Algebraically Simple Chaotic Flows

20
INTERNATIONAL JOURNAL OF CHAOS THEORY AND APPLICATIONS Volume 5 (2000), No. 2 _________________________________________________________________________________________________ Invited paper Algebraically Simple Chaotic Flows J. C. Sprott Department of Physics, University of Wisconsin, Madison, WI 53706 USA Stefan J. Linz Theoretische Physik I, Institut für Physik, Universität Augsburg, D-86135 Augsburg, Germany Abstract 1 It came as a surprise to most scientists when Lorenz in 1963 discovered chaos in a simple system of three autonomous ordinary differential equations with two quadratic nonlinearities. This paper reviews efforts over the subsequent years to discover even simpler examples of chaotic flows. There is reason to believe that the algebraically simplest examples of chaotic flows with quadratic and piecewise linear nonlinearities have now been identified. The properties of these and other simple systems will be described. Keywords: chaos, flow, jerk, strange attractor, differential equations, fractal, bifurcation, circuits 1 Manuscript invited: November 27, 1999; accepted: ? ???, 2000. 1. Introduction Some aspects of chaos have been known for over a hundred years. Isaac Newton was said to get headaches thinking about the 3-body problem (Sun, Moon, and Earth). In 1890, King Oscar II of Sweden announced a prize for anyone who could solve the n-body problem and hence demonstrate stability of the solar system. The prize was awarded to Jules Henri Poincaré who showed that even the 3-body problem has no analytical solution [1, 2]. He went on to deduce many of the properties of chaotic systems including the sensitive dependence on initial conditions. With the successes of linear models in the sciences and the lack of powerful computers, the work of these early nonlinear dynamists went largely unnoticed and undeveloped for many decades. In 1963, Lorenz published a seminal paper [3] in which he showed that chaos can occur in systems of autonomous (no explicit time dependence) ordinary differential equations (ODEs) with as few as three variables and two quadratic nonlinearities. For continuous flows, the Poincaré-Bendixson theorem [4] implies the necessity of three variables, and chaos requires at least one nonlinearity. More explicitly, the theorem states that the long-time

Transcript of Algebraically Simple Chaotic Flows

Page 1: Algebraically Simple Chaotic Flows

INTERNATIONAL JOURNAL OF CHAOS THEORY AND APPLICATIONSVolume 5 (2000), No. 2

_________________________________________________________________________________________________

Invited paper

Algebraically Simple Chaotic Flows

J. C. SprottDepartment of Physics, University of Wisconsin, Madison, WI 53706 USA

Stefan J. LinzTheoretische Physik I, Institut für Physik, Universität Augsburg, D-86135 Augsburg, Germany

Abstract1

It came as a surprise to most scientists when Lorenz in 1963 discovered chaos in a simplesystem of three autonomous ordinary differential equations with two quadratic nonlinearities. Thispaper reviews efforts over the subsequent years to discover even simpler examples of chaoticflows. There is reason to believe that the algebraically simplest examples of chaotic flows withquadratic and piecewise linear nonlinearities have now been identified. The properties of theseand other simple systems will be described.

Keywords: chaos, flow, jerk, strange attractor, differential equations, fractal, bifurcation,circuits

1 Manuscript invited: November 27, 1999; accepted: ????, 2000.

1. Introduction

Some aspects of chaos have been known forover a hundred years. Isaac Newton was saidto get headaches thinking about the 3-bodyproblem (Sun, Moon, and Earth). In 1890,King Oscar II of Sweden announced a prize foranyone who could solve the n-body problemand hence demonstrate stability of the solarsystem. The prize was awarded to Jules HenriPoincaré who showed that even the 3-bodyproblem has no analytical solution [1, 2]. Hewent on to deduce many of the properties ofchaotic systems including the sensitive

dependence on initial conditions. With thesuccesses of linear models in the sciences andthe lack of powerful computers, the work ofthese early nonlinear dynamists went largelyunnoticed and undeveloped for many decades.

In 1963, Lorenz published a seminal paper[3] in which he showed that chaos can occur insystems of autonomous (no explicit timedependence) ordinary differential equations(ODEs) with as few as three variables and twoquadratic nonlinearities. For continuous flows,the Poincaré-Bendixson theorem [4] impliesthe necessity of three variables, and chaosrequires at least one nonlinearity. Moreexplicitly, the theorem states that the long-time

Page 2: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows2

limit of any “smooth” two-dimensional flow iseither a fixed point or a periodic solution. Withthe growing availability of powerfulcomputers, many other examples of chaoswere subsequently discovered in algebraicallysimple ODEs. Yet the sufficient conditions forchaos in a system of ODEs remain unknown.

This paper will review early serendipitousand insightful discoveries as well as laterextensive computer searches for thealgebraically simplest chaotic flows. There arereasons to believe that the simplest suchexamples with quadratic and piecewise linearnonlinearities have now been identified. Theproperties of these systems will be described.

2. Lorenz and Rössler

The celebrated Lorenz equations are:

bzxyz

yrxxzy

yxx

−=−+−=

σ+σ−=

&

&

&

(1)

where the dot denotes a first time derivative( dtdxx /=& , etc.). Note that there are seventerms on the right-hand side of these equations,two of which are nonlinear (xz and xy). Alsonote that there are three parameters, for whichLorenz found chaos with ó = 10, r = 28, and b= 8/3. The number of independent parametersis generally d+1 less than the number of termsfor a d-dimensional system, since each of thevariables (x, y, and z in this case) and time (t)can be arbitrarily rescaled. The Lorenz systemhas been extensively studied, and there is anentire book by Sparrow [5] devoted to it.

Although the Lorenz system is often takenas the prototypical chaotic flow, it is not thealgebraically simplest such system. In 1976,Rössler [6] proposed the following equations:

czxzbz

ayxy

zyx

−+=+=

−−=

&

&

&

(2)

This example also has seven terms and threeparameters, which Rössler took as a = b = 0.2and c = 5.7, but only a single quadraticnonlinearity (xz).

As recently as 1993, Lorenz [7] wrote: “Oneother study left me with mixed feelings. OttoRössler of the University of Tübingen hadformulated a system of three differentialequations as a model of a chemical reaction.By this time, a number of systems ofdifferential equations with chaotic solutionshad been discovered, but I felt I still had thedistinction of having found the simplest.Rössler changed things by coming along withan even simpler one. His record still stands.”

What Lorenz apparently did not realize wasthat Rössler himself had much earlier (in 1979)found an even simpler system [8] given by:

bzayayz

xy

zyx

−−=

=−−=

2&

&

&

(3)

This system has only six terms, a singlequadratic nonlinearity (y2), and twoparameters, giving chaos with a = 0.386 and b= 0.2. For some other values of the parameters,the dynamics are quasiperiodic, with atrajectory that lies on an invariant torus.

However, note that the simplicity of asystem can be measured in various ways.Algebraic simplicity is one such way;topological simplicity is another. The Rösslerattractor and most of the others in this paperare topologically simpler than the double-lobedattractor of Lorenz, but they are roughlyequivalent in that they all tend to resemble thesingle folded-band structure produced by Eq.(2).

