Chaos and Order
eBook - ePub

Chaos and Order

Complex Dynamics in Literature and Science

  1. English
  2. ePUB (mobile friendly)
  3. Available on iOS & Android
eBook - ePub

Chaos and Order

Complex Dynamics in Literature and Science

Book details
Book preview
Table of contents
Citations

About This Book

The scientific discovery that chaotic systems embody deep structures of order is one of such wide-ranging implications that it has attracted attention across a spectrum of disciplines, including the humanities. In this volume, fourteen theorists explore the significance for literary and cultural studies of the new paradigm of chaotics, forging connections between contemporary literature and the science of chaos. They examine how changing ideas of order and disorder enable new readings of scientific and literary texts, from Newton's Principia to Ruskin's autobiography, from Victorian serial fiction to Borges's short stories.N. Katherine Hayles traces shifts in meaning that chaos has undergone within the Western tradition, suggesting that the science of chaos articulates categories that cannot be assimilated into the traditional dichotomy of order and disorder. She and her contributors take the relation between order and disorder as a theme and develop its implications for understanding texts, metaphors, metafiction, audience response, and the process of interpretation itself. Their innovative and diverse work opens the interdisciplinary field of chaotics to literary inquiry.

Frequently asked questions

Simply head over to the account section in settings and click on “Cancel Subscription” - it’s as simple as that. After you cancel, your membership will stay active for the remainder of the time you’ve paid for. Learn more here.
At the moment all of our mobile-responsive ePub books are available to download via the app. Most of our PDFs are also available to download and we're working on making the final remaining ones downloadable now. Learn more here.
Both plans give you full access to the library and all of Perlego’s features. The only differences are the price and subscription period: With the annual plan you’ll save around 30% compared to 12 months on the monthly plan.
We are an online textbook subscription service, where you can get access to an entire online library for less than the price of a single book per month. With over 1 million books across 1000+ topics, we’ve got you covered! Learn more here.
Look out for the read-aloud symbol on your next book to see if you can listen to it. The read-aloud tool reads text aloud for you, highlighting the text as it is being read. You can pause it, speed it up and slow it down. Learn more here.
Yes, you can access Chaos and Order by N. Katherine Hayles, N. Katherine Hayles in PDF and/or ePUB format, as well as other popular books in Biological Sciences & Science General. We have over one million books available in our catalogue for you to explore.

