the
boundary layer, although it was not via a standard bifurcation.
The following section gives the most famous example of this
route to chaos. The similarity in detail of this simple system to
many far more complicated real systems is simply amazing. The
reader is encouraged to try out some of these.
7.1 The Logistic Map: A Prototype for Period-Doubling
Bifurcations. Human population growth seems unlimited even
today. It appears to be following what Malthus described as an exponential growth. Malthusian law does not take into consideration
the finiteness of natural resources. Verhulst suggested a different
law where growth rate is proportional to the existing population
and at the same time is limited by the carrying capacity of the
environment. Populations of almost all species appear to follow
this law. The discrete version of this law is the famous logistic
map
xn1 rxn 1 xn
(30)
This is a quadratic first-order difference equation. Population evolution as predicted by Eq. (30) strongly depends on the value of
the parameter r. For r < 2.992 trajectories evolve toward a fixed
point x*. The approach to the fixed point is monotonic when
r < 2.35 (refer to Fig. 20(a)). For r lying in the range
2.35 < r < 2.992, the trajectory still approaches its stable fixed
point, but in an oscillatory manner (Fig. 20(b)). As we increase r
further till it reaches the value 2.992, oscillation amplitudes
increase and the decay rate decreases. The fixed point x*
approaches a value of 0.643 as r approaches the numerical value
2.992 from below. As r increases steadily past the value of 2.992,
trajectories of increasing period, doubling at every bifurcation
point, are displayed. A period-k orbit is a trajectory that eventually
settles down to dynamics given by xnk xn for every n > n0 for
some finite n0. It may be given by, for example,
xn0 1 ; xn0 2 ; ; xn0 k . The period is 2 for r up to 3.448. Beyond
r 3.448, the dynamics changes from that on a period-2 orbit
(Fig. 20(c)) to that on a period-4 orbit (Fig. 20(d)). Again as r
crosses 3.544, the steady state changes from period-4 orbit to
period-8 orbit (Fig. 20(e)). Similarly as r crosses 3.568, the steady
state changes from period-8 orbit to period-16 orbit. At r
Transactions of the ASME
rar(i);
a0.1;
for j1:900
afeval(@logis,a);
end;
a(1)a;
for j2:length(r)
a(j)feval(@logis,a(j-1));
end
rpra*ones(1,length(r));
bdplotplot(rp,a,.);
end;
%font, fontsize of title and labels
set(gca,fontsize,18);
set(bdplot,MarkerSize,1)
xlabel(r), ylabel(steady state);
title(Bifurcation diagram of a Logistic map)
axis([2.6 4 0 1.1]);
Fig. 20 Trajectories of the logistic map for different values of r
Lorenz in his remarkable 1963 paper described a simple threedimensional approximation of an infinite-dimensional model of a
weather system, which is now famous as the Lorenz equations. It
describes the convective motion of a 2D fluid cell that is warmed
from below and cooled from above. The resulting partial differential equation is approximated as
x_ ry x
y_ rx y xz
z_ xy bz
where x physically describes the rate of convective overturning, y
is the horizontal temperature variation, and z the vertical temperature variation. Real positive parameters r and r are proportional to
the Prandtl number, Rayleigh number, while b is proportional to
physical dimensions of the system under consideration (Fig. 21).
Since the system is dissipative all trajectories eventually settle
down to a bounded set of zero volume lying in phase space. This
set may be an equilibrium point, a periodic orbit, or some complicated set of noninteger dimension. The last of these is called as a
strange attractor. Figure 22 shows a trajectory of the system starting with point (0.01,0,0) in the phase portrait. The trajectory is not
periodic. Continuing the numerical integration, the trajectory continues to wind around on one side and then switches from that
side to the other, and continues to wind around that side until the
next switch, without settling down to either periodic or stationary
behavior. This aspect is not a function of a particular choice of
6 br 1; 6 br 1; r 1 [24]. The origin loses its stability. At r 13:926 ; unstable limit cycles are created (due to
a homoclinic bifurcation, which we have not discussed). At this
point, a strange invariant set is born. For 1 < r < 24:74 ; the
other two equilibrium points are stable and the origin is unstable.