3. Computer Search

Also unaware of the simpler Rösslerexample, Sprott [9] embarked on an extensivesearch for autonomous three-dimensionalchaotic systems with fewer than seven terms

Page 3: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)3

and a single quadratic nonlinearity and systemswith fewer than six terms and two quadraticnonlinearities. The brute-force method [10, 11]involved the numerical solution of about 108

systems of autonomous ODEs with randomlychosen real coefficients and initial conditions.The criterion for chaos was the existence of adecidedly positive Lyapunov exponent [12].

He found fourteen algebraically distinctcases with six terms and one nonlinearity, andfive cases with five terms and twononlinearities. One case was volume-conserving (conservative), and all the otherswere volume-contracting (dissipative),implying the existence of a strange attractor.Sprott provided a table of the spectrum ofLyapunov exponents, the Kaplan-Yorkedimension [13], and the types and eigenvaluesof the unstable fixed points for each of thenineteen cases. Interestingly, the Rösslerexample in Eq. (3) was not found, suggestingthat even this extensive search was notexhaustive.

Subsequently, Hoover [14] pointed out thatthe conservative case A found by Sprott

21 yz

yzxy

yx

−=

+−==

&

&

&

(4)

is a special case of the Nosé-Hooverthermostated dynamic system that had earlierbeen shown [15] to exhibit time-reversibleHamiltonian chaos. Note that this case ingeneral needs an adjustable parameter, but itturns out that chaos occurs for all coefficientsequal to unity, and so it is especially simple inthat sense. None of the fourteen cases with asingle quadratic nonlinearity share thatproperty, although there are two other chaoticcases with all unity coefficients and twoquadratic nonlinearities with strange attractors.Chaos is observed in Eq. (4) for only a smallrange of initial conditions, one choice of whichis (x, y, z) = (0, 5, 0). The other eighteen

chaotic cases were apparently previouslyunknown.

This search for algebraically simple chaoticsystems was an outgrowth of earlier studies[16] in which Sprott showed that three-dimensional ODEs with quadraticnonlinearities and bounded solutions arechaotic for 0.38±0.02 percent of the cases withuniform randomly chosen coefficients. Theprobability of chaos increases approximatelyas the square root of the dimension up to atleast d = 8 as shown in Fig. 1. Also shown inFig. 1 is the probability of chaos in systemsgoverned by difference equations (iteratedmaps), whose behavior is contrary to the caseof ODEs (flows) for reasons that are onlypartly understood. Whereas ODEs withquadratic nonlinearities require threedimensions to exhibit chaos, iterated maps canbe chaotic with only one dimension. Similarstudies [17] with randomly connected,discrete-time, artificial neural networks with ahyperbolic tangent squashing function showthat the probability of chaos is small at lowdimension and increases to nearly 100% at adimension of about 100. The relative rarity ofchaos in low-dimensional ODE systems is thereason chaos went largely unnoticed for solong and why new examples of simple chaoticsystems are still being discovered. In a relatedstudy [18] Sprott showed that the averagecorrelation dimension of chaotic d-dimensionalflows with quadratic nonlinearities anduniform randomly chosen coefficients isapproximately 1.07d 0.3 and the average largestLyapunov exponent is approximately 1.15d-0.84

4. Jerk Functions

In response to Sprott’s work, Gottlieb [19]pointed out that Eq. (4) could be recast in theexplicit third-order form

xxxxxx &&&&&&&&& /)(3 ++−= (5)

Page 4: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows4

which he called a “jerk function” since itinvolves a third derivative of x, which in amechanical system is the rate of change of theacceleration, sometimes called the “jerk” [20].It is known that any explicit ODE can be castin the form of a system of coupled first-orderODEs, but the converse does not hold ingeneral. Even if one can reduce the dynamicalsystem to a jerk form for each of the phasespace variables, the resulting differentialequations may look quite different. Gottliebasked the provocative question “What is thesimplest jerk function that gives chaos?”

Figure 1. Probability of chaos for mapsand flows of various dimensions.

One response was provided by Linz [21]who showed that the original Rössler model,the Lorenz model, and Sprott’s case R can bereduced to jerk forms, albeit of very differentcomplexity. The Rössler model in slightlymodified form can be written as

0)(

])1(1[

][2

=ε++ε+ε+ε+−ε−+

−ε+ε−+

ycy

yyyc

yyycy

&&

&&&&&&

(6)

where å (= a = b) = 0.2 and c = 5.7 giveschaos. Note that the jerk form of the Rösslerequation is a rather complicated quadraticpolynomial with 10 terms. Figure 2 shows that

the attractor in the yy −& phase space is the

familiar folded band.

Figure 2. Phase-space plot of the jerkrepresentation of the Rössler attractorfrom Eq. (6) with å = 0.2 and c = 5.7.

The Lorenz model in Eq. (1) can be writtenas

0)1(

]/)1()1([

]/1[

2

2

=−−σ−

σ+−+σ++

−+σ++

xxrb

xxxxb

xxxbx

&&

&&&&&&

(7)

The jerk form of the Lorenz equation is not apolynomial since it contains terms proportionalto xx /& as is typical of dynamical systems withmultiple nonlinearities. Its jerk form containseight terms. The phase-space plot of Eq. (7) inFig. 3 shows the familiar double-lobedattractor.

Linz showed that Sprott’s case R can bewritten as a polynomial with only five termsand a single quadratic nonlinearity

0=++−+ baxxxxx &&&&&& (8)

Page 5: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)5

with chaos for a = 0.9 and b = 0.4. Its attractoras shown in Fig. 4 is a folded band similar tothe Rössler attractor.

Figure 3. Phase-space plot of the jerkrepresentation of the Lorenz attractor fromEq. (7) with ó = 10, r = 28, and b = 8/3.

Figure 4. Attractor for Sprott’s Case Rfrom Eq. (8) with a = 0.9 and b = 0.4.

Clearly, the Lorenz and Rössler models arenot candidates for Gottlieb’s simplest jerk

function that gives chaos, and Sprott’s modelsdemonstrate the existence of much simplerexamples.

Meanwhile, Sprott also took up Gottlieb’schallenge and embarked on an extensivenumerical search for chaos in systems of theexplicit form ),,( xxxJx &&&&&& = , where the (jerk)function J is a simple quadratic or cubicpolynomial. He found a variety of cases [22],including two with three terms and twoquadratic nonlinearities in their jerk function,

02 =+−+ xxxaxx &&&&&& (9)

with a = 0.645 and

0=+−+ xxxxaxx &&&&&& (10)

with a = - 0.113, and a particularly simple casewith three terms and a single quadraticnonlinearity [23],

02 =+±+ xxxax &&&&&& (11)

with a = 2.017. Its attractor is shown in Fig. 5.For this value of a, the Lyapunov exponents(base-e) are (0.0550, 0, -2.0720) and theKaplan-Yorke dimension is DKY = 2.0265. Healso found systems of the form

)(xGxxax =++ &&&&&& (12)

where G(x) is a second-degree (or higher)polynomial such as x2 – b or x(x - b).