Information

1
Introduction: Complex Dynamics in Literature and Science
N. Katherine Hayles
The law of chaos is the law of ideas, of improvisations and seasons of belief.
Wallace Stevens, “Extracts from Addresses to the Academy of Fine Ideas”
I
This volume focuses on a question that is attracting increasing attention within the sciences and humanities: what is the relation between order and disorder? Traditionally, of course, they have been regarded as opposites. Order was that which could be classified, analyzed, encompassed within rational discourse; disorder was allied with chaos and by definition could not be expressed except through statistical generalizations. The last twenty years have seen a radical reevaluation of this view. In both contemporary literature and science, chaos has been conceptualized as extremely complex information rather than an absence of order. As a result, textuality is conceived in new ways within critical theory and literature, and new kinds of phenomena are coming to the fore within an emerging field known as the science of chaos.
A science of chaos may seem to be a contradiction in terms. In the scientific sense, however, chaos means something different than it does in common usage. At the center of chaos theory is the discovery that hidden within the unpredictability of chaotic systems are deep structures of order. “Chaos,” in this usage, denotes not true randomness but the orderly disorder characteristic of these systems. The science of chaos seeks to understand behavior so complex that it defeats the usual methods of formalizing a system through mathematics. Hence the science of chaos has also been called the science of complexity—or more precisely the sciences of complexity, for fields as diverse as meteorology, irreversible thermodynamics, epidemiology, and nonlinear dynamics are included within the rubric. The kinds of systems to which chaos models have been successfully applied range from dripping faucets to measles epidemics, schizophrenic eye movements to fluctuations in fish populations.1 Recently there has been considerable interest in using chaos theory to understand the stock market.2 Although it is too soon to say where the discoveries associated with complex systems will end, it is already apparent that chaos theory is part of a paradigm shift of remarkable scope and significance.
Among the controversial issues within the science of chaos is the word “chaos” itself. No sooner did the term become widely associated with nonlinear dynamics than practitioners in the field began to shy away from it, regarding it as an imprecise, even sensationalized, word that was unnecessarily confusing. As the term gained notoriety, chiefly through James Gleick’s popular book Chaos: Making a New Science, it lost credibility within the scientific community. To many, the word has now become so thoroughly deprofessionalized that its use is regarded as a signal that one is in the presence of a dilettante rather than an expert. Nevertheless, it will be retained in this volume precisely because of the ambiguous meanings that inhere within it. Marked by scientific denotations as well as historical and mythic interpretations, it serves as a crossroads, a juncture where various strata and trends within the culture come together.
The cultural attitudes that mark the word “chaos” can be read in its etymology. The word derives from a Greek verb-stem, KHA, meaning “to yawn, to gape”; from this comes the meaning given by the Oxford English Dictionary, “a gaping void, yawning gulf, chasm, or abyss.” Creation myths in the West, from the Babylonian epic Enuma Elish to Milton’s Paradise Lost, depict chaos as a negative state, a disordered void which must be conquered for creation to occur. Eugene Eoyang (“Chaos Misread”), writing about Enuma Elish, recalls the dramatic moment in the epic when the demoness Tiamat is slain by the young warrior Marduk. He explains that Tiamat stands for the entropic tendency toward dissolution, the force of decay that would return everything to darkness. Her death signals the beginning of a new era. Marduk, exultant over Tiamat’s carcass, exclaims that he will create a new “savage,” and that “‘man’ shall be his name” (quoted in Eoyang 272). The emphasis on man as savage implies that chaos is the aboriginal foe which must be vanquished before the civilizing process can begin. In this semiotic, chaos is opposed to civilized values as well as to the initiating act of creation.
Eoyang contrasts these assumptions with those encoded into the ancient Taoist creation myth recounted in the seventh chapter of the Zhuangzi. Shu (Brief) and Hu (Sudden) go to visit Hun-dun (Chaos), who graciously offers them his hospitality. Observing that Hun-dun lacks the seven openings through which men see, hear, eat, and breathe, Shu and Hu determine to create them. Each day they bore a new hole. On the seventh day, Hun-dun dies. Here the destruction of chaos, far from marking the beginning of civilization, bespeaks a provinciality unable to accept an other different than the self. In contrast to the triumphal climax of the Western epic, the Taoist story ends with an ironic twist.
The etymology of “Hun-dun” provides a similar contrast. Eoyang explains that the word belongs to a category of rhyming compounds whose sounds enact their meanings. The onomatopoetic quality that Hun-dun suggests is whirling water, flowing turbulence, swirling action. According to Eoyang, the word is not so much a nominative as an attributive adjective, a quality rather than an object (275). He quotes a scholar from the Jin Dynasty, Du Yü, who wrote that Hun-dun was “something whose appearance cannot be seen through” (quoted in Eoyang 275). Thus in the myth Hun-dun has no openings, and when his visitors insist on providing him with some, he dies. There is the delicate suggestion that Hun-dun stands outside the world of reified concepts Taoist irony tries to puncture. Chaos remains the necessary other, the opaque turbulence that challenges and complements the transparency of order.
