Chaos theory

Source: Wikipedia, the free encyclopedia.
Lorenz attractor
for values r = 28, σ = 10, b = 8/3
An animation of a double-rod pendulum at an intermediate energy showing chaotic behavior. Starting the pendulum from a slightly different initial condition would result in a vastly different trajectory. The double-rod pendulum is one of the simplest dynamical systems with chaotic solutions.

Chaos theory is an

fractals and self-organization.[2] The butterfly effect, an underlying principle of chaos, describes how a small change in one state of a deterministic nonlinear system can result in large differences in a later state (meaning there is sensitive dependence on initial conditions).[3] A metaphor for this behavior is that a butterfly flapping its wings in Brazil can cause a tornado in Texas.[4][5][6]

Small differences in initial conditions, such as those due to errors in measurements or due to rounding errors in

Chaos: When the present determines the future but the approximate present does not approximately determine the future.

Chaotic behavior exists in many natural systems, including fluid flow, heartbeat irregularities, weather and climate.

processes.

Introduction

Chaos theory concerns deterministic systems whose behavior can, in principle, be predicted. Chaotic systems are predictable for a while and then 'appear' to become random. The amount of time for which the behavior of a chaotic system can be effectively predicted depends on three things: how much uncertainty can be tolerated in the forecast, how accurately its current state can be measured, and a time scale depending on the dynamics of the system, called the Lyapunov time. Some examples of Lyapunov times are: chaotic electrical circuits, about 1 millisecond; weather systems, a few days (unproven); the inner solar system, 4 to 5 million years.[18] In chaotic systems, the uncertainty in a forecast increases exponentially with elapsed time. Hence, mathematically, doubling the forecast time more than squares the proportional uncertainty in the forecast. This means, in practice, a meaningful prediction cannot be made over an interval of more than two or three times the Lyapunov time. When meaningful predictions cannot be made, the system appears random.[19]

Chaos theory is a method of qualitative and quantitative analysis to investigate the behavior of dynamic systems that cannot be explained and predicted by single data relationships, but must be explained and predicted by whole, continuous data relationships.

Chaotic dynamics

mod
1 displays sensitivity to initial x positions. Here, two series of x and y values diverge markedly over time from a tiny initial difference.

In common usage, "chaos" means "a state of disorder".[20][21] However, in chaos theory, the term is defined more precisely. Although no universally accepted mathematical definition of chaos exists, a commonly used definition, originally formulated by Robert L. Devaney, says that to classify a dynamical system as chaotic, it must have these properties:[22]

  1. it must be
    sensitive to initial conditions
    ,
  2. it must be topologically transitive,
  3. it must have
    periodic orbits
    .

In some cases, the last two properties above have been shown to actually imply sensitivity to initial conditions.[23][24] In the discrete-time case, this is true for all continuous maps on metric spaces.[25] In these cases, while it is often the most practically significant property, "sensitivity to initial conditions" need not be stated in the definition.

If attention is restricted to intervals, the second property implies the other two.[26] An alternative and a generally weaker definition of chaos uses only the first two properties in the above list.[27]

Sensitivity to initial conditions

Lorenz equations used to generate plots for the y variable. The initial conditions for x and z were kept the same but those for y were changed between 1.001, 1.0001 and 1.00001. The values for , and were 45.91, 16 and 4 respectively. As can be seen from the graph, even the slightest difference in initial values causes significant changes after about 12 seconds of evolution in the three cases. This is an example of sensitive dependence on initial conditions.

Sensitivity to initial conditions means that each point in a chaotic system is arbitrarily closely approximated by other points that have significantly different future paths or trajectories. Thus, an arbitrarily small change or perturbation of the current trajectory may lead to significantly different future behavior.[2]

Sensitivity to initial conditions is popularly known as the "

Edward Lorenz in 1972 to the American Association for the Advancement of Science in Washington, D.C., entitled Predictability: Does the Flap of a Butterfly's Wings in Brazil set off a Tornado in Texas?.[28]
The flapping wing represents a small change in the initial condition of the system, which causes a chain of events that prevents the predictability of large-scale phenomena. Had the butterfly not flapped its wings, the trajectory of the overall system could have been vastly different.

As suggested in Lorenz's book entitled The Essence of Chaos, published in 1993,[5] "sensitive dependence can serve as an acceptable definition of chaos". In the same book, Lorenz defined the butterfly effect as: "The phenomenon that a small alteration in the state of a dynamical system will cause subsequent states to differ greatly from the states that would have followed without the alteration." The above definition is consistent with the sensitive dependence of solutions on initial conditions (SDIC). An idealized skiing model was developed to illustrate the sensitivity of time-varying paths to initial positions.[5] A predictability horizon can be determined before the onset of SDIC (i.e., prior to significant separations of initial nearby trajectories).[29]

A consequence of sensitivity to initial conditions is that if we start with a limited amount of information about the system (as is usually the case in practice), then beyond a certain time, the system would no longer be predictable. This is most prevalent in the case of weather, which is generally predictable only about a week ahead.

geologic era
), but we cannot predict exactly which day will have the hottest temperature of the year.