It is interesting to rewrite Eq. (11) as adynamical system in the variables x, y, and z:

xyazz

zy

yx

−±−=

==

2&

&

&

(13)

In this form, it is apparent that it has two fewerterms than the Lorenz or Rössler models andonly a single quadratic nonlinearity (y2). As a

Page 6: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows6

consequence, it is characterised by a singleparameter (a). It is unlikely that a simplerquadratic form exists because it would have noadjustable parameters. The number ofpossibilities is quite small, and a systematicnumerical check revealed that none of themexhibits chaos.

Figure 5. Attractor for the simplest chaoticflow with a quadratic nonlinearity fromEq. (11) with a = 2.017.

Equation (11) is simpler than any previouslydiscovered case, both in its jerk representationand in its representation as a dynamicalsystem. It was apparently overlooked in earliersearches because the range of a over whichchaos occurs is quite narrow (2.0168… < a <2.0577…). It also has a relatively small basinof attraction, so that initial conditions must bechosen carefully. One choice of initialconditions that lies in the basin of attraction is(x, y, z) = (0, 0, ±1), where the sign is chosenaccording to the sign of the y2 term in Eq. (13).

There is an alternate form for the simplestquadratic jerk function that can be written as

0=+±+ xxxxax &&&&&& (14)

but this case is equivalent to Eq. (11) to withina constant as can be seen by differentiating Eq.(11) with respect to time and then renamingthe variable x& to x.

These systems, and most of the others inthis paper, share a common route to chaos. Thecontrol parameter a can be considered adamping rate for the nonlinear oscillator. Forlarge values of a, there are one or more stableequilibrium points. As a decreases, a Hopfbifurcation occurs in which the equilibriumbecomes unstable, and a stable limit cycle isborn. The limit cycle grows in size until itbifurcates into a more complicated limit cyclewith two loops, which then bifurcates into fourloops, and so forth, in a sequence of perioddoublings until chaos finally onsets. A furtherdecrease in a causes the chaotic attractor togrow in size, passing through infinitely manyperiodic windows, and finally becomingunbounded when the attractor grows to touchthe boundary of its basin of attraction (a crisis).A bifurcation diagram for Eq. (11), which istypical, is shown in Fig. 6. In this figure, thelocal maxima of x are plotted as the damping ais gradually decreased. Note that the scales areplotted backwards to emphasise the similarityto the logistic map, xn+1 = Axn(1 – xn) [24].Indeed, a plot the maximum of x versus theprevious maximum shows an approximateparabolic dependence, albeit with a very small-scale fractal structure. No cases were foundwith a toroidal attractor. Apparently, three-dimensional systems with a single quadraticnonlinearity cannot produce toroidal attractors.

Sprott also found a variety of chaotic jerkfunctions with cubic nonlinearities, thesimplest of which have three terms and twononlinearities. Two examples are

023 =+++ axxxxx &&&&&& (15)

with a = 0.25 and

032 =+−+ xxxxax &&&&&& (16)

Page 7: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)7

with a = 3.6. He also found cases with fourterms and one nonlinearity, including the old(circa 1966), but little known, Moore-Spiegeloscillator [25]

0

)( 2

=++−++

Tx

xRxRTxx &&&&&&(17)

which models the inviscid convection of arotating fluid, where T is analogous to thePrandtl number times the Taylor number and Ris analogous to the Prandtl number times theRayleigh number. Moore and Spiegel reported“aperiodic behaviour” for T = 6 and R = 20;the term “chaos” was not coined until 1975[26].

Figure 6. Bifurcation diagram for Eq.(11) as the damping is reduced.

A phase-space plot of the Moore-Spiegelattractor for these parameters is shown in Fig.7. The regions in R-T space over which chaosoccurs are shown in Fig 8, in which chaos isassumed to exist if the largest Lyapunovexponent exceeds 0.005 after 4×105 fourth-order Runge-Kutta iterations with a step size of0.05. A more general equation of the sameform as Eq. (17) was derived by Auvergne andBaglin [27] to model the motion of theionisation zone of a star, and they also reported“irregular behaviour” and a broad powerspectrum.

Figure 7. Attractor for the Moore-Spiegel oscillator in Eq. (17) with T = 6and R = 20.

Sprott did not find dissipative chaotic jerkfunctions with fewer than four terms and asingle cubic nonlinearity. The absence of chaosin Eq. (11) with a cubic instead of quadraticnonlinearity is curious since it contradicts theconventional wisdom that increasing thenonlinearity enhances the likelihood of chaos.It is evident that a certain amount ofnonlinearity is required for chaos, but more isnot necessarily better.

Recently, Malasoma [28] joined the searchfor simple chaotic jerk functions and found acubic case as simple as Eq. (11) but of adifferent form

02 =+−+ xxxxax &&&&&& (18)

which exhibits chaos for a = 2.05. Its attractoris shown in Fig 9. For this value of a, theLyapunov exponents (base-e) are (0.0541, 0,-2.1041), and the Kaplan-Yorke dimension isDKY = 2.0257. This case follows the usualperiod-doubling route to chaos, culminating ina boundary crisis and unbounded solutions as ais lowered. The range of a over which chaos

Page 8: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows8

occurs is very narrow, 2.0278… < a <2.0840…, which probably explains why it wasmissed in earlier numerical searches. There isalso a second extraordinarily small window ofchaos for 0.0753514… < a < 0.0753624…,which is five thousand times smaller than theprevious case. Malasoma points out that thissystem is invariant under the paritytransformation xx −→ and speculates thatthis system is the simplest such example.

Figure 8. Regions of R-T space overwhich chaos occurs for the Moore-Spiegel oscillator in Eq. (17).

Both Linz and Sprott pointed out that if thejerk function is considered the time derivativeof an acceleration of a particle of mass m,Newton’s second law implies a force F whosetime derivative is dF/dt = mJ. If the force hasan explicit dependence on only x& , x, and time,it is considered to be “Newtonian jerky”. Thecondition for ),,( txxFF &= is that J depends

only linearly on x&& . In such a case the force ingeneral includes a memory term of the form

∫ ττ=t

dxGM ))(( (19)

that depends on the dynamical history of themotion.

Figure 9. Attractor for the simple cubicflow in Eq. (18) with a = 2.05.

The jerk papers by Linz and by Sprottappeared in the same issue of the AmericanJournal of Physics and prompted von Baeyer[29] to comment: “The articles with thosefunny titles are not only perfectly serious, butthey also illustrate in a particularly vivid waythe revolution that is transforming the ancientstudy of mechanics into a new science—onethat is not just narrowly concerned with themotion of physical bodies, but that deals withchanges of all kinds.” He goes on to say thatthe method of searching for chaos in a largeclass of systems “is not just emptymathematical formalism. Rather it illustratesthe arrival of a new level of abstraction inphysical science… At that higher level ofabstraction, dynamics has returned to theclassical Aristotelian goal of trying tounderstand all change.”

