Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

A Final Story: Science, Myth, and Beginnings
A Final Story: Science, Myth, and Beginnings
A Final Story: Science, Myth, and Beginnings
Ebook1,002 pages15 hours

A Final Story: Science, Myth, and Beginnings

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Popular science readers embrace epics—the sweeping stories that claim to tell the history of all the universe, from the cosmological to the biological to the social. And the appeal is understandable: in writing these works, authors such as E. O. Wilson or Steven Weinberg deliberately seek to move beyond particular disciplines, to create a compelling story weaving together natural historical events, scientific endeavor, human discovery, and contemporary existential concerns.

In A Final Story, Nasser Zakariya delves into the origins and ambitions of these scientific epics, from the nineteenth century to the present, to see what they reveal about the relationship between storytelling, integrated scientific knowledge, and historical method. While seeking to transcend the perspectives of their own eras, the authors of the epics and the debates surrounding them are embedded in political and social struggles of their own times, struggles to which the epics in turn respond. In attempts to narrate an approach to a final, true account, these synthesizing efforts shape and orient scientific developments old and new. By looking closely at the composition of science epics and the related genres developed along with them, we are able to view the historical narrative of science as a form of knowledge itself, one that discloses much about the development of our understanding of and relationship to science over time.
LanguageEnglish
Release dateNov 14, 2017
ISBN9780226500737
A Final Story: Science, Myth, and Beginnings

Related to A Final Story

Related ebooks

Science & Mathematics For You

View More

Related articles

Reviews for A Final Story

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    A Final Story - Nasser Zakariya

    A Final Story

    A Final Story

    Science, Myth, and Beginnings

    Nasser Zakariya

    The University of Chicago Press

    CHICAGO AND LONDON

    The University of Chicago Press, Chicago 60637

    The University of Chicago Press, Ltd., London

    © 2017 by The University of Chicago

    All rights reserved. No part of this book may be used or reproduced in any manner whatsoever without written permission, except in the case of brief quotations in critical articles and reviews. For more information, contact the University of Chicago Press, 1427 E. 60th St., Chicago, IL 60637.

    Published 2017

    Printed in the United States of America

    26 25 24 23 22 21 20 19 18 17    1 2 3 4 5

    ISBN-13: 978-0-226-47612-4 (cloth)

    ISBN-13: 978-0-226-50073-7 (e-book)

    DOI: 10.7208/chicago/9780226500737.001.0001

    Published with the support of the Susan E. Abrams Fund.

    Library of Congress Cataloging-in-Publication Data

    Names: Zakariya, Nasser, author.

    Title: A final story : science, myth, and beginnings / Nasser Zakariya.

    Description: Chicago ; London : The University of Chicago Press, 2017. | Includes bibliographical references and index.

    Identifiers: LCCN 2017015353 | ISBN 9780226476124 (cloth : alk. paper) | ISBN 9780226500737 (e-book)

    Subjects: LCSH: Science—Historiography. | Science—Philosophy.

    Classification: LCC Q125 .Z353 2017 | DDC 507.2/2—dc23 LC record available at https://lccn.loc.gov/2017015353

    This paper meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper).

    To Rachel Zakariya, who took a different view

    Contents

    Introduction

    PART I

    1  Varieties of Natural History: The Whole of the Natural and the Known

    2  Dogmas of Unity and Questions of Expertise

    3  The Many Faces of Force and the Mutability of Energy

    4  Schisms

    PART II

    5  Undoing and Reassembling Scales and Histories

    6  Other Emerging Genres of Synthesis: From the Fabulaic to the Foundational

    7  Humanisms, Nuclear Histories, and Nuclear Ages

    8  Scientific Myth and Mysticism

    PART III

    9  Scientific Tribes and Totalizing Myths

    10  Cosmos and the Structure of Epic Myth

    11  Political Cosmologies

    12  A New Version of Genesis

    Coda: Epic Humanisms

    Acknowledgments

    Notes

    Index

    Introduction

    Relating a natural history of the world is a canonical and iconic scientific practice. In the present, its scientific and scientifically invested chroniclers promote a narrative in a mythic register in which they adopt a self-avowedly epic form. The epic relates a history of the universe from its earliest conceivable moments up until and often past the emergence of humanity, performing what the narrators perceive as humanity’s own evolving task of determining the breadth and truth of that story. This historical narrative bears different names indicating different emphases, including the scientific epic, the evolutionary epic, cosmic evolution, the new cosmic myth, the new Genesis, or simply the history of the universe.¹ The historical analysis of the variants of this story, defined through the coevolution of matter and life, including the acts of its authorship and its anticipations of the future of the living and the physical world, is the central subject of this work.