In more mathematical terms, the Lyapunov exponent measures the sensitivity to initial conditions, in the form of rate of exponential divergence from the perturbed initial conditions.[31] More specifically, given two starting trajectories in the phase space that are infinitesimally close, with initial separation , the two trajectories end up diverging at a rate given by

where is the time and is the Lyapunov exponent. The rate of separation depends on the orientation of the initial separation vector, so a whole spectrum of Lyapunov exponents can exist. The number of Lyapunov exponents is equal to the number of dimensions of the phase space, though it is common to just refer to the largest one. For example, the maximal Lyapunov exponent (MLE) is most often used, because it determines the overall predictability of the system. A positive MLE is usually taken as an indication that the system is chaotic.[8]

In addition to the above property, other properties related to sensitivity of initial conditions also exist. These include, for example, measure-theoretical mixing (as discussed in ergodic theory) and properties of a K-system.[11]

Non-periodicity

A chaotic system may have sequences of values for the evolving variable that exactly repeat themselves, giving periodic behavior starting from any point in that sequence. However, such periodic sequences are repelling rather than attracting, meaning that if the evolving variable is outside the sequence, however close, it will not enter the sequence and in fact, will diverge from it. Thus for almost all initial conditions, the variable evolves chaotically with non-periodic behavior.

Topological mixing

Six iterations of a set of states passed through the logistic map. The first iterate (blue) is the initial condition, which essentially forms a circle. Animation shows the first to the sixth iteration of the circular initial conditions. It can be seen that mixing occurs as we progress in iterations. The sixth iteration shows that the points are almost completely scattered in the phase space. Had we progressed further in iterations, the mixing would have been homogeneous and irreversible. The logistic map has equation . To expand the state-space of the logistic map into two dimensions, a second state, , was created as , if and otherwise.
topological mixing
. Here, the blue region is transformed by the dynamics first to the purple region, then to the pink and red regions, and eventually to a cloud of vertical lines scattered across the space.

Topological mixing (or the weaker condition of topological transitivity) means that the system evolves over time so that any given region or open set of its phase space eventually overlaps with any other given region. This mathematical concept of "mixing" corresponds to the standard intuition, and the mixing of colored dyes
or fluids is an example of a chaotic system.

Topological mixing is often omitted from popular accounts of chaos, which equate chaos with only sensitivity to initial conditions. However, sensitive dependence on initial conditions alone does not give chaos. For example, consider the simple dynamical system produced by repeatedly doubling an initial value. This system has sensitive dependence on initial conditions everywhere, since any pair of nearby points eventually becomes widely separated. However, this example has no topological mixing, and therefore has no chaos. Indeed, it has extremely simple behavior: all points except 0 tend to positive or negative infinity.

Topological transitivity

A map is said to be topologically transitive if for any pair of non-empty open sets , there exists such that . Topological transitivity is a weaker version of

topological mixing. Intuitively, if a map is topologically transitive then given a point x and a region V, there exists a point y near x whose orbit passes through V. This implies that it is impossible to decompose the system into two open sets.[32]

An important related theorem is the Birkhoff Transitivity Theorem. It is easy to see that the existence of a dense orbit implies topological transitivity. The Birkhoff Transitivity Theorem states that if X is a second countable, complete metric space, then topological transitivity implies the existence of a dense set of points in X that have dense orbits.[33]

Density of periodic orbits

For a chaotic system to have

periodic orbits means that every point in the space is approached arbitrarily closely by periodic orbits.[32] The one-dimensional logistic map
defined by x → 4 x (1 – x) is one of the simplest systems with density of periodic orbits. For example,  →  → (or approximately 0.3454915 → 0.9045085 → 0.3454915) is an (unstable) orbit of period 2, and similar orbits exist for periods 4, 8, 16, etc. (indeed, for all the periods specified by Sharkovskii's theorem).[34]

Sharkovskii's theorem is the basis of the Li and Yorke[35] (1975) proof that any continuous one-dimensional system that exhibits a regular cycle of period three will also display regular cycles of every other length, as well as completely chaotic orbits.

Strange attractors

Lorenz attractor
displays chaotic behavior. These two plots demonstrate sensitive dependence on initial conditions within the region of phase space occupied by the attractor.

Some dynamical systems, like the one-dimensional logistic map defined by x → 4 x (1 – x), are chaotic everywhere, but in many cases chaotic behavior is found only in a subset of phase space. The cases of most interest arise when the chaotic behavior takes place on an attractor, since then a large set of initial conditions leads to orbits that converge to this chaotic region.[36]

An easy way to visualize a chaotic attractor is to start with a point in the

Lorenz
weather system. The Lorenz attractor is perhaps one of the best-known chaotic system diagrams, probably because it is not only one of the first, but it is also one of the most complex, and as such gives rise to a very interesting pattern that, with a little imagination, looks like the wings of a butterfly.

Unlike

strange attractors, have great detail and complexity. Strange attractors occur in both continuous dynamical systems (such as the Lorenz system) and in some discrete systems (such as the Hénon map). Other discrete dynamical systems have a repelling structure called a Julia set, which forms at the boundary between basins of attraction of fixed points. Julia sets can be thought of as strange repellers. Both strange attractors and Julia sets typically have a fractal structure, and the fractal dimension
can be calculated for them.