Eichhorn, Linz and Hänggi [30]summarized the situation with quadratic jerkfunctions that exhibit chaos. They used themethod of comprehensive Gröbner bases [31]

Page 9: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)9

to show that all the previously mentionedchaotic flows with a single quadraticnonlinearity and some of those with multiplenonlinearities can be organized into ahierarchy of quadratic jerk equations withincreasingly many terms. A slightly modifiedversion of their results is shown in Table 1with simplified parameters that produce chaos.Case JD0 (not included by Eichhorn, et. al.) is aspecial case of JD1 in which the constant termis zero. Such a categorization provides asimple means to compare the functionalcomplexity of different systems anddemonstrates the equivalence of cases nototherwise apparent. In a subsequent paper [32],the authors examined the simple cases JD1 andJD2 in more detail and identified the regions ofparameter space over which they exhibit chaos,albeit with different forms chosen so that thelocations of the fixed points are independent ofthe parameter values. In particular, JD1

becomes

xxbxxaxx &&&&&&& +−−−= (20)

and JD2 becomes

)4/1( 2 −−−−= xbxaxx &&&&&& (21)

They also derived criteria for functionalforms of the jerk function that cannot exhibitchaos. In particular, they showed that a jerkequation whose integral can be written as

∫ ττ=Ω+t

dxGxxx ))((),( &&& (22)

with G being either a positive or a negativesemidefinite function for all x cannot exhibitchaos. Moreover, if G is of the form

cxGxG += )(~

)( with a positive (negative)constant c and a positive (negative)

semidefinite function G~

, the dynamicseventually diverge for all initial values, exceptfor those that coincide with fixed points.

Table 1. Classification of simplechaotic polynomial jerk systems

Model Equation ParametersJD0 xxxax −+= 2&&&&&& a = -2.017

JD1

1−++=

xx

bxxax

&

&&&&& a = -1.8b = -2

JD2

12 −+

+=

x

xbxax &&&&&& a = - 0.5b= - 1.9

JD3

12 −++

+=

xxcx

xbxax

&

&&&&&& a = - 0.6b = -3c = 5

JD4

12 −++

+=

xxcx

xbxax

&&

&&&&&& a = - 0.6b = -2c = 3

JD5

xxx

bxxax

&&&

&&&&

−+

+=2

2 a = 0.5b = -1

JD6

12

2

−++

++=

xxxd

cxxbxax

&&&

&&&&&& a = -1b = -1c = 2d = 2

JD7

1

2

2

−+++

++=

xx

xexxd

cxxbxax

&&

&&

&&&&&& a = -1b = 1c = 2d = -3e = 1

Concurrently, Fu and Heidel [33], with atechnical correction by Gascon [34], provedthat all three-dimensional, dissipative,dynamical systems with quadraticnonlinearities and fewer than five terms cannotexhibit chaos. They subsequently extendedtheir results [35] to include conservative cases.More precisely, they rigorously proved theirresults for almost all conservative cases. The

lone exception, 22 zzz −= &&&& , appearsnumerically to have only periodic andunbounded solutions. Their work lendscredence to the claim that Eq. (11) is thesimplest chaotic flow with a quadraticnonlinearity.

Working independently, Thomas [36, 37]considered the terms in the Jacobian matrix of

Page 10: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows10

the system as feedback loops from which hewas able to deduce candidate chaotic systemsand the required signs of the coefficients. Onesuch system is

czxz

ayxy

zyx

−=

+=−−=

2&

&

&

(23)

with a = 0.385 and c = 2. This system can bereduced to the jerk form

02

)1()(2 =+−+

−+−+

cxaxxx

xacxacx

&

&&&&&&(24)

which by a transformation of variables isequivalent to case JD3 in Table 1. The x2 termin Eq. (23) can be replaced with othernonlinearities, including x3, tanh(x), sin(x), andsgn(x), yielding chaos for appropriateparameter values. These cases have multipleequilibrium points.

Thomas also proposed a case with a singleequilibrium (at the origin) given by

czxz

xy

zyaxx

−=

=−−=

2&

&

&

(25)

with a = 0.25 and c = 2. Its jerk form is

02

)1()(

=++−+−+

cxxx

xacxacx

&

&&&&&&(26)

which is a generalization of Eq. (14). By alinear transformation and rescaling, Eq. (26) isequivalent to case JD1 in Table 1. The x2 termin Eq. (25) can be replaced with x3, yieldingchaos for a = 3.3 and c = 4, and a jerk functionsimilar to but slightly more complicated thanEq. (18).

Thomas also suggested symmetric equationsof the form

)(

)(

)(

xfazz

zfayy

yfaxx

+−=+−=+−=

&

&

&

(27)

for which he found chaotic attractors for cubicpolynomial and sinusoidal functions f. Forexample, a = 0.18 and f(x) = sin(x) gives chaosas shown in Fig. 10. With f(x) = sin(x), he alsoobserved chaos in the conservative limit of a =0. This particularly simple and elegantexample has a trajectory that percolateschaotically within the infinite three-dimensional lattice of unstable steady states.Thomas calls this “labyrinth chaos.”Unfortunately, its jerk representation iscomplicated because of the threenonlinearities. Equation (27) is a special caseof the more general, cyclically symmetricsystem

),,(

),,(

),,(

yxzfz

xzyfy

zyxfx

===

&

&

&

(28)

which has chaotic solutions for many choicesof the nonlinear function f.

A convenient feature of chaotic jerkequations is that all three of the Lyapunovexponents can be determined from a numericalcalculation of only the largest exponent. Thisexponent (ë1) must be positive for chaos, andthere must be one zero exponent (ë2)corresponding to the direction of the flow.However, for a bounded system, the sum of theexponents ë1 + ë2 + ë3 is the rate of volumeexpansion averaged along the orbit and mustbe negative or zero and given by xJ &&∂∂ / ,which is equal to a for the Newtonian jerkycases JD0 though JD3 in Table 1. Hence thenegative exponent is easily found from ë3 = a- ë1. The Kaplan-Yorke dimension thenfollows from DKY = 2 - ë1/ ë3.

Page 11: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)11

Figure 10. Attractor for the symmetricsystem of Thomas in Eq. (27) with a =0.18 and f(x) = sin(x).

Cases whose exponents sum to zero areconservative, and their orbits fill a three-dimensional volume. Cases whose exponentssum to a negative value are dissipative, andthey have a strange attractor with a dimensionbetween 2 and 3. In either case, initialconditions must be chosen appropriately toensure that they are in the basin of attractionfor the dissipative systems and in the stochasticsea for the conservative systems. The emphasishere is on dissipative systems since they aremore mathematically tractable and are bettermodels of most natural systems.

5. Piecewise Linear Jerk Functions

Having found what appears to be thesimplest jerk function with a quadraticnonlinearity that leads to chaos, it is natural toask whether the nonlinearity can be weakened.