    The terms of this historical narrative as an object of analysis declare the tensions running through its many exemplars. The locution history of the universe implies the related and still more pedigreed universal history. And the invocation of natural history suggests that any historical analysis of a historicizing narrative cannot be confined to histories alone. Natural history still retains museological connotations, suggesting how much its own history as a category was also invested in apparently nonnarrative representations of the world and in the institutions and practices that formed the context of such representations.

    Given the unqualified embrace of the totality that the epic of a universal history details, the analysis of the history of such narratives therefore requires taking account of other modes in which such a totality might be embraced, other natural historical attempts to synthesize the results of what were once often referred to as the special sciences. From the varied contributions that the different scientific disciplines might make to capturing that totality, these attempts stage the construction of a whole of knowledge. As a result, the historical analysis here must treat different, overlapping modes of synthesis. These modes are themselves suggestive of how scientists and other historical actors concede disunity (sometimes explicitly) while attempting to work toward unification. Their efforts are a reminder that the historiographic analysis of the diversity of the sciences does not come to an end with the lesson that there may be no one scientific method that the natural sciences practice in common. Such analysis also requires attending to how historical actors attempt to synthesize disparate disciplines, to answer fears of disunity or promote hopes of a future unity.

    GENRES OF SYNTHESIS

    A central argument pursued here is that despite apparently divergent paths, human/political universal history and the natural history of the universe share recurrent narrative structures and topics, as well as recurrent ambitions and central assumptions, even while their idioms and modes of proof or compulsion may differ vastly. I will call the orienting frameworks structuring these totalizing ambitions genres of synthesis. These are synthesizing modes calling upon a constellation of representational devices including narrative forms, structures of formal and more informal argument, and visual or visually evocative schemes of survey and coordination. Once the basic vocabulary of modern scientific concern with a totalized account of the universe was established in the nineteenth century, these genres have been employed as different strategies for advancing and constituting an organized and unifying synthesis of scientific and broader knowledge. The strategies at play in the construction of synthetic frames have posited ways in which the apparently balkanizing trends in the sciences could still be regarded as enfolded into one whole project, even as the articulation of that project reinflected the understanding of the individual sciences it molded together.

    Four genres of synthesis have largely shaped attempts at achieving a final, natural scientific synthetic frame: the scalar, the historical, the foundational, and the fabulaic. Scalar syntheses construct accounts via a diagramming of space and/or time, often involving the expansion and contraction of spatial scales, or scales of both space and time. Historical syntheses deploy some version of a timeline or thicker historical narrative. Foundational syntheses attempt to reduce (via presumed or implied mathematical and logical consequence) their accounts to the existence of atemporal universal laws. And fabulaic syntheses often rely on quasi-fictional narratives that employ storytelling devices such as a journey taken through different domains. These representations are discernable to an individual understanding in a crisp enough fashion that they can supply coordinates for specific individual research within the disciplines and in relation to each other. More broadly, they position a conception of the human within the genre, in the role that this emplacement may dictate. In professional, generalist, or popular representations, they thus establish understandings of the categories of the scientist and of science.

    They consequently orient all knowledge-producing scientific work as such, which, in the abstract, always undertakes the same task: to sharpen the contours of the very representation upon which the given genre itself is founded. For the cases explored here, the tasks they set include attempting to cover all the scales, from the microcosmic to the macrocosmic, from instant to eon; to tell the full universal history; to drill down to the enduring foundations of things or to construct a scaffold on which they can be said to rest; or to come to the end of the road or quest, passing every station in between. Generically, the scalar is thus often linked to a spectrum, map, or outline; the historical is linked to a timeline, a chronological spine revealed in the growth of scientific knowledge in epic terms; the foundationalist is linked to postulations or claims to the discovery of first principles and the potential ladder of deductive argument extending from them; and the fabulaic is related to and told along the path of a science adventure story.