Coexisting attractors

Coexisting chaotic and non-chaotic attractors within the generalized Lorenz model.[37][38][39] There are 128 orbits in different colors, beginning with different initial conditions for dimensionless time between 0.625 and 5 and a heating parameter r = 680. Chaotic orbits recurrently return close to the saddle point at the origin. Nonchaotic orbits eventually approach one of two stable critical points, as shown with large blue dots. Chaotic and nonchaotic orbits occupy different regions of attraction within the phase space.

In contrast to single type chaotic solutions, recent studies using Lorenz models [40][41] have emphasized the importance of considering various types of solutions. For example, coexisting chaotic and non-chaotic may appear within the same model (e.g., the double pendulum system) using the same modeling configurations but different initial conditions. The findings of attractor coexistence, obtained from classical and generalized Lorenz models,[37][38][39] suggested a revised view that "the entirety of weather possesses a dual nature of chaos and order with distinct predictability", in contrast to the conventional view of "weather is chaotic".

Minimum complexity of a chaotic system

Bifurcation diagram of the logistic map xr x (1 – x). Each vertical slice shows the attractor for a specific value of r. The diagram displays period-doubling as r increases, eventually producing chaos. Darker points are visited more frequently.

Discrete chaotic systems, such as the

nonlinear
or infinite-dimensional.

The Poincaré–Bendixson theorem states that a two-dimensional differential equation has very regular behavior. The Lorenz attractor discussed below is generated by a system of three differential equations such as:

where , , and make up the

system state
, is time, and , , are the system
Rössler equations, which have only one nonlinear term out of seven. Sprott[42] found a three-dimensional system with just five terms, that had only one nonlinear term, which exhibits chaos for certain parameter values. Zhang and Heidel[43][44] showed that, at least for dissipative and conservative quadratic systems, three-dimensional quadratic systems with only three or four terms on the right-hand side cannot exhibit chaotic behavior. The reason is, simply put, that solutions to such systems are asymptotic
to a two-dimensional surface and therefore solutions are well behaved.

While the Poincaré–Bendixson theorem shows that a continuous dynamical system on the Euclidean plane cannot be chaotic, two-dimensional continuous systems with non-Euclidean geometry can still exhibit some chaotic properties.[45] Perhaps surprisingly, chaos may occur also in linear systems, provided they are infinite dimensional.[46] A theory of linear chaos is being developed in a branch of mathematical analysis known as functional analysis.

The above set of three ordinary differential equations has been referred to as the three-dimensional Lorenz model.[47] Since 1963, higher-dimensional Lorenz models have been developed in numerous studies[48][49][37][38] for examining the impact of an increased degree of nonlinearity, as well as its collective effect with heating and dissipations, on solution stability.

Infinite dimensional maps

The straightforward generalization of coupled discrete maps[50] is based upon convolution integral which mediates interaction between spatially distributed maps: ,

where kernel is propagator derived as Green function of a relevant physical system,[51] might be logistic map alike or

complex map. For examples of complex maps the Julia set
or Ikeda map may serve. When wave propagation problems at distance with wavelength are considered the kernel may have a form of Green function for Schrödinger equation:.[52][53]

.

Jerk systems

In

position
, with respect to time. As such, differential equations of the form

are sometimes called jerk equations. It has been shown that a jerk equation, which is equivalent to a system of three first order, ordinary, non-linear differential equations, is in a certain sense the minimal setting for solutions showing chaotic behavior. This motivates mathematical interest in jerk systems. Systems involving a fourth or higher derivative are called accordingly hyperjerk systems.[54]

A jerk system's behavior is described by a jerk equation, and for certain jerk equations, simple electronic circuits can model solutions. These circuits are known as jerk circuits.

One of the most interesting properties of jerk circuits is the possibility of chaotic behavior. In fact, certain well-known chaotic systems, such as the Lorenz attractor and the

Rössler map
, are conventionally described as a system of three first-order differential equations that can combine into a single (although rather complicated) jerk equation. Another example of a jerk equation with nonlinearity in the magnitude of is:

Here, A is an adjustable parameter. This equation has a chaotic solution for A=3/5 and can be implemented with the following jerk circuit; the required nonlinearity is brought about by the two diodes:

In the above circuit, all resistors are of equal value, except , and all capacitors are of equal size. The dominant frequency is . The output of op amp 0 will correspond to the x variable, the output of 1 corresponds to the first derivative of x and the output of 2 corresponds to the second derivative.

Similar circuits only require one diode[55] or no diodes at all.[56]

See also the well-known Chua's circuit, one basis for chaotic true random number generators.[57] The ease of construction of the circuit has made it a ubiquitous real-world example of a chaotic system.

Spontaneous order

Under the right conditions, chaos spontaneously evolves into a lockstep pattern. In the Kuramoto model, four conditions suffice to produce synchronization in a chaotic system. Examples include the

Josephson junctions.[58]

Moreover, from the theoretical physics standpoint, dynamical chaos itself, in its most general manifestation, is a spontaneous order. The essence here is that most orders in nature arise from the spontaneous breakdown of various symmetries. This large family of phenomena includes elasticity, superconductivity, ferromagnetism, and many others. According to the supersymmetric theory of stochastic dynamics, chaos, or more precisely, its stochastic generalization, is also part of this family. The corresponding symmetry being broken is the topological supersymmetry which is hidden in all stochastic (partial) differential equations, and the corresponding order parameter is a field-theoretic embodiment of the butterfly effect.[59]

History

Barnsley fern created using the chaos game. Natural forms (ferns, clouds, mountains, etc.) may be recreated through an iterated function system (IFS).