In particular, the 2x& in Eq. (11) might be

replaced with x& . A numerical search did not

reveal any such chaotic solutions.However, one can formulate the question

differently. Consider the system,

0=+±+ xxxaxb&&&&&& (29)

which is equivalent to Eq. (11) when b = 2. Forwhat values of a and b does this system exhibitchaos? Figure 11 shows the result of anumerical search in which chaos is assumed toexist if the largest Lyapunov exponent exceeds0.005 after 4×105 fourth-order Runge-Kuttaiterations with a step size of 0.05. There areindeed regions of chaos for 1 < b 2 as well asfor larger values of b. However, there do notappear to be chaotic solutions for b = 3 asmentioned earlier. The spiral structure of thechaotic region in a-b space begs for anexplanation. It appears that the region of chaosdoes not extend down to b = 1, but this isbecause systems with |x| + |y| + |z| > 106 havebeen discarded since they are consideredunbounded. In fact, the attractor size is foundto increase approximately as exp[1/(b-1)] for 1< b < 2.

Figure 11. Regions of a-b space forwhich chaos occurs in Eq. (29).

In an extensive numerical search for thealgebraically simplest dissipative chaotic flow

Page 12: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows12

with an absolute-value nonlinearity, Linz andSprott [38] discovered the case

01 =+−++ xxbxax &&&&&& (30)

which exhibits chaos for a = 0.6 and b = 1.Chaos also occurs if the signs of both the lasttwo terms are reversed, with an attractor that isa mirror image of the original about the x=0plane. The attractor for this case as shown inFig. 12 resembles the folded-band structure ofthe Rössler attractor.

Figure 12. Attractor for the simplestchaotic system with an absolute-valuenonlinearity from Eq. (30) with a = 0.6and b = 1.

For these parameters, the Lyapunovexponents (base-e) are (0.035, 0, - 635), andthe Kaplan-Yorke dimension is DKY = 2.055.The abrupt change in direction of the flow at x= 0 is not evident in the figure because thediscontinuity occurs only in the fourth timederivative of x.

The constant 1 in Eq. (30) affects only thesize of the attractor. Chaos exists for arbitrarilysmall values of this constant, but the attractorand its basin of attraction shrink

proportionally. This system exhibits a period-doubling route to chaos as shown in Fig. 13and otherwise resembles the quadratic chaoticjerk functions previously described. Thisexample relates to the quadratic flows as thetent map does to the logistic map. Linz andSprott claim it is the most elementarypiecewise linear chaotic flow and point out thatthe piecewise linear nature of the nonlinearityallows for an analytic solution to Eq. (30) bysolving two linear equations and matching theboundary conditions at x = 0. Figure 14 showsthe regions in a-b space for which chaosoccurs in Eq. (30), in which chaos is assumedto exist if the largest Lyapunov exponentexceeds 0.005 after 4×105 fourth-order Runge-Kutta iterations with a step size of 0.05.

Figure 13. Bifurcation diagram for Eq.(30) with b = 1 as the damping isreduced.

More recently, Linz [39] has proved thatchaos cannot exist in Eq. (30) if any of theterms are set to zero. He also notes that chaosis possible if the |x| term in Eq. (30) is replacedwith |xn|, |x|n, or x2n, with n a positive integer,or more generally with any inversionsymmetric function f(x) = f(-x) that is eitherpositive or negative for all x. Numericalexperiments indicate that chaotic solutionswith f(x) = |x|n exist for all nonzero n, includingnon-integer and negative values.

Page 13: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)13

Figure 14. Regions of a-b space forwhich chaos occurs in Eq. (30).

In a recent paper, Fischer, Weiler, Fröhlich,and Rössler [40] propose a piecewise linearsystem whose jerk representation can bewritten as

0=+−+++ cxxxbxax &&&&&& (31)

which exhibits chaos for a = c = 0.1 and b =0.3. It is intended to model an experimentalexample of chaos discovered in an electroniccircuit [41]. They claim their system is “amaximally simple 3-variable ODE with asingle letter-V shaped nonlinearity”.Unfortunately, it has one more term and onemore parameter than the case in Eq. (30), andthus it is not “maximally simple” in this sense,although the term x - |x| can be written as 2min(x, 0).

Equation (30) and (31) are special cases ofEq. (12) in which G(x) is a general nonlinearfunction with the properties described below.Integrating each term in Eq. (12) reveals thatthis system is a damped harmonic oscillatordriven by a nonlinear memory term thatinvolves the integral of G(x) as shown in Eq.(22). Such an equation often arises in the

feedback control of an oscillator in which theexperimentally accessible variable is atransformed and integrated version of thefundamental dynamical variable. Despite itsimportance and the richness of its dynamics,this system has been relatively little studied.Coullet, Tresser, and Arneodo observed chaosin numerical simulations with a cubicnonlinearity [42] of the form G(x) = bx(1 – x2)with a = 0.1 and b = 0.44 and with a specialpiecewise linear [43, 44] form

≥++−≤−≤−−−

=1 if

1 if

1 if

)(

xcbbx

xcx

xcbbx

xG (32)

with a = 0.1, b = 0.2061612, and c =0.2171604 that models a cubic nonlinearityand satisfies the Sil’nikov conditions for ahomoclinic orbit [45-47]. Thus it is one casefor which chaos can be rigorously provedrather than numerically indicated.

It does not appear generally known thatchaos accompanies many functions G(x), someexamples of which are listed in Table 2. Thesesystems are elementary, both in the sense ofhaving the algebraically simplest autonomousODE and in the form of the nonlinearity. Thetable lists typical values of b that give chaosfor arbitrary values of c with a = 0.6, alongwith the numerically calculated largestLyapunov exponents (LE) in base-e. Theconstant c is arbitrary and only affects the sizeof the attractor.

For bounded solutions, G(x) must averageto zero along the orbit, which means that anycontinuous G(x) must have at least one zero atx = x*. The stability of the fixed point at (x*, 0,0) is determined by the solutions of the

eigenvalue equation 023 =′−λ+λ+λ Ga ,where 'G = dG/dx, evaluated at x = x*. Thispoint is locally stable for –a 'G 0 andundergoes a Hopf bifurcation at 'G = -a,where ë = ± i. Thus, one would expect chaoticsystems of this form to require nonlinearity

Page 14: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows14

with either a positive slope at its zero crossing,or a large negative slope. Systems with 'G > 0apparently require at least two fixed points forchaos, but systems with 'G < -a only needone. All the cases in Table 2 have theseproperties. A scaling that preserves 'G and theshape of G(x) only affect the size of theattractor.

Table 2. Some simple functions G(x)that produce chaos in Eq. (12) with a =

0.6.

G(x) b LE)||( cxb −± 1.0 0.036

cxb +− )0,max( 6.0 0.093

)sgn(xcbx − 1.2 0.657

)sgn(xcbx +− 1.2 0.162

)/( 2 ccxb −± 0.58 0.073

)1/( 2 −cxbx 1.6 0.103

)1/( 2 −− cxbx 0.9 0.126

]/)tanh(2[ ccxxb −− 2.2 0.221

ccxb /)sin(± 2.7 0.069

ccxb /)cos(± 2.7 0.069

It is interesting to identify the maximallychaotic piecewise linear system. Of the casesin Table 2, the largest Lyapunov exponentsoccur for systems with )sgn()( xcbxxG −= .Using a variant of simulated annealing [48],the parameters a and b were adjusted tomaximise the Lyapunov exponent. The resultwas a = 0.55 and b = 2.84, for which theLyapunov exponents (base-e) are (1.055, 0,-1.655), giving an attractor with a Kaplan-Yorke dimension of DKY = 2.637. The attractoras shown in Fig. 15 is contained within anextremely thin torus that nearly touches theboundary of its small basin of attraction so thatinitial conditions must be chosen carefully toproduce bounded solutions. Initial conditionsthat suffice are (x, y, z) = (0.03, -0.33, -0.3).