    In addition, more recently, interdisciplinary and mathematical/statistical analyses promote still other, related candidate syntheses, themselves promoted as sciences of knowledge visualization, potentially establishing new orientations for knowledge-making and new modes for addressing a potential totality. These can appeal to cartographic and organic imaginaries (maps, landscapes, trees of knowledge) in generally descriptive modes attempting to capture the active and evolving links between scholars. But these totalizing visualizations can be prescriptive, as well, in attempting to predict trends and promote research innovations. A proliferation of potential visualizations emerges, for example, on the basis of algorithms appealing to bibliographic data, structured through mathematically developed metaphors of foraging or searching. These may result in new synthesizing genres to a degree independent in character from those that precede them and, perhaps, in the celebration of their plasticity and adaptability to use, offering new orienting and meaning-making frames.²

    In this regard, the synthesis of a body of knowledge tends itself to be rendered as the multidimensional object to be discovered (what scientists are up to, captured by varying measures emphasizing different units of analysis) according to what are regarded as the best principles to determine active research pathways and potential sites of meaning. What such contemporary representations or form-producing algorithms point to in their apparent flexibility is not only the extent to which the entirety of active scientific practices cannot easily be reconciled to any synthetic form, but also the extent to which any such forms must to varying degrees be dynamic and fluid if they are to sustain a connection to the ongoing research, purposes, and desires of scientific practitioners. There is a tension in synthesizing genres, then, between motioning toward the finalization of their forms, as toward a final theory or story, on the one hand, and the project of an open-ended inquiry into ultimate truths, on the other. The notion of finalizability can be anathema to more freely evolving and proliferating representations, the referents for which might prove to be more plastic and therefore less conforming than prior genres of synthesis already being employed. When plasticity is celebrated over and above the retention of a specific form, as may be the case with knowledge visualization, making meaning can tend to center on the promise of innovation as such. Addressing a potential totality of knowledge, this science in particular concentrates more on the organization of a given network through the pragmatics of connecting to information and knowledge understood as useful to some potential searcher. This pragmatism can in turn stipulate that the value of information or knowledge be continually redetermined by the actual or potential use of an actual or potential user.

    Even in such a case, treating abstract principles of use and innovation as fundamental can constrain consequent representations of knowledge as a function of the conventional choices available for operationalizing these principles (as, for example, in measures of use/usefulness based on the circulation or impact of papers and journals). And these visualizing algorithms can extend to representations of the using and innovating human subject as information hunter-gatherer set to the task of rendering that hunting-gathering ever more comprehensive and efficient, pointing to a notion of finalization at work in and beyond the operationalization of these attributes. At the same time, structuring metaphors, such as trees and landscapes, may constrain representational algorithms in such a way as to produce visualizations no more plastic than all the varied modulations of histories or maps of the world, which such visualizations also may to a degree rely on or reproduce.

    A vision of a scientific enterprise/practice as a whole is at play in how, more generally, the genres of synthesis link elements they posit as pivotal to different disciplines. The question of organization therefore is central: the genres conceptualize and have increasingly informed organizational principles and policies for science. The generic representation of the scientific product thus has entailed the representation of scientific organization. But then what constitutes scientific synthesis is itself at stake through that organization, articulating the relationship between disciplines. In the nineteenth century, a central issue of scientific historical debate was the primacy of physical versus biological reasoning, and contested pivotal elements were historical events—a conflict between arguments of physical cosmology and human evolution. In decades thereafter, as we will see, synthesizers posited several different ways of unifying these once historically conflicting sciences to the satisfaction of scientific expert opinion. And as we will also see, these genres influenced resolutions of how research projects should be designed, advocated for, defended, and funded. In providing frameworks for the whole of knowledge, these syntheses established putative objectives for that knowledge. In narrow terms, the syntheses could articulate specific targets: some specific scientific research that should determine this or that chapter in the history of the world, or unexplored region of spacetime or energy frontier.

    In short, then, genres of synthesis structure an approach to the production of knowledge. But they also provide a kind of grammar of ignorance, delimiting what it means not to know, shaping what progress or its lack amounts to and further indicating what might be in principle unknowable: those parts of the representation that can never be resolved or will forever be open to further resolution.

    None of these frameworks has existed in purity, however. Scalar mappings of the universe assume the operation of fundamental laws at one or a multitude of scales, with central sites and often central historical events providing salient markers, while the narratological principles of historical accounts are expressed through the unfolding of the consequences of such laws. Foundational accounts privilege certain ultimate phenomena and laws—ultimate domains of experience revealed historically in space and time—even if this privileging is understood as the universe’s doing. When those experiential foundations are indexed to different scales, so that changes in scale are changes in fundamentals, different scales become different worlds. In turn, fables recounting travels between worlds generally require not only calculations to chart that travel, but the election of fundamental parameters deciding how to reckon that a domain is a world. Given their interweaving, these genres of synthesis can be understood as principles active within myriad instances of universalizing accounts, rather than as pure syntheses in themselves. Nevertheless, one or another is often the ruling figure, with the logics of other genres enabling and coordinated within that figure. A synthesis might for example be housed and structured within a history, the individual chapters of which relate an ongoing emergence at different scales, or the plot and causal structure of which may conform or appeal to fundamental laws.