Hadamard's billiards".[66] Hadamard was able to show that all trajectories are unstable, in that all particle trajectories diverge exponentially from one another, with a positive Lyapunov exponent
.

Chaos theory began in the field of

Mary Lucy Cartwright and John Edensor Littlewood,[71] and Stephen Smale.[72] Except for Smale, these studies were all directly inspired by physics: the three-body problem in the case of Birkhoff, turbulence and astronomical problems in the case of Kolmogorov, and radio engineering in the case of Cartwright and Littlewood.[citation needed
] Although chaotic planetary motion had not been observed, experimentalists had encountered turbulence in fluid motion and nonperiodic oscillation in radio circuits without the benefit of a theory to explain what they were seeing.

Despite initial insights in the first half of the twentieth century, chaos theory became formalized as such only after mid-century, when it first became evident to some scientists that

linear theory, the prevailing system theory at that time, simply could not explain the observed behavior of certain experiments like that of the logistic map. What had been attributed to measure imprecision and simple "noise" was considered by chaos theorists as a full component of the studied systems. In 1959 Boris Valerianovich Chirikov proposed a criterion for the emergence of classical chaos in Hamiltonian systems (Chirikov criterion). He applied this criterion to explain some experimental results on plasma confinement in open mirror traps.[73][74] This is regarded as the very first physical theory of chaos, which succeeded in explaining a concrete experiment. And Boris Chirikov himself is considered as a pioneer in classical and quantum chaos.[75][76][77]

The main catalyst for the development of chaos theory was the electronic computer. Much of the mathematics of chaos theory involves the repeated iteration of simple mathematical formulas, which would be impractical to do by hand. Electronic computers made these repeated calculations practical, while figures and images made it possible to visualize these systems. As a graduate student in Chihiro Hayashi's laboratory at Kyoto University, Yoshisuke Ueda was experimenting with analog computers and noticed, on November 27, 1961, what he called "randomly transitional phenomena". Yet his advisor did not agree with his conclusions at the time, and did not allow him to report his findings until 1970.[78][79]

strange attractor
, a main concept of chaos theory.

Lorenz attractors
, showed that even detailed atmospheric modeling cannot, in general, make precise long-term weather predictions.

In 1963,

Koch curve or snowflake, which is infinitely long yet encloses a finite space and has a fractal dimension of circa 1.2619). In 1982, Mandelbrot published The Fractal Geometry of Nature, which became a classic of chaos theory.[86]

In December 1977, the New York Academy of Sciences organized the first symposium on chaos, attended by David Ruelle, Robert May, James A. Yorke (coiner of the term "chaos" as used in mathematics), Robert Shaw, and the meteorologist Edward Lorenz. The following year Pierre Coullet and Charles Tresser published "Itérations d'endomorphismes et groupe de renormalisation", and Mitchell Feigenbaum's article "Quantitative Universality for a Class of Nonlinear Transformations" finally appeared in a journal, after 3 years of referee rejections.[87][88] Thus Feigenbaum (1975) and Coullet & Tresser (1978) discovered the universality in chaos, permitting the application of chaos theory to many different phenomena.

In 1979,

Mitchell J. Feigenbaum for their inspiring achievements.[89]

In 1986, the New York Academy of Sciences co-organized with the National Institute of Mental Health and the Office of Naval Research the first important conference on chaos in biology and medicine. There, Bernardo Huberman presented a mathematical model of the eye tracking dysfunction among people with schizophrenia.[90] This led to a renewal of physiology in the 1980s through the application of chaos theory, for example, in the study of pathological cardiac cycles.

In 1987,

Chao Tang and Kurt Wiesenfeld published a paper in Physical Review Letters[91] describing for the first time self-organized criticality (SOC), considered one of the mechanisms by which complexity
arises in nature.

Alongside largely lab-based approaches such as the

biological evolution (where SOC has been invoked, for example, as the dynamical mechanism behind the theory of "punctuated equilibria" put forward by Niles Eldredge and Stephen Jay Gould). Given the implications of a scale-free distribution of event sizes, some researchers have suggested that another phenomenon that should be considered an example of SOC is the occurrence of wars
. These investigations of SOC have included both attempts at modelling (either developing new models or adapting existing ones to the specifics of a given natural system), and extensive data analysis to determine the existence and/or characteristics of natural scaling laws.

Also in 1987 James Gleick published Chaos: Making a New Science, which became a best-seller and introduced the general principles of chaos theory as well as its history to the broad public.[93] Initially the domain of a few, isolated individuals, chaos theory progressively emerged as a transdisciplinary and institutional discipline, mainly under the name of nonlinear systems analysis. Alluding to Thomas Kuhn's concept of a paradigm shift exposed in The Structure of Scientific Revolutions (1962), many "chaologists" (as some described themselves) claimed that this new theory was an example of such a shift, a thesis upheld by Gleick.

The availability of cheaper, more powerful computers broadens the applicability of chaos theory. Currently, chaos theory remains an active area of research,

etc.