Figure 15. Attractor for the maximallychaotic system given by Eq. (12) with a= 0.55 and G(x) = 2.84x – sgn(x).

It is also interesting to identify the leastnonlinear form of G(x) for which chaos occurs,which we take to mean the two-part piecewiselinear function with the smallest bend at theknee, è. Of the cases in Table 2, this conditionoccurs for G(x) = ±(b|x| - c) with a = 0.025 andb = 0.468, for which è (= 2tan-1b) is about50.2. The basin of attraction is very small, andthe chaotic attractor coexists with a nearbylimit cycle. Initial conditions that suffice are(x, y, z) = (0.9, 0, 0). Its attractor is shown inFig. 16.

The chaotic cases described above by nomeans exhaust the list of simple jerk functionswith chaotic solutions. In an extensive searchfor chaos in equations of the form

gxfexxd

xcxbxax

+ϕ++ϕ++ϕ+=

)()(

)(

&

&&&&&&&&(33)

where )(xϕ is one of a variety of simplenonlinear functions, several dozenalgebraically distinct cases were found withthree terms on the right-hand side, and severalhundred cases were found with four terms on

Page 15: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)15

the right-hand side. Cases whose coefficientshave different signs are considered distinct, butnot cases in which only the values differ. Table3 lists a small selection of those cases, chosento exclude ones that have already beendescribed, ones with multiple nonlinearities,and ones that are a superset of a simplerequation listed in the table. These cases havenot been carefully verified, and no values aregiven for the coefficients a and b that givechaotic solutions, except that they are positive.The coefficients are arbitrarily put into theleading terms. They are presented here toencourage further study and experimentation.Note that many of the cases are conservativeand those cases require a careful choice ofinitial conditions and tend to have very smallLyapunov exponents.

Figure 16. Attractor for the leastnonlinear chaotic system given by Eq.(12) with a = 0.025 and G(x) = 0.468|x| -1

6. Electrical Circuit Implementations

The piecewise linear jerk functionsdescribed above are ideally suited for

electronic implementation [49] because theycan be accurately represented with resistors,capacitors, diodes and operational amplifiers.The general procedure is to start with thequantity x&&&− at the circuit input, and thensuccessively generate x&& , x&− , and x withinverting integrators. These signals, perhapswith an additional constant voltage, are thenappropriately combined to form the function

),,( xxxJ &&& , which is then fed back to the input

of the circuit. Such a circuit can be considereda nonlinear oscillator with positive feedback.

Table 3. Some additional simplesystems with chaotic solutions forappropriate values of a and b (not

given).

xxxbxax −−−= &&&&&&&& )sgn(

1)cosh( −+−−= xbxxax &&&&

)1( −±−= xxax &&&&

)( 3xxxax −±−= &&&&

)1)0,max(( +−±−= xxbxax &&&&

)1)0,min(( −−±−= xxbxax &&&&

))sinh(( xxxax −±−= &&&&

2xxxax +±−= &&&&

1)cosh( +−±−− xbxxax &&&&

xxxbxax −−+−= 3&&&&&&&

xxxbxax −−+−= )sinh( &&&&&&&

xxxax −−−= )exp( &&&&&&

xxxbxax −+−−= )cos( &&&&&&&

12 −+−−= xxbxax &&&&&&

)cosh(xxxbxax −+−−= &&&&&&

)exp(xxxbxax ±+−−= &&&&&&

1)0,min( −−−−= xxbxax &&&&&&

)cosh(xxxbxax ±−−−= &&&&&&

)1)(cosh( −±−−= xxbxax &&&&&&

xxxbxax −±−−= )cosh( &&&&&&&

2xxxbxax −±−−= &&&&&&

xxbxax −−±−= )1)(cosh( &&&&&&

Page 16: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows16

As an example, Fig. 17 shows a circuit thatsolves Eq. (30) with b = 1. In this and thefollowing circuit, only the inverting input tothe operational amplifiers are shown; thenoninverting inputs are grounded. If the fixedresistors are 1 Ù, the capacitors are 1 F, andthe battery is 1 V, the circuit should work inreal time and should produce chaoticoscillations when the variable resistor isadjusted to a value of 1/a 1.67 Ù. However,the frequency at the first Hopf bifurcation at a= 1 is only 1/2ð Hz. A more practicalimplementation uses resistors of 1 kÙ andcapacitors of 0.1 ìF, giving a fundamentalfrequency of f = 104/2ð 1592 Hz at the firstbifurcation. This frequency is well into theaudio range so that the period doublings,periodic windows, and chaos, as shown in Fig.13, are easily heard in the signal x(t). Theperiod doublings are even more pronouncedwhen the signal x is integrated beforeamplification to enhance audibly the lowfrequencies.

Figure 17. A chaotic circuit usinginverting operational amplifiers and idealdiodes that solves Eq. (30) with b = 1.

The operation of the circuit in Fig. 17should be apparent to anyone with operationalamplifier design experience. However, it is notthe simplest circuit that solves Eq. (12) with G(x)as shown in Fig. 18 (a). One of the activeintegrators can be replaced with a passiveintegrator, and one of the diodes can beeliminated, resulting in a circuit with fifteencomponents rather than eighteen [50]. Elwakil

and Soliman [51] have also devised a chaoticoperational amplifier circuit with fifteencomponents using resistors, capacitors, anddiodes, but the equations required to model itare much more complicated.

Figure 18. Some functions G(x) in Eq.(12) that lead to chaos and are easilyimplemented electronically.

The functions shown in Table 2 suggestother nonlinear circuits. For example, thefunction sgn(x) is easily implemented with anoperational amplifier that has no feedback andthus acts as a comparator, abruptly switchingoutput from a large positive to a large negativevalue as the input voltage crosses zero. Specialoperational amplifier comparators are availablethat have orders of magnitude better frequencyresponse and slew rates than those designed forlinear operations. Figure 19 shows a circuitthat solves the equation

0)sgn( =+−++ xxxxax &&&&&& (34)

with a = 0.5, whose nonlinearity is of the formshown in Fig. 18 (b). In this circuit, capacitorsare in microfarads and all resistors are 1 kÙ.

This circuit has eleven components andmight be the simplest chaotic circuit using onlyinverting operational amplifiers, resistors andcapacitors. Its attractor is a single folded bandsimilar to the Rössler attractor and otherexamples previously discussed. Equation (34)with the last two signs reversed as in Fig. 18(c) also has chaotic solutions, with a double-scroll attractor similar to the Lorenz attractor,as shown in Fig. 20, but its circuitimplementation requires an additional

Page 17: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)17

inverting amplifier. There is not a directcorrespondence between simple equations andsimple circuits, although the identification ofone may suggest the other.