    The various topics characteristic of these genres of synthesis also were often implicit in each genre and implicitly run together by both historical actors and analysts. The earliest, the greatest, the most distant, the most exotic, the most profound—these were contrasted to the present, the mundane, the near, the familiar, the superficial. But these topical correlations also could be broken. In some of the narratives that resulted, humanity was represented as being of cosmic consequence in its ability to destroy its own and potentially all life; in others, such self-destruction would instead represent nothing more than the disappearance of creatures of no cosmic significance whatever. The correlations could break down still further when the representation threatened to dismiss the importance of an entire discipline, as foundational accounts might relegate to secondary status the importance of anything but the study of the most energetic physics.³

    Importantly, in all these schemata for achieving an account of or revealing an underlying unity in all phenomena, the genres are essentially aspirational. Such syntheses may promise a unity to come, but it is a unity that may or may not be inherent in knowledge itself or in those who author it. It is a unity that may turn out never to exist. In deference to the tension between the potential unity and increasing fragmentation of scientific knowledge and of science as a practice, these synthetic efforts have functioned as a reminder of the acknowledged absence of unity within the sciences.

    To treat accounts of universal history in terms of genres of synthesis thus is in effect simply to name the commitments a number of existing works have in common, works that are perceived to be similar in some respects. Scientific epics or natural historical texts (or, more recently, films and other mixed media) establish a canonical set—a scientific canon of texts or materials—the abstracted features of which amount to a genre. In this sense, many of these synthesizing works were to become or are becoming classics, simultaneously outdated and celebrated by research carrying out the very goals they had enshrined.

    Alternatively, these evolving architectonic frames involve a number of related, consonant investigations or dimensions of analysis that are re-encodings of each other. By representing the effort of scientists and other cultural actors trying to enact, express, and/or represent a meaningful totality, genres of synthesis invoke varied modes for attempting to address totality as such—a totality of nature as given or nature as it becomes known—suggesting how such modes stand in relation to each other. These modes therefore constitute a set of synonyms for the same effort, each taking as central or as most salient a different aspect of scientific materials.

    The first of these saliencies or synonyms is a history of a narrative as synthesis, the development of a story that is itself a contemporary history of the world. As we will see in the first chapters of what follows, in the modes of natural history posited in the nineteenth century by John Herschel, William Whewell, or Mary Somerville, universal history could not provide a synthetic frame. But across the nineteenth century, attempts to claim a unity for the sciences and to provide a synthetic account of the nature of the universe became expressly narratological, set alongside and struck through with other synthetic principles and possibilities. These attempts helped frame the terminological and narrative elements of the genres of synthesis that were to come.

    Intimately related to the historical trajectory of ideas and images involved in the developing genres of syntheses since then are the contestations that also reveal to what extent the synthesizing project was itself a historiographic and more broadly cultural enterprise, an extended debate over whether to address scientific universal history and how such a history should be drafted—whether to construct a synthesis as a whole of knowledge and/or simply to invest faith in a unity that is to be revealed or that transcends human understanding.

    The question of what it means for scientists to carry out and write a history of a totality becomes itself a historiographic concern. Natural history, in its later historical variations, required deciding or discovering (depending on philosophical viewpoint) what might be pivotal universal-historical events and pivotal structures needing elaboration and explanation, what would cause that history to unfold and even allow the universe to be seen as given to a history subject to unfolding. Alternatively, other modes of synthesis, such as the scalar, might establish a spatiotemporal frame and a chain of being housed in that frame. Whatever the genre, the possible and/or implicit limits of such representations have tended to be overshadowed by the project’s totalizing ambitions—ambitions that at the same time must treat as so many unregarded ellipses the majority of events or scales, objects, implications of fundamental laws, or minor stations on the road to scientific insight into the nature of things.