Lorenz's pioneering contributions to chaotic modeling

Throughout his career, Professor Lorenz authored a total of 61 research papers, out of which 58 were solely authored by him.[97] Commencing with the 1960 conference in Japan, Lorenz embarked on a journey of developing diverse models aimed at uncovering the SDIC and chaotic features. A recent review of Lorenz's model[98][99] progression spanning from 1960 to 2008 revealed his adeptness at employing varied physical systems to illustrate chaotic phenomena. These systems encompassed Quasi-geostrophic systems, the Conservative Vorticity Equation, the Rayleigh-Bénard Convection Equations, and the Shallow Water Equations. Moreover, Lorenz can be credited with the early application of the logistic map to explore chaotic solutions, a milestone he achieved ahead of his colleagues (e.g. Lorenz 1964[100]).

In 1972, Lorenz coined the term "butterfly effect" as a metaphor to discuss whether a small perturbation could eventually create a tornado with a three-dimensional, organized, and coherent structure. While connected to the original butterfly effect based on sensitive dependence on initial conditions, its metaphorical variant carries distinct nuances. To commemorate this milestone, a reprint book containing invited papers that deepen our understanding of both butterfly effects was officially published to celebrate the 50th anniversary of the metaphorical butterfly effect.[101]

A popular but inaccurate analogy for chaos

The sensitive dependence on initial conditions (i.e., butterfly effect) has been illustrated using the following folklore:[93]

For want of a nail, the shoe was lost.

For want of a shoe, the horse was lost.

For want of a horse, the rider was lost.

For want of a rider, the battle was lost.

For want of a battle, the kingdom was lost.

And all for the want of a horseshoe nail.

Based on the above, many people mistakenly believe that the impact of a tiny initial perturbation monotonically increases with time and that any tiny perturbation can eventually produce a large impact on numerical integrations. However, in 2008, Lorenz stated that he did not feel that this verse described true chaos but that it better illustrated the simpler phenomenon of instability and that the verse implicitly suggests that subsequent small events will not reverse the outcome.[102] Based on the analysis, the verse only indicates divergence, not boundedness.[6] Boundedness is important for the finite size of a butterfly pattern.[6][102][103] In a recent study,[104] the characteristic of the aforementioned verse was recently denoted as "finite-time sensitive dependence".

Applications

A conus textile shell, similar in appearance to Rule 30, a cellular automaton with chaotic behaviour.[105]

Although chaos theory was born from observing weather patterns, it has become applicable to a variety of other situations. Some areas benefiting from chaos theory today are geology, mathematics, biology, computer science, economics,[106][107][108] engineering,[109][110] finance,[111][112][113][114][115] meteorology, philosophy, anthropology,[15] physics,[116][117][118] politics,[119][120] population dynamics,[121] and robotics. A few categories are listed below with examples, but this is by no means a comprehensive list as new applications are appearing.

Cryptography

Chaos theory has been used for many years in

diffusion and confusion, which is modeled well by chaos theory.[124] Another type of computing, DNA computing, when paired with chaos theory, offers a way to encrypt images and other information.[125] Many of the DNA-Chaos cryptographic algorithms are proven to be either not secure, or the technique applied is suggested to be not efficient.[126][127][128]

Robotics

Robotics is another area that has recently benefited from chaos theory. Instead of robots acting in a trial-and-error type of refinement to interact with their environment, chaos theory has been used to build a predictive model.[129] Chaotic dynamics have been exhibited by passive walking biped robots.[130]

Biology

For over a hundred years, biologists have been keeping track of populations of different species with population models. Most models are continuous, but recently scientists have been able to implement chaotic models in certain populations.[131] For example, a study on models of Canadian lynx showed there was chaotic behavior in the population growth.[132] Chaos can also be found in ecological systems, such as hydrology. While a chaotic model for hydrology has its shortcomings, there is still much to learn from looking at the data through the lens of chaos theory.[133] Another biological application is found in cardiotocography. Fetal surveillance is a delicate balance of obtaining accurate information while being as noninvasive as possible. Better models of warning signs of fetal hypoxia can be obtained through chaotic modeling.[134]

As Perry points out,

chaotic- oscillations in pathogen population.[137]
: 169 

Economics

It is possible that economic models can also be improved through an application of chaos theory, but predicting the health of an economic system and what factors influence it most is an extremely complex task.[138] Economic and financial systems are fundamentally different from those in the classical natural sciences since the former are inherently stochastic in nature, as they result from the interactions of people, and thus pure deterministic models are unlikely to provide accurate representations of the data. The empirical literature that tests for chaos in economics and finance presents very mixed results, in part due to confusion between specific tests for chaos and more general tests for non-linear relationships.[139]

Chaos could be found in economics by the means of recurrence quantification analysis. In fact, Orlando et al.[140] by the means of the so-called recurrence quantification correlation index were able detect hidden changes in time series. Then, the same technique was employed to detect transitions from laminar (regular) to turbulent (chaotic) phases as well as differences between macroeconomic variables and highlight hidden features of economic dynamics.[141] Finally, chaos theory could help in modeling how an economy operates as well as in embedding shocks due to external events such as COVID-19.[142]

AI-Extended Modeling Framework

In AI-driven large language models, responses can exhibit sensitivities to factors like alterations in formatting and variations in prompts. These sensitivities are akin to butterfly effects.[143] Although classifying AI-powered large language models as classical deterministic chaotic systems poses challenges, chaos-inspired approaches and techniques (such as ensemble modeling) may be employed to extract reliable information from these expansive language models (see also "Butterfly Effect in Popular Culture").