That chaos has been negatively valued in the Western tradition may be partly due, Eoyang suggests (“Heuristics”), to the predominance of binary logic in the West. If order is good, chaos is bad because it is conceptualized as the opposite of order. By contrast, in the four-valued logic characteristic of Taoist thought, not-order is also a possibility, distinct from and valued differently than anti-order. The science of chaos draws Western assumptions about chaos into question by revealing possibilities that were suppressed when chaos was considered merely as order’s opposite. It marks the validation within the Western tradition of a view of chaos that constructs it as not-order. In chaos theory chaos may either lead to order, as it does with self-organizing systems, or in yin/yang fashion it may have deep structures of order encoded within it. In either case, its relation to order is more complex than traditional Western oppositions have allowed.
This cultural history helps to explain why the science of chaos should seem revolutionary within the Western tradition. Many of the methods of nonlinear dynamics are not new, going back to the nineteenth century; and many of its central insights had been anticipated in the nineteenth century and before. Yet calling the object of inquiry “chaos” caused something to click that had not before. In 1975 when James Yorke, with characteristic flair, chose the flashy title “Period Three Implies Chaos” for a seminal paper on deterministic disorder (Yorke and Li), he tapped into a network of presuppositions that may soon put the emerging science of chaos on a par with evolution, relativity, and quantum mechanics in its impact on the culture. The furor over chaos has made many practitioners in nonlinear dynamics feel it would be best to avoid the word altogether. They see themselves as solving practical and technical problems, not rewriting cultural history.
Yet cultural traditions are not so easily escaped. They are encoded not merely into words but also into practices, institutions, and material conditions. If on the one hand this means that chaos theory can scarcely avoid having implications for the culture beyond its technical achievements, on the other it implies that existing practices exert an inertial pull on new ideas. The embedding of chaos theory into existing disciplinary contexts makes it difficult to determine how revolutionary it is. The claim is sometimes heard, for example, that the science of chaos challenges traditional ideas of how science is done. But the science of chaos is not opposed to normal science. It is normal science. Its criteria for evaluating evidence, reproducing results, credentialling investigators, and so on, differs not at all from those of other physical sciences. Thus its insights have evolved within contexts that partially reinscribe the very assumptions these insights draw into question. The result is no simple revolution in which a new view replaces an old. Rather, change occurs through negotiations at multiple sites among those who generate data, interpret them, theorize about them, and extrapolate beyond them to broader cultural and philosophical significances.
This volume aims to investigate these negotiations. It treats chaos both as a subject of scientific inquiry and a crossroads where various paths within the culture converge. It explores how insights that work against the grain of Western culture mingle with other cultural currents, changing and being changed by the resistances they encounter. It articulates chaos theory together with developments in the human sciences and postmodern culture. Such an approach implies, of course, that the science of chaos is part of the culture, and that scientists, like everyone else, are affected by the culture in which they are immersed. Yet so strong is the ideology of scientific objectivity that practitioners and laymen alike often speak as if scientists were hermetically sealed within the laboratory, isolated from and immune to the thousands of experiences that constitute the fabric of everyday life.
This hermetic view of scientific inquiry has recently been challenged on a number of fronts. Among the challengers are feminist critiques that explore gender issues, sociological studies that investigate how science is socially constructed, and ethnographic analyses of scientific communities as tribes with their own vocabularies, rituals, and social practices.3 These studies show that even if we leave aside such obvious and important influences on scientific enterprises as sources of money, governmental regulations, commercial possibilities, and popular opinion, there remains a host of other cultural factors less visible but no less constitutive of scientific inquiry.
The one most relevant here is language. In general, scientific discourse adopts as its ideal univocality—one word, one meaning. Closely related to this goal is the belief that a language exists, or can be forged, that is purely instrumental. Clearly and unambiguously, it will communicate to the world what the speaker or writer intends to say. Roland Barthes (Rustle) has ironically called this the belief that science can own a slave language, docile and obedient to its demands.4 Anyone who has seriously studied how language works is aware, however, that it shapes even as it articulates thought. There is now an impressive body of work exploring how metaphors, narrative patterns, rhetorical structures, syntax, and semantic fields affect scientific discourse and thought. Prominent examples include Gillian Beer’s Darwin’s Plots, Donald McCloskey’s The Rhetoric of Economics, Michael Arbib and Mary Hesse’s The Construction of Reality, Charles Bazerman’s Shaping Written Knowledge, and Bruno Latour’s Science in Action. These studies, along with many others, demonstrate that language is not a passive instrument but an active engagement with a vital medium that has its own currents, resistances, subversions, enablings, pathways, blockages. As soon as discovery is communicated through language, it is also constituted by language.