    For the question of universal history, the first and last elements might instead be the most important to understand, the impossible alpha and omega implied by any presentation of a narrative based on the unity of science. Or alternatively, a missing chain in a sequence of what are regarded as formative events might be crucial to the historical account. Here the ratio between the known and unknown, the representation and its completion, is an epic ratio in which a few events structure an immensity of time.⁴ The many architects and draftspeople of varied synthesizing and synthetic genres have attempted to summarize all natural history and to determine specific elements they affirm as pivotal, as with the place of the emergence of life in universal history. Universal historical authors have worked in concert and parallel to stabilize that history by schematizing it through narrative elements it refines and reiterates: its own formative moments (the sequence of its pivotal events) and its shaping principles (how its story unfolds and resolves).

    EPIC, SCIENCE, AND MYTH

    As explanatory models, the relations between such attempts at universal synthesis and myths thus become evident. Functionally, the stories that science has told accounting for beginnings and endings, for the nature of everywhere and everywhen, can be seen as scientific myths. New accounts replacing old ones were being put forward, not stories in a vacuum. Earlier generations had embraced myths as a way to root explanations of what is in the heavens, how old the world is, where humanity—or some part of it—came from, what produced human difference. But part of what became only more distinctive in the period from the mid-nineteenth century and into the earlier part of the twentieth in relation to the natural historical or synthetic work was that there was no simple, certain, or secure cosmological myth, scientific or otherwise, in which to root explication—not in the US and European contexts, at the very least. The multiplicity exhibited by the various genres of synthesis and announced with increasing self-consciousness was often interpreted as deprivation, the absence of a collective myth, a unifying truth, rather than as plenitude, the celebration of many diverse accounts, the acceptance of the multiplicity of truths. New stories were being argued for, while old ones—often understood as well-defined religious accounts—failed to carry the conviction that might have made the composition of new accounts seem foolish.

    The frameworks articulated in the different genres of synthesis mattered on this level because they raised the question of whether it is the business of science, or of the tribe of scientists, to supply society with myths, and as we will see, the supporters of a historical synthesis answered this question with increasing affirmation. Their stories articulated positions as to what falls within the province of science, what could belong only to faith, and what could be rendered into a naturalistic story, what could count as a new secular, rational account, and what was smuggled into such accounts in obscurity. Put another way, the question was how comprehensive an epic history science could produce, and increasingly not whether it should or could seek comprehensiveness. And here, the presumption was that the answers that science can provide to questions about the origins, ends, and dynamics of the universe are all-important, that there is something for science to be presumptuous about: its ability to answer questions previously answered in essentially mythic terms, whether metaphysical, spiritual, or religious.

    If science is seen as the prime figure of institutionalized, experientially determined rationality, and myth as a prescientific explanation of things in the form of a story, long since untethered from direct, publicly producible experience, then a scientific myth presents itself as an enlightened and reasonable tale, a self-interrogated superstition, the rational submission of reason to the need for meaning. Judged by the central sense of its individual terms, the notion of a scientific myth is oxymoronic, or nearly so. But in the multitude of concepts that the term could house, these are only the most immediate tensions. More salient tensions and cultural resonances lay within the inversion of the terms: mythic science.

    Mythic science is an already tamed, if multivalent phrase, the adjectival form of myth bearing little suggestion of objectively false, but rather the sense of epically scaled or famously successful. Both of these latter senses are active in the circulation and reception of these natural histories. But scientific epic is an incongruity in a different way. Like myths, epic tales recount struggles above experience. Alternatively put, they split experience into what is true for their heroic kings and what is true for the subjects who celebrate and suffer the hero.⁵ Science apostrophized as an ambition, by contrast, is alien to such a rupture. The reality to which this scientific ambition still directs itself, regardless of all the half-heard or misheard critiques voiced by critical disciplines, is a reality it posits as shared by all, whether or not that ultimate experience can be made intelligible only by or to a few.

    These contraries within terms mixing science and myth depend on committed distinctions or at least conceptual purifications, science held up on the ground of reason that can penetrate as far as it can latch itself to an actual or even potential experience, uprooting any unnecessary assumption, false presumption, conventional belief—any nonscience. As a conceptual matter, therefore, science is represented as grounding itself. Whatever the breadth of the sources stitched into it, however much its knowledge might re-encode conventions, political resolutions, projected social orderings, or historical contingencies great and small—not generally understood themselves as entangled in the practice or knowledge of science—scientific results are represented as self-justifying.