Other areas

In chemistry, predicting gas solubility is essential to manufacturing

Josephson junctions benefitted greatly from chaos theory.[146] Closer to home, coal mines have always been dangerous places where frequent natural gas leaks cause many deaths. Until recently, there was no reliable way to predict when they would occur. But these gas leaks have chaotic tendencies that, when properly modeled, can be predicted fairly accurately.[147]

Chaos theory can be applied outside of the natural sciences, but historically nearly all such studies have suffered from lack of reproducibility; poor external validity; and/or inattention to cross-validation, resulting in poor predictive accuracy (if out-of-sample prediction has even been attempted). Glass[148] and Mandell and Selz[149] have found that no EEG study has as yet indicated the presence of strange attractors or other signs of chaotic behavior.

Researchers have continued to apply chaos theory to psychology. For example, in modeling group behavior in which heterogeneous members may behave as if sharing to different degrees what in Wilfred Bion's theory is a basic assumption, researchers have found that the group dynamic is the result of the individual dynamics of the members: each individual reproduces the group dynamics in a different scale, and the chaotic behavior of the group is reflected in each member.[150]

Redington and Reidbord (1992) attempted to demonstrate that the human heart could display chaotic traits. They monitored the changes in between-heartbeat intervals for a single psychotherapy patient as she moved through periods of varying emotional intensity during a therapy session. Results were admittedly inconclusive. Not only were there ambiguities in the various plots the authors produced to purportedly show evidence of chaotic dynamics (spectral analysis, phase trajectory, and autocorrelation plots), but also when they attempted to compute a Lyapunov exponent as more definitive confirmation of chaotic behavior, the authors found they could not reliably do so.[151]

In their 1995 paper, Metcalf and Allen[152] maintained that they uncovered in animal behavior a pattern of period doubling leading to chaos. The authors examined a well-known response called schedule-induced polydipsia, by which an animal deprived of food for certain lengths of time will drink unusual amounts of water when the food is at last presented. The control parameter (r) operating here was the length of the interval between feedings, once resumed. The authors were careful to test a large number of animals and to include many replications, and they designed their experiment so as to rule out the likelihood that changes in response patterns were caused by different starting places for r.

Time series and first delay plots provide the best support for the claims made, showing a fairly clear march from periodicity to irregularity as the feeding times were increased. The various phase trajectory plots and spectral analyses, on the other hand, do not match up well enough with the other graphs or with the overall theory to lead inexorably to a chaotic diagnosis. For example, the phase trajectories do not show a definite progression towards greater and greater complexity (and away from periodicity); the process seems quite muddied. Also, where Metcalf and Allen saw periods of two and six in their spectral plots, there is room for alternative interpretations. All of this ambiguity necessitate some serpentine, post-hoc explanation to show that results fit a chaotic model.

By adapting a model of career counseling to include a chaotic interpretation of the relationship between employees and the job market, Amundson and Bright found that better suggestions can be made to people struggling with career decisions.[153] Modern organizations are increasingly seen as open complex adaptive systems with fundamental natural nonlinear structures, subject to internal and external forces that may contribute chaos. For instance, team building and group development is increasingly being researched as an inherently unpredictable system, as the uncertainty of different individuals meeting for the first time makes the trajectory of the team unknowable.[154]

Some say the chaos metaphor—used in verbal theories—grounded on mathematical models and psychological aspects of human behavior provides helpful insights to describing the complexity of small work groups, that go beyond the metaphor itself.[155]

The red cars and blue cars take turns to move; the red ones only move upwards, and the blue ones move rightwards. Every time, all the cars of the same colour try to move one step if there is no car in front of it. Here, the model has self-organized in a somewhat geometric pattern where there are some traffic jams and some areas where cars can move at top speed.
The red cars and blue cars take turns to move; the red ones only move upwards, and the blue ones move rightwards. Every time, all the cars of the same colour try to move one step if there is no car in front of it. Here, the model has self-organized in a somewhat geometric pattern where there are some traffic jams and some areas where cars can move at top speed.

Traffic forecasting may benefit from applications of chaos theory. Better predictions of when a congestion will occur would allow measures to be taken to disperse it before it would have occurred. Combining chaos theory principles with a few other methods has led to a more accurate short-term prediction model (see the plot of the

BML traffic model at right).[156]

Chaos theory has been applied to environmental

hydrological data), such as rainfall and streamflow.[157] These studies have yielded controversial results, because the methods for detecting a chaotic signature are often relatively subjective. Early studies tended to "succeed" in finding chaos, whereas subsequent studies and meta-analyses called those studies into question and provided explanations for why these datasets are not likely to have low-dimension chaotic dynamics.[158]

In art (predominately art theory) a possible postpostmodern era has been outlined with emphasis on multiple narratives and the notion that every fictional angle is a possibility. In part this is therefor of a bisociate (trissociative) discourse and can be explained within emphasis on an institutional interchange of subjectivistic agents.[159]