Figure 19. A chaotic circuit usinginverting saturating operationalamplifiers that solves Eq. (34) with a =0.5.

Figure 20. Attractor for the chaoticsystem given by Eq. (12) with a = 0.5and G(x) = sgn(x) – x.

For some purposes, such as where size andexpense are crucial, it may be important tominimize the required total capacitance and to

keep the resistances small so that the wholecircuit can be fabricated on a single chip.There is no fundamental reason why thiscannot be done, although the challenge wouldbe in simultaneously keeping the frequencylow. Note that the initial charge on the threecapacitors corresponds to the three initialconditions in the equations. Thus the circuitsmay not start properly if the initial values lieoutside the basin of attraction for the desiredsolution.

These circuits are similar in spirit to Chua’scircuit [52, 53] that uses two capacitors, aninductor, and diodes with operationalamplifiers or transistors to provide a piecewiselinear approximation to a cubic nonlinearity. Aversion of Chua’s circuit using saturatingoperational amplifiers is shown in Fig. 21. Theequations modelling this circuit can be reducedto

byz

zyxy

xhyax

−=+−=

−=

&

&

& ))((

(35)

which is simple in the sense of having sixterms and one nonlinearity, h(x), whose form issimilar to G(x) in Eq. (32). However, its jerkrepresentation is much more complicated:

)]()()([

)(

xbhxhxha

xabxx

++−

=−++&&&

&&&&&&(36)

Because of the discontinuities in h& and h&& , thedynamics are not continuous in the space of

),,( xxx &&& . Since the contraction is not constantalong the trajectory, it is more difficult toverify the Lyapunov exponents. Chua’s circuitis more difficult to construct, scale to arbitraryfrequencies, and analyse because of theinductor with its frequency-dependent losses,although a variant of Chua’s circuit with onlycapacitors is possible [54]. Three reactivecomponents (capacitors or inductors) are

Page 18: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows18

required for chaos in systems with continuousflows so that the Kirchhoff representation ofthe circuit contains three first-order ODEs.

Figure 21. A version of Chua’s circuitusing saturating operational amplifiers.

Systems involving delta functions(derivatives of the step function) and hysteretic(double-valued) functions can also lead tochaos. Since the flow is discontinuous for suchcases, chaos is possible with fewer than threevariables. A 2-D chaotic circuit with thirteencomponents based on this idea has beendeveloped by Tamaševièius, et al. [55].

A further 2-D chaotic example is providedby Dixon, et al. [56], in which the flow

)1(22

2

22

bbzzx

zz

axzx

xzx

−−−+

=

−+

=

&

&

(37)

is singular at the origin and all orbits are forcedto approach the singularity. It is probably notthe simplest such case. Such examples will notbe further discussed because they are usuallynot good models for natural phenomena.

7. Conclusions

Many autonomous chaotic systems havebeen discovered and studied that arealgebraically simpler than the Lorenz andRössler systems that are usually cited as

prototypical dissipative chaotic flows. Therepresentation of these systems in terms of asingle, autonomous, third-order, scalar ODE (ajerk equation) has simplified theiridentification and classification. Candidateequations have been found for the simplestsuch systems with quadratic and piecewiselinear nonlinearities, Eq. (11) and Eq. (30),respectively. Those systems with piecewiselinear functions are especially suited forelectronic implementation, and several suchcircuits have been described. The simplestsuch circuit may not yet have been identified.

References

[1] Peterson, I.: Newton's Clock: Chaos in theSolar System. W. H. Freeman, San Francisco,1993

[2] Barrow-Green, J.: Poincaré and the ThreeBody Problem. American MathematicalSociety, Providence, RI, 1997

[3] Lorenz, E. N.: Deterministic NonperiodicFlow. J. Atmos. Sci., vol. 20, pp. 130-141, 1963

[4] Hirsch, M. W., and Smale, S.: DifferentialEquations, Dynamical Systems and LinearAlgebra. Chapter 11, pp. 239-254, AcademicPress, New York, 1974

[5] Sparrow, C.: The Lorenz Equations:Bifurcations, Chaos, and Strange Attractors.Springer-Verlag, New York, 1982

[6] Rössler, O. E.: An Equation for ContinuousChaos. Phys. Lett. A, vol. 57, nr. 5, pp. 397-398, 1976

[7] Lorenz, E. N.: The Essence of Chaos. Chapter4, pp. 148, University of Washington Press,Seattle, 1993

[8] Rössler, O. E.: Continuous Chaos – FourPrototype Equations. Ann.(N.Y.) Acad. Sci., vol.316, pp. 376-392, 1979

[9] Sprott, J. C.: Some Simple Chaotic Flows.Phys. Rev. E, vol. 50, nr. 2, pp. R647-R650,1994

[10] Sprott, J. C.: Automatic Generation of StrangeAttractors. Comput. & Graphics, vol. 17, nr. 3,pp. 325-332, 1993

Page 19: Algebraically Simple Chaotic Flows

Sprott and Linz: Algebraically Simple Chaotic Flows JCTA 5 (2000)19

[11] Sprott, J. C.: Strange Attractors: CreatingPatterns in Chaos. M&T Books, New York,1993

[12] Wolf, A, Swift, J. B., Swinney, H. L., andVastano, J. A.: Determining LyapunovExponents from a Time Series. Physica D, vol.16, nr. 3, pp. 285-317, 1985

[13] Kaplan, J. L. and Yorke, J. A.: NumericalSolution of a Generalized Eigenvalue Problemfor Even Mapping. Peitgen, H. –O. andWalther, H. –O. (Eds.): Functional DifferentialEquations and Approximations of Fixed Points,Lecture Notes in Mathematics, vol. 730, pp.228-256, Springer-Verlag, Berlin, 1979

[14] Hoover, W. G.: Remark on “Some SimpleChaotic Flows”. Phys. Rev. E, vol. 51, nr. 1,759-760, 1995

[15] Posch, H. A., Hoover, W. G., and Vesely, F. J.:Canonical Dynamics of the Nosé Oscillator:Stability, Order, and Chaos. Phys. Rev. A, vol.33, nr. 6, pp. 4253-4265, 1986

[16] Sprott, J. C.: How Common is Chaos? Phys.Lett. A, vol. 173, nr. 1, pp. 21-24, 1993

[17] Dechert, W. D., Sprott, J. C., and Albers, D. J.:On the Probability of Chaos in LargeDynamical Systems: A Monte Carlo Study.Journal of Economic Dynamics & Control, vol.23, nr. 8, pp. 1197-1206, 1999

[18] Sprott, J. C.: Predicting the Dimension ofStrange Attractors. Phys. Lett. A, vol. 192, nr.5-6, pp. 355-360, 1994

[19] Gottlieb, H. P. W.: Question #38. What is thesimplest jerk function that gives chaos? Am. J.Phys., vol. 64, nr. 5, pp. 525, 1996

[20] Schot, S. H.: Jerk: The Time Rate of Change ofAcceleration. Am. J. Phys., vol. 46, nr. 11, pp.1090-1094, 1978