At r 24:74 ; a subcritical Hopf bifurcation occurs, the two
equilibrium points other than the origin lose their stability by
absorbing unstable limit cycles. As we increase r further, the Lorenz attractor is displayed and the system continues to be chaotic.
Finally as r is increased beyond 24:74 ; all three equilibrium
points are unstable, the system undergoes period doubling bifurcations and intermittent chaos is observed.
The dynamics on the strange invariant set shows sensitive dependence on initial conditions, usually quantified by Lyapunov
exponents. Figure 23 shows the divergence of two trajectories that
start quite close to each other. The figure is generated by running
script file sensitivity_IC_script.m, which calls function file Lorenz.m. The script file numerically integrates the Lorenz system
with two nearby initial conditions. Lyapunov exponents are a
measure of the exponential rate of divergence of two trajectories
that begin arbitrarily close to each other. A Lyapunov exponent
greater than zero signifies the system to be chaotic.
The aim of this exercise is to generate phase portraits of the
Lorenz system for various parameter values.
Exercise (xi): Generate phase portraits of the Lorenz equations
for the parameter values r 10:0; b 8=3 and for different r
values such as r 0:5; 2; 15 using Lorenz.m for initial conditions
such as (0.01,0,0).
024802-14 / Vol. 66, MARCH 2014
%Function le Lorenz.m
function ydotLorenz(t,y)
global z
% parameter values
sigma10;
r28;
b8/3;
% Lorenz system
ydot[sigma*(y(2)-y(1));y(1)*(r-y(3))-y(2);
y(1)*y(2)-b*y(3)];
%Script le sensitivity_IC_script.m
epsil0.001;
% numerical integration of Lorenz system
%for initial condition (1.5,0,0)
[t,x]ode45(Lorenz1,linspace(0,50,1000)
,[1.5 0 0]);
%plotting the solution
plot(t,x(:,1));
hold on;
% numerical integration of Lorenz system
%for initial condition (1.501,0,0)
[t,x]ode45(Lorenz1,linspace(0,50,1000)
,[1.5epsil 0 0]);
%plotting the solution over
the earlier solution
plot(t,x(:,1),r);
%Script le Buttery_script.m
%numerically integrate Lorenz
equations using ode45
[t,x]ode45(Lorenz1,linspace(0,50,30000)
,[0.01 0 0]);
%3-D plot of the phase space
Butteryplot3(x(:,1),x(:,2),x(:,3));
%setting font and fontsize of label and title
set(gca,fontsize,18); set(Buttery,
MarkerSize,5);
xlabel(x(t)), ylabel(y(t)),
zlabel(z(t));
title(Phase portrait for Lorenz equations)
axis([10,20,20,20,0,50]);
grid on;
s 0ek1 s ;
s k2 s
e ; ;
2s
jjsjj
n 1s kn s
e
ns
jjn 1sjj
We define k as
k lim
Consider two trajectories that start from nearby initial conditions. Let t be the distance between them. We write it in the linearized form
_ Jt
(31)
r
6
Jt 4 r zt
yt
r
1
xt
n
1X
n!1 ns
k1
lnjjksjj lim
n
1X
n!1 n
kk
k1
Figure 25 shows all the three Lyapunov exponents for the Lorenz
system. The largest is greater than zero and, therefore, the system
is chaotic. The system displays a strange attractor for the chosen
values of the parameters, in accordance with the largest exponent
being positive, the second one zero and the smallest negative. The
magnitude of the largest negative exponent is greater than that of
the largest positive one, given that the Lorenz system is dissipative. To learn more about the Lorenz system and related chaos
theory, it is suggested to the reader to study Ref. [24].
7
xt 5
b
Conclusions
Acknowledgment
Sharath Jose and Mamta Jotkar are thanked for their contributions. The anonymous referees have contributed significantly to
improving the article.
References
[1] Strogatz, S., 2001, Nonlinear Dynamics and Chaos: with Applications to Physics,
Biology, Chemistry and Engineering, 1st ed., Westview Press Boulder, CO.
[2] Schneider, T., Marinc, D., and Eckhardt, B., 2010, Localized Edge States Nucleate Turbulence in Extended Plane Couette Cells, J. Fluid Mech., 646, pp.
441451.