    But even according to the logic of these representations of science, the divisions between science and myth or between science and epic can stand firm only in certain frameworks. For Theodor Adorno and Max Horkheimer, myth and the Enlightenment remained tightly and problematically intertwined, despite the belief that the inductive sciences were based on the rejection of traditional, mythic authority. Just as myths already entail enlightenment, with every step enlightenment entangles itself more deeply in mythology. Receiving all its subject matter from myths, in order to destroy them, it falls as judge under the spell of myth.⁶ Likewise, epic and myth diverge and are bound together. For these critics, the epic was demonstrably linked to classifying reason, Odysseus’s journey anticipating a scientific calculating mastery of nature achieved through estrangement from nature—an epic as pregnant with a future science and technology as was the modernity Francis Bacon found in the printing press, gunpowder, and the nautical compass.⁷ Successful journeying entails empirical knowledge and at times a self-renouncing, (partial) mastery of nature, the latter itself once a popular formula for the project of science itself. The metaphor of the road to knowledge was present in antique thought and remains a controlling metaphor of knowledge acquisition. The absence of a royal road to geometry revealed the presence of a rougher, longer road, closer to epic travel. The epic understood in terms of quest, from a standpoint taking account of the history and pervasiveness of that metaphor alone, seems to slide closer to science to the degree that science is structured by metaphors of quest. Examined in terms of the history of the coevolutionary account of humanity and the universe, this sliding becomes still more precipitous.

    Regardless of common usage in distinct discourses, myth is often less distant from science than the denotations of the individual terms and their cultural connotations suggest. The terms are repeatedly juxtaposed in a way that does not imagine either dialectic or dissonance.⁸ The expansive present at times elects as the aim of the totality of scientific knowledge the bridging of the distance between the scientific representation of the universe, on the one side, and the resonance, promise, and subject matters of myth, on the other. In this line of thought, origin is a middle term between science and myth, science seeking the ultimate origins (of matter, of life, of thought) once apparently imagined as disclosed by older mythologies. Myth can establish a discourse representing what, within the mythic representation, are the ultimate origins of everything, origins pregnant with ultimate ends. These mythic origins need not narrate a beyond-themselves; myth voices its truth out of a silence that it does not narrate or approach. Just as science provides its own ground, myth likewise can remain silent over the possibility of a further origin, over another ground in which to root out a further license to relate its account. Once science, as celebrated by prominent expositors, explicitly includes in its manifest ambitions the discovery of such an ultimate ground for everything, a ground that itself can be expressed narratively and that does not admit an outside of its own story, the designation of myth is to hand, and indeed, it fired the imagination of a plurality of scientists and commentators over the past several decades.⁹

    But myth and epic each bear heavy conceptual and cultural loads separate from science. Myths shoulder histories, persons, and practices foreign to even the most expansive formulation of the task of science. The concepts of scientific myth or evolutionary tales are by these lights generic monsters that themselves bear prodigies. Odysseus becomes Sisyphus, the hours of Gilgamesh’s dark run are prolonged: the ritual practices and finalizations posited by nonscientific cosmologies or secured by the homecoming of the king are not as clearly won by a scientific history looking with less certainty to the future. For some practitioners, the central reward of a scientific history is simply the possibility of determining more of it.

    If the relevant concepts are rendered crisply enough, scientific myth can be included in the wider category of scientific narrative or scientific story. These more general terms are suffused with their own, more dispersed tensions, depending on how close or far apart truth and story are held for consideration. Narrative is no more foreign to truth, and no less biased by individual or collective perspective, than is history. Like myth, story in reference to science is used both to advocate scientific history and to dismiss it. A history so extensive may, as in some exemplars, dizzyingly project a historicization of time, exciting in some of its producers and consumers awe for what science has been able to accomplish or for the vision it projects. For others, that same extensiveness immediately negates itself, revealing how far beyond the reach of empirical license or conceptual integrity such a history must apparently extend and provoking a diagnosis of scientific storytelling or nonsense.¹⁰

    But a totalizing or universalizing history is not the only conceptual framework providing such an inclusive shelter for the varied products and producers of knowledge. Myth and epic as forms have themselves been theorized in terms of their aggregative or totalizing character. That character is also found in synthesizing genres or frameworks often not overtly presented in narrative terms. For example, the mapping of the world or of the knowable or surveyable universe engages a similarly totalizing ambition. A second, more fantastic example is the possibility of visiting every world, with or without a map, as a mode to capture everything, an idea itself constituting a fable with an indefinite number of ellipses. Mathematics or mathematical logic holds out the hope of a still more abbreviated form of synthesis, in which some ultimate predicate or law is valid without qualification and out of which the domains of knowledge and experience can be derived by actual or in-principle mathematical and logical manipulations.