See also

Examples of chaotic systems

Other related topics

People

References

  1. ^ "chaos theory | Definition & Facts". Encyclopedia Britannica. Retrieved 2019-11-24.
  2. ^ a b c "What is Chaos Theory? – Fractal Foundation". Retrieved 2019-11-24.
  3. ^ Weisstein, Eric W. "Chaos". mathworld.wolfram.com. Retrieved 2019-11-24.
  4. ^ Boeing, Geoff (26 March 2015). "Chaos Theory and the Logistic Map". Retrieved 2020-05-17.
  5. ^ a b c Lorenz, Edward (1993). The Essence of Chaos. University of Washington Press. pp. 181–206.
  6. ^
    ISSN 2673-8392. Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
    .
  7. .
  8. ^ a b c d Bishop, Robert (2017), "Chaos", in Zalta, Edward N. (ed.), The Stanford Encyclopedia of Philosophy (Spring 2017 ed.), Metaphysics Research Lab, Stanford University, retrieved 2019-11-24
  9. ^ Kellert 1993, p. 56
  10. ^ Kellert 1993, p. 62
  11. ^
    S2CID 354849
    .
  12. ^ Danforth, Christopher M. (April 2013). "Chaos in an Atmosphere Hanging on a Wall". Mathematics of Planet Earth 2013. Retrieved 12 June 2018.
  13. ^ .
  14. .
  15. ^ a b Mosko M.S., Damon F.H. (Eds.) (2005). On the order of chaos. Social anthropology and the science of chaos. Oxford: Berghahn Books.
  16. ^ a b Piotrowski, Chris. "Covid-19 Pandemic and Chaos Theory: Applications based on a Bibliometric Analysis". researchgate.net. Retrieved 2020-05-13.
  17. ^ .
  18. .
  19. ^ Sync: The Emerging Science of Spontaneous Order, Steven Strogatz, Hyperion, New York, 2003, pages 189–190.
  20. ^ Definition of chaos at Wiktionary;
  21. ^ "Definition of chaos | Dictionary.com". www.dictionary.com. Retrieved 2019-11-24.
  22. .
  23. .
  24. .
  25. .
  26. .
  27. .
  28. ^ "Edward Lorenz, father of chaos theory and butterfly effect, dies at 90". MIT News. 16 April 2008. Retrieved 2019-11-24.
  29. ISSN 2073-4433
    .
  30. ^ Watts, Robert G. (2007). Global Warming and the Future of the Earth. Morgan & Claypool. p. 17.
  31. ^ Weisstein, Eric W. "Lyapunov Characteristic Exponent". mathworld.wolfram.com. Retrieved 2019-11-24.
  32. ^ a b Devaney 2003
  33. ^ Robinson 1995
  34. ^ Alligood, Sauer & Yorke 1997
  35. JSTOR 2318254. Archived from the original
    (PDF) on 2009-12-29.
  36. .
  37. ^ .
  38. ^ .
  39. ^ .
  40. .
  41. .
  42. .
  43. .
  44. .
  45. .
  46. .
  47. .
  48. .
  49. .
  50. .
  51. .
  52. .
  53. .
  54. ^ K. E. Chlouverakis and J. C. Sprott, Chaos Solitons & Fractals 28, 739–746 (2005), Chaotic Hyperjerk Systems, http://sprott.physics.wisc.edu/pubs/paper297.htm
  55. ^ "A New Chaotic Jerk Circuit", J. C. Sprott, IEEE Transactions on Circuits and Systems,2011.
  56. ^ "Simple Autonomous Chaotic Circuits", J. C. Sprott, IEEE Transactions on Circuits and Systems—II: Express Briefs, 2010.
  57. ^ "Secure Image Encryption Based On a Chua Chaotic Noise Generator", A. S. Andreatos*, and A. P. Leros, Journal of Engineering Science and Technology Review, 2013.
  58. ^ Steven Strogatz, Sync: The Emerging Science of Spontaneous Order, Hyperion, 2003.
  59. ISSN 0960-0779
    .
  60. ^ Hunt, Brian R.; Yorke, James A. (1993). "Maxwell on Chaos" (PDF). Nonlinear Science Today. 3 (1).
  61. ^ Everitt, Francis (2006-12-01). "James Clerk Maxwell: a force for physics". Physics World. Retrieved 2023-11-03.
  62. S2CID 225246631
    .
  63. .
  64. .
  65. ^ Diacu, Florin; Holmes, Philip (1996). Celestial Encounters: The Origins of Chaos and Stability. Princeton University Press.
  66. ^ Hadamard, Jacques (1898). "Les surfaces à courbures opposées et leurs lignes géodesiques". Journal de Mathématiques Pures et Appliquées. 4: 27–73.
  67. ^ George D. Birkhoff, Dynamical Systems, vol. 9 of the American Mathematical Society Colloquium Publications (Providence, Rhode Island: American Mathematical Society, 1927)
  68. S2CID 123612939
    .
  69. .
  70. .
  71. ^ Chirikov, Boris. "РЕЗОНАНСНЫЕ ПРОЦЕССЫ В МАГНИТНЫХ ЛОВУШКАХ" (PDF). Атомная энергия. 6.
  72. S2CID 59483478
    .
  73. ^ Jean, Bellissard; Dima, Shepelyansky (27 February 1998). "Boris Chirikov, a pioneer in classical and quantum chaos" (PDF). Annales Henri Poincaré. 68 (4): 379.
  74. S2CID 119107150
    .
  75. .
  76. ^ Abraham & Ueda 2000, See Chapters 3 and 4