[3] Willis, A., and Kerswell, R., 2009, Turbulent Dynamics of Pipe Flow Captured
in a Reduced Model: pu? Relaminarization and Localized Edge States,J. Fluid
Mech., 619, pp. 213233.
[4] Duguet, Y., Schlatter, P., and Henningson, D. S., 2009, Localised Edge States
in Plane Couette Flow, Phys. Fluids, 21, p. 111701.
[5] Manneville, P., and Pomeau, Y., 2009, Transition to Turbulence, Scholarpedia, 4(3), p. 2072.
[6] Subramanian, P., Mariappan, S., Sujith, R. I., and Wahi, P., 2010, Bifurcation
Analysis of Thermoacoustic Instability in a Horizontal Rijke Tube, Int. J.
Spray Combus. Dyn., 2, pp. 325355.
[7] Gaster, M., 1969, Vortex Shedding From Slender Cones at Low Reynolds
Numbers, J. Fluid Mech., 38(3), pp. 565576.
[8] Venkatraman, D., 2013, Computational Models of Dorsal Coverts on Birds
Wings, Ph.D. thesis, University of Genova, Italy.
[9] Rand, R., Lecture notes on Nonlinear Vibrations. Version 52. Available at:
http://www.tam.cornell.edu/randdocs/ed
[10] Hinch, E. J., 1991, Perturbation Methods, Cambridge University Press, Cambridge, UK.
[11] Johnson, R. S., 2005, Singular Perturbation Theory, Springer ebooks.
[12] Holmes, M. H., 1991, In Introduction to Perturbation Methods, Springer-Verlag,
New York.
[13] Kevorkian, J., and Cole, J. D., 1996, Multiple Scale and Singular Perturbation
Methods, Springer-Verlag, New York.
[14] Mickens, R. E., 1981, An Introduction to Nonlinear Oscillations, 1st ed., Cambridge University Press Cambridge, UK.
[15] Mickens, R. E., 1996, Oscillations in Planar Dynamic Systems, World Scientific Publishing Co., River Edge, New Jersey.
[16] Chatterjee, A., 2002, An Elementary Continuation Technique, http://home.iitk.
ac.in/anindya/continuation.pdf
[17] Morgan, A., 2009, Solving Polynomial Systems using Continuation for Engineering and Scientific Problems, Classics in Applied Mathematics, SIAM.
[18] Stoker, J. J., 1950, Nonlinear Vibrations in Mechanical and Electrical Systems,
John Wiley & Sons, New York.
[19] Verhulst, F., 1996, Nonlinear Differential Equations and Dynamical Systems,
2nd ed., Springer, New York.
[20] Lanczos, C., 1986, The Variational Principles of Mechanics, Dover Publications, New York.
[21] Hill, G. W., 1886, On the Part of the Motion of the Lunar Perigee is a Function
of the Mean Motions of the Sun and the Moon, Acta Math., 8, pp. 136.
[22] Magnus, W., and Winkler, S., 2004, Hills equation, Dover Publications, Mineola, NY.
[23] Rana, N., and Joag, P., 2011, Classical Mechanics, 30th ed., Tata-McGraw
Hill.
[24] Sparrow, C., 1982, The Lorenz Equations: Bifurcations, Chaos and
Strange Attractors, Applied Mathematical Sciences 41, Springer-Verlag,
New York.
[25] Rugonyi, S., and Bathe, K., 2003, An Evaluation of the Lyapunov Characteristic Exponent of Chaotic Continuous Systems, Int. J. Numer. Methods Eng., 56,
pp. 145163.
[26] Wiggins, S., 2003, Introduction to Applied Nonlinear Dynamics and Chaos,
2nd ed., Springer, New York.
[27] Hale, J. K., 1963, Oscillations in Nonlinear Systems, 1st ed., Dover Publications
Mineola, NY.
[28] Drazin, P. G., 1992, Nonlinear Systems, 1st ed., Cambridge University Press,
Cambridge, UK.
[29] Grimshaw, R., 1991, Nonlinear Ordinary Differential Equations, Taylor &
Francis, London.
[30] Glendinning, P., 1994, Stability, Instability and Chaos, Cambridge University
Press, Cambridge, UK.