[21] Linz, S. J.: Nonlinear Dynamical Models andJerky Motion. Am. J. Phys., vol. 65. nr. 6, pp.523-526, 1997

[22] Sprott, J. C.: Some Simple Chaotic JerkFunctions. Am. J. Phys., vol. 65, nr. 6, pp. 537-543, 1997

[23] Sprott, J. C.: Simplest Dissipative ChaoticFlow. Phys. Lett. A, vol. 228, nr. 4-5, pp. 271-274, 1997

[24] May, R: Simple Mathematical Models withVery Complicated Dynamics. Nature vol. 261,nr. 5560, pp. 459-467, 1976

[25] Moore, D. W. and Spiegel, E. A.: A ThermallyExcited Non-Linear Oscillator. Astrophys. J.,vol. 143, nr. 3, pp. 871-887, 1966

[26] Li, T. Y, and Yorke, J. A.: Period ThreeImplies Chaos. Amer. Math. Monthly, vol. 82,nr. 10, pp. 985-992, 1975

[27] Auvergne, M. and Baglin, A.: A DynamicalInstability as a Driving Mechanism for StellarOscillations. Astron. Astrophysics, vol. 142, pp.388-392, 1985

[28] Malasoma, J. –M.: What is the SimplestDissipative Chaotic Jerk Equation which isParity Invariant? Phys. Lett. A, vol. 264, nr. 5,pp. 383-389

[29] von Baeyer, H. C.: All Shook Up. TheSciences, vol. 38, nr. 1, pp. 12-14, 1998

[30] Eichhorn, R., Linz, S. J., and Hänggi, P.:Transformations of Nonlinear DynamicalSystems to Jerky Motion and its Application toMinimal Chaotic Flows. Phys. Rev. E, vol. 58,nr. 6, pp. 7151-7164, 1998

[31] Becker, T. and Weispfenning, V.: GröbnerBases: A Computational Approach toCommutative Algebra. Springer, New York,1993

[32] Eichhorn, R., Linz, S. J., and Hänggi, P.:Simple Polynomial Chaotic Jerky Dynamics.Submitted for publication

[33] Fu, Z. and Heidel, J.: Non-Chaotic Behavior inThree-Dimensional Quadratic Systems.Nonlinearity, vol 10, nr. 5, pp. 1289, 1997

[34] Fu, Z. and Heidel, J.: Erratum: Non-ChaoticBehaviour in Three-Dimensional QuadraticSystems. Nonlinearity, vol. 12, nr. 3, pp. 739,1999

[35] Heidel, J, and Fu, Z.: Nonchaotic Behaviour inThree-Dimensional Quadratic Systems II. TheConservative Case. Nonlinearity, vol. 12, nr. 3,pp. 617-633, 1999

[36] Thomas, R.: Analyse et Synthèse de Systèmesà Dynamique Chaotique en Terme de Circuitsde Rétroaction (Feedback Circuits). Bull. de laClasse des Sciences, vol. VII, nr. 6, pp. 101-124, 1996

Page 20: Algebraically Simple Chaotic Flows

JCTA 5 (2000) Sprott and Linz: Algebraically Simple Chaotic Flows20

[37] Thomas, R: Deterministic Chaos Seen inTerms of Feedback Circuits: Analysis,Synthesis, “Labyrinth Chaos”. Int. J.Bifurcation and Chaos Appl. Sci. Eng., vol. 9,nr. 10, pp. 1889-1905, 1999

[38] Linz, S. J., and Sprott, J. C.: ElementaryChaotic Flow. Phys. Lett. A, vol. 259, nr. 3-4,pp. 240-245, 1999

[39] Linz, S. J., No-chaos Criteria for Certain JerkyDynamics. Submitted for publication

[40] Fischer, S., Weiler, A., Fröhlich, D., andRössler, O.: Kleiner_Attractor in a Piecewise-Linear C1-System. Z. Naturforsch, vol.54a, pp.268-269, 1999

[41] Kleiner, N.: Das Deterministische Chaos imExperiment. Krapp, H. and Wagenbaur, F.(Eds.): Komplexiat und Selbstorganisation, pp.259-266, 1997

[42] Coullet, P., Tresser, C., and Arneodo, A.:Transition to Stochasticity for a Class of ForcedOscillators. Phys. Lett. A, vol. 72, nr. 4-5, pp.268-270, 1979

[43] Arneodo, A., Coullet, P., and Tresser, C:Possible New Strange Attractors with SpiralStructure. Commun. Math. Phys., vol. 79, pp.573-579, 1981

[44] Arneodo, A, Coullet, P., and Tresser, C.:Oscillators with Chaotic Behavior: AnIllustration of a Theorem by Shil’nikov. J. Stat.Phys., vol. 27, nr. 1, pp. 171-182, 1982

[45] Sil’nikov, L. P.: A Case of the Existence of aDenumerable Set of Periodic Motions. Sov.Mat. Kokl., vol. 6, pp. 163-166, 1965

[46] Sil’nikov, L. P.: A Contribution to the Problemof the Structure of an Extended Neighborhoodof a Rough Equilibrium State of Saddle-FocusType. Math. USSR Sbornik., vol. 10, pp. 91-102, 1970

[47] Flendinning, P. and Sparrow, C.: Local andGlobal Behavior near Homoclinic Orbits. J.Stat. Phys., vol. 35, nr. 5, pp. 645-696, 1984

[48] Press, W. H., Teukolsky, S. A., Vetterling, W.T., and Flannery, B. P.: Numerical Recipes inC: The Art of Scientific Computing, Chapter 10,pp. 444-455, Cambridge University Press,Cambridge, 1993

[49] Sprott, J. C: Simple Chaotic Systems andCircuits. Am. J. Phys., in press

[50] Sprott, J. C.: A New Class of Chaotic Circuit.Phys. Lett. A, vol. 266, nr. 1., pp. 19-23, 2000

[51] Elwakil, A. S., and Soliman, A. M.: TwoModified for Chaos Negative ImpedanceConverter Op Amp Oscillators withSymmetrical and AntisymmetricalNonlinearities. Int. J. Bifurcation and ChaosAppl. Sci. Engr., vol. 8, nr. 6, pp. 1335-1346,1998

[52] Matsumoto, T., Chua, L. O., and Komoro, M:The Double Scroll. IEEE Trans. Circuits Syst.,svol. CAS-32, pp. 797-818, 1985

[53] Matsumoto, T., Chua, L. O., and Komoro, M:Birth and Death of the Double Scroll. PhysicaD, vol. 24, nr. 1-3, pp. 97-124, 1987

[54] Morgül, Ö.: Inductorless Realisation of ChuaOscillator. Electronics Letters, vol. 31, pp.1303-1304, 1995

[55] Tamaševièius, A., Mykolaitis, G., andNamajûnas, A.: Double Scroll in a Simple ‘2D’Chaotic Oscillator. Electronics Letters, vol. 32,pp. 1250-1251, 1996

[56] Dixon, D. D., Cummings, F. W., and Kaus, P.E.: Continuous “Chaotic” Dynamics in TwoDimensions. Physica D, vol. 65, nr. 1-2, pp.109-116, 1993