  77. ^ Sprott 2003, p. 89
  78. ^ Sokol, Joshua (May 20, 2019). "The Hidden Heroines of Chaos". Quanta Magazine. Retrieved 2022-11-09.
  79. .
  80. .
  81. ^ Mandelbrot, B. (1977). The Fractal Geometry of Nature. New York: Freeman. p. 248.
  82. .
  83. S2CID 15662830. Archived from the original
    on 19 October 2021. Retrieved 31 January 2022.
  84. .
  85. .
  86. ^ Coullet, Pierre, and Charles Tresser. "Iterations d'endomorphismes et groupe de renormalisation." Le Journal de Physique Colloques 39.C5 (1978): C5-25
  87. ^ "The Wolf Prize in Physics in 1986". Archived from the original on 2012-02-05. Retrieved 2008-01-17.
  88. S2CID 42733652
    .
  89. S2CID 7674321. However, the conclusions of this article have been subject to dispute. "?". Archived from the original
    on 2007-12-14.. See especially: Laurson, Lasse; Alava, Mikko J.; Zapperi, Stefano (15 September 2005). "Letter: Power spectra of self-organized critical sand piles". Journal of Statistical Mechanics: Theory and Experiment. 0511. L001.
  90. ^ Omori, F. (1894). "On the aftershocks of earthquakes". Journal of the College of Science, Imperial University of Tokyo. 7: 111–200.
  91. ^ .
  92. .
  93. .
  94. .
  95. ^ Chen, G.-R. (2020-01-01). "Butterfly Effect and Chaos" (PDF). Retrieved 1 July 2023.
  96. .
  97. .
  98. .
  99. .
  100. ^ a b Lorenz, E. N. (December 2008). "The butterfly effect. In Premio Felice Pietro Chisesi E Caterina Tomassoni Award Lecture; University of Rome: Rome, Italy" (PDF). Archived from the original (PDF) on June 10, 2023. Retrieved January 29, 2023.
  101. ^ Shen, Bo-Wen. "A Popular but Inaccurate Analogy for Chaos and Butterfly Effect". YouTube. Retrieved 2023-02-21.
  102. ISSN 2073-4433
    .
  103. ^ Stephen Coombes (February 2009). "The Geometry and Pigmentation of Seashells" (PDF). www.maths.nottingham.ac.uk. University of Nottingham. Archived (PDF) from the original on 2013-11-05. Retrieved 2013-04-10.
  104. .
  105. .
  106. ^ Kyrtsou, C.; Vorlow, C. (2005). "Complex dynamics in macroeconomics: A novel approach". In Diebolt, C.; Kyrtsou, C. (eds.). New Trends in Macroeconomics. Springer Verlag.
  107. ISSN 1367-2630
    .
  108. ^ "Applying Chaos Theory to Embedded Applications". Archived from the original on 9 August 2011.
  109. . 138547.
  110. .
  111. .
  112. .
  113. .
  114. .
  115. .
  116. .
  117. .
  118. .
  119. .
  120. ^ .
  121. .
  122. .
  123. .
  124. .
  125. .
  126. .
  127. (PDF) on 2017-08-12. Retrieved 2017-10-25.
  128. .
  129. .
  130. .
  131. .
  132. .
  133. ^ .
  134. ^ a b c Thompson, John; Burdon, Jeremy (1992). "Gene-for-gene coevolution between plants and parasites". Review Article.
    S2CID 4346920
    .
  135. ^ Jones, Gareth (1998). Jones, D. Gareth (ed.). The Epidemiology of Plant Diseases (1 ed.).
    S2CID 1793087
    .
  136. .
  137. (PDF) from the original on 2017-08-09.
  138. .
  139. .
  140. .
  141. ].
  142. .
  143. .
  144. ^ Steven Strogatz, Sync: The Emerging Science of Spontaneous Order, Hyperion, 2003
  145. .
  146. ^ Glass, L (1997). "Dynamical disease: The impact of nonlinear dynamics and chaos on cardiology and medicine". In Grebogi, C; Yorke, J. A. (eds.). The impact of chaos on science and society. United Nations University Press.
  147. ^ Mandell, A. J.; Selz, K. A. (1997). "Is the EEG a strange attractor?". In Grebogi, C; Yorke, J. A. (eds.). The impact of chaos on science and society. United Nations University Press.
  148. PMID 23517610
    .
  149. .
  150. ^ Metcalf, B. R.; Allen, J. D. (1995). "In search of chaos in schedule-induced polydipsia". In Abraham, F. D.; Gilgen, A. R. (eds.). Chaos theory in psychology. Greenwood Press.
  151. .
  152. .
  153. .
  154. .
  155. ^ "Dr. Gregory B. Pasternack – Watershed Hydrology, Geomorphology, and Ecohydraulics :: Chaos in Hydrology". pasternack.ucdavis.edu. Retrieved 2017-06-12.
  156. .
  157. ^ Thomas III, Fred Charles (July 1990). "Design of a Sinusoidally-Force Pendulum for the Demonstration of Chaotic Phenomena". Bucknell University MS Engineering Thesis – via Bucknell University.

Further reading

Articles

Textbooks

Semitechnical and popular works

External links

Copyright note