    These different modes of synthesis all have shared histories, foremost because, as we have noted, they rely on each other. But in relation specifically to the history of the universe, treating these genres together also serves to illuminate how narrative itself can function as a synthetic form—how story can achieve or constitute what scientific theory promises. To make this latter claim, however, requires positing a slippery distinction between synthesizing or generic activity (efforts to actualize such syntheses) and dogmatic belief that the unity promised by these syntheses is already active, without any need for extended proof or elaboration. A belief in a foundational law and its mathematical consequences might be dogmatic, a matter of belief in no derisive sense. But such a dogmatic belief can be distinguished from any scientific practice actively attempting to establish a more thorough, foundationalist frame, theorizing and organizing scientific practice according to that conception. Similarly, the idea that the universe as a whole has a history can be a dogmatic assertion, one that has at times been rejected as, for example, an anthropomorphic conceit or as hypostatizing the totalizing, possibly fictive idea of world at play within it. By contrast, presenting, promoting, and working to establish the aim of the sciences by way of authoring a historical account reveals and turns on the synthetic character of narrative itself. The history and functions of totalizing scientific narratives underscore differences between belief in unity and synthesizing practice.

    BETWEEN STORY AND THEORY: THREE EXAMPLES

    To clarify and concretize these arguments and distinctions, it is necessary to examine at length and in depth the extended development of contemporary genres of synthesis, their composers and composition, their refinement, distribution, promised finality and form, the dilemmas at stake within them, the organizational structures their architects promoted through these genres and that promoted, shaped, and refined them in turn. That is the task of the many pages that will follow. However, a particularly crisp, accessible moment of roughly simultaneous synthetic effort in the late 1970s offers a preface to this more extended history of synthetic genres, underscoring the issues broached here insofar as many of the terms of analysis are explicitly invoked.

    At a moment of summation in his first generalist, popularist text, The First Three Minutes: A Modern View of the Origin of the Universe (1977), the physicist Steven Weinberg stated: We are now prepared to follow the course of cosmic evolution through its first three minutes.¹¹ Having presented his reader with the relevant scientific background, Weinberg steps back in a kind of dramatic suspension before delivering the central matter of the work, a fluid exposition of the first moments in the history of the universe as the temperature of the universe fell quickly with its expansion. It is not an even, straightforward chronological presentation: Events move much more swiftly at first than later, so it would not be useful to show pictures spaced at equal time intervals, like an ordinary movie. Instead, I will adjust the speed of our film to the falling temperature of the universe, stopping the camera to take a picture each time that the temperature drops by a factor of about three.¹² Weinberg details the first six frames, from a period when the universe was a hot cosmic soup of elementary particles to three and three-quarter minutes later, when the basic ingredients of matter formed. At the end of those minutes, Weinberg observes: The universe will go on expanding and cooling, but not much of interest will occur for 700,000 years.¹³ After this, conditions will change in such a way as to allow matter to begin to form stars and galaxies, but even this is of less interest to Weinberg, since the ingredients with which the stars would begin their life would be just those prepared in the first three minutes.¹⁴ Concluding in an autobiographical register that includes all of humanity, Weinberg dryly observes that after another 10,000 million years or so, living beings will begin to reconstruct this story.¹⁵

    Weinberg is still more precise with his timeline, specifying more exactly the moment when the authorship of that scientific story became possible. According to him, during most of the history of modern science, there simply has not existed an adequate observational and theoretical foundation on which to build a history of the early universe.¹⁶ But from the point of view of the middle 1970s, now, in just the past decade, all this has changed, and advances in scientific theory and practice allow the scientific story to be told. As a story with adequate observational and theoretical foundation, it has to Weinberg’s mind greater merit than the origin myths that preceded it—Norse myths are his chief example—while looking to resolve the same questions: What could be more interesting than the problem of Genesis?¹⁷

    In a similar spirit, in the same period, the sociobiologist E. O. Wilson found the scientific story to be one that had virtues beyond those of the accounts that preceded it: Indeed, the origin of the universe in the big bang of fifteen billion years ago, as deduced by astronomers and physicists, is far more awesome than the first chapter of Genesis or the Ninevite epic of Gilgamesh. When the scientists project physical processes backward to that moment with the aid of mathematical models they are talking about everything—literally everything—and when they move forward in time to pulsars, supernovas, and the collision of black holes they probe distances and mysteries beyond the imaginings of earlier generations.¹⁸ Wilson quotes the biblical Book of Job, where Job is chastised by God, asked to give an answer to whether he has descended to the springs of the sea, walked in the unfathomable deep, comprehended the vast expanse of the world. Come, tell me all this, if you know. To which Wilson in the place of Job responds, "Yes, we do know and we have told.¹⁹ For Wilson, the fundamental theoretical approach of science, its material and unifying basis, or as he terms it, scientific materialism, has provided and told all these answers, allowing him and the modern generation to answer the biblical God. And for Wilson, the core of scientific materialism is the evolutionary epic," his term for humanity’s story of the world and humanity’s place in it.