That language is interactive rather than inert implies that chaos theory is influenced by the culture within which it arose. In my view, it is a mistake to assume that the science of chaos has initiated the attitudes that have made it an object of popular fascination. Rather, it is one site within the culture where the premises characteristic of postmodernism are inscribed. The postmodern context catalyzed the formation of the new science by providing a cultural and technological milieu in which the component parts came together and mutually reinforced each other until they were no longer isolated events but an emergent awareness of the constructive roles that disorder, nonlinearity, and noise play in complex systems. The science of chaos is new not in the sense of having no antecedents in the scientific tradition, but of having only recently coalesced sufficiently to articulate a vision of the world. It is no accident that this vision has deep affinities with other articulations that have emerged from the postmodern context.
What are these affinities? The material conditions under which the science evolved are relevant as well as linguistic and cultural factors. The microcomputer, for example, has been extremely important in the development of chaos theory, for it allows mathematics to be practiced as an experimental science. It has also affected how people have imaged themselves and their relation to the world. The two come together when someone sits down in front of a computer to model a dynamical nonlinear system. Because the computer permits interaction, the practitioner need not proceed through the traditional mathematical method of theorem-proof. Instead she can set up a recursive program that begins when she feeds initial values for the equations into the computer. Then she watches as the screen display generated by the recursion evolves into constantly changing, often unexpected patterns. As the display continues, she adjusts the parameters to achieve different effects. With her own responses in a feedback loop with the computer, she develops an intuitive feeling for how the display and parameters interact. She notices that small changes in initial values can lead to large changes in the display. She also sees that, although the displays are complex, there are underlying symmetries that impart a pleasing, sometimes a startlingly beautiful quality to them. And she is subliminally aware that her interaction with the display could be thought of as one complex system (the behavior described by a set of nonlinear differential equations) interfaced with another (the human neural system) through the medium of the computer.
What presuppositions does this situation embody and reinforce? Perhaps the most obvious is that the connective tissue holding the system together is the flow of information circulating through it. William Gibson, author of the cyberpunk novel Neuromancer, remarked in an interview that a teenager playing an arcade video game illustrates how an informational feedback loop connects human and machine.5 Photons leave the screen, enter the teenager’s eyes, and trigger neural responses that coordinate with hand movements, which in turn cause the electronic circuitry of the machine to produce more photons. When such experiences are everyday events, a context is created that makes information flow seem as real as the matter and energy that carries it—or more real.
This context affects how theoretically abstract ideas such as chaos are invested with meaning. An important turning point in the science of chaos occurred when complex systems were conceptualized as systems rich in information rather than poor in order. One could construct a history that would treat this development as if it emerged solely from the internal logic of the scientific tradition, for example by analyzing competing definitions of entropy and examining their relation to information theory.6 But the context that made disorder appear as complex information is not confined to scientific inquiry alone. It is part of a cultural milieu that included World War II, which among other things was an object lesson in the importance of information; consolidation of power by multinational corporations and the accompanying sense that the world was growing at once more chaotic and more totalized; increasing economic interdependencies between nations, which brought home to nearly everyone that small causes could lead to large-scale effects; and rapid expansion of information technologies.7 All of these factors, and more, contributed to the cultural matrix out of which the science of chaos grew.
It should not be surprising, then, to find other sites within the culture that also embody the presuppositions informing chaos theory. To distinguish between the science of chaos narrowly defined and this broader cultural phenomenon, I will appropriate a term suggested to me by Ihab Hassan and call the latter chaotics.8 The term signifies certain attitudes toward chaos that are manifest at diverse sites within the culture, among them poststructuralism and the science of chaos. The question of how such isomorphisms arise is not easily answered. Let me say at the outset, however, that I do not assume they are the result of direct influence between one site and another. In particular, I am not arguing that the science of chaos is the originary site from which chaotics emanates into the culture. Rather, both the literary and scientific manifestations of chaotics are involved in feedback loops with the culture. They help to create the context that energizes the questions they ask; at the same time, they also ask questions energized by the context.
The impact of chaotics derives less from specific theories than from the general awareness it fosters of nonlinear processes and forms. To some extent, we see what we are taught to see. How does the world envisioned by chaotics differs from that of Newtonian mechanics? The Newtonian paradigm emphasizes predictability. Such a mindset is exemplified by Laplace’s famous boast that, given the initial conditions and an intelligence large e...

Table of contents

  1. Cover
  2. Copyright
  3. Title Page
  4. Series Page
  5. Contents
  6. Acknowledgments
  7. 1. Introduction: Complex Dynamics in Literature and Science
  8. I. Chaos: More than Metaphor
  9. II. Order: Revisioning Form
  10. III. Chaos and Order: Probing the Limits
  11. Notes
  12. Contributors
  13. Index