    Like Wilson, the astronomer Carl Sagan in his 1977 Dragons of Eden: Speculations on the Evolution of Human Intelligence also found a resonant figure in Job, invoking in an epigraph words in which Job bemoans his condition: I am a brother to dragons, and a companion to owls. That epigraph, cueing it would seem the title of the text,²⁰ reflected Sagan’s belief in the deep-seated place of myth in humanity’s self-understanding, a vision in a somewhat different spirit, but consonant with Wilson’s connection of myth and science. In sentiments on similar themes made in an address that same year and republished in 1979 in an essay entitled The Golden Age of Planetary Exploration, Sagan lists what he saw as the many virtues of planetary exploration. The last dwells on the present moment in history and the histories that science could tell: And it [exploration] permits us, for the first time in history, to approach with rigor, with a significant chance of finding out the true answers, questions on the origins and destinies of worlds, the beginnings and ends of life, and the possibility of other beings who live in the skies—questions basic to the human enterprise as thinking is, as natural as breathing.²¹ For Sagan, Interplanetary unmanned spacecraft of the modern generation extend the human presence to bizarre and exotic landscapes far stranger than any in myth or legend.²²

    This reference point of myth and legend involves not only the origins and fates of other worlds, but those of this world and the history of the universe as a whole. When Sagan comes elsewhere to modern cosmological theories and finds in them uncomfortable analogies and connections to the experience of human birth—analogies that he ultimately takes to be at the root of religion—he queries almost anxiously: Are we such limited creatures that we are unable to construct a cosmology that differs significantly from one of the perinatal stages? Is our ability to know the universe hopelessly ensnared and enmired in the experiences of birth and infancy? Are we doomed to recapitulate our origins in a pretense of understanding the universe? But his final question in this same series is aligned more with an optimism for an unconstrained capacity to know, an optimism that was the dominant key of his works: Or might the emerging observational evidence gradually force us into accommodation with and an understanding of that vast and awesome universe in which we float, lost and brave and questing?²³

    In the composition and reception of this history of the world, the word story, its cognates and synonyms and the ambitions they express for the rendering scientific truths are pervasive.²⁴ For those advocating a new history or myth, those terms were generally not dismissive; they did not suggest that scientists and other synthesizing architects attempted to build a false structure. For these authors, the opposite was the case: the use of mythic categories and registers reflected a belief in and an ambition to relate a true story of the universe, a story that might translate, structure, capture, or serve as the theory that finally gives account of everything.²⁵

    MODELS AND OBSTACLES: HISTORY AND EPISTEMOLOGY

    In their primary representations, the scientific ways for addressing this narrative were neither remade from whole cloth nor simple iterations of past efforts. The language they employed in relation to natural science, the language and structures of story, of history, of myth, and of epic, clearly did not begin with the twentieth century or with the nineteenth. Generic models of scientific narrative existed well before modern scientific histories offered the possibility of concise, thoroughly naturalistic accounts. The cultural reception of these scientific histories would not only juxtapose them to sacred histories, but compare them with the scholarly and popular tradition of universal history.

    Universal histories and their connection to sacred history itself were under ongoing extensive interrogation and critique. From earlier universal histories on, there was a pattern of fluctuation and diversity in natural scientific views of the history of the world—even of whether the world has a history and whether science could speak to it—varying from acceptance to denial and from contestation, to indifference, to celebration. For the larger part of the nineteenth and twentieth centuries, no one view commanded so much respect as to render its skeptical responses unscientific, even if there were dominant claims. Likewise, the advisability and true possibility of world/human/species history would become subject to extensive debate and critique.

    At the time when the primary account here begins, in the second third of the nineteenth century, deliberations and debates over natural history established strong conceptual obstacles to framing a naturalistic synthesis of natural knowledge in the form of history. Natural history, understood in the sense of providing an intelligible order of the natural world, had included elements in that order—the pieces of its imagined museum—understood as unchanging. A stable solar system or the fixed stars did not betray or disclose their origin. Whereas the eighteenth-century French mathematician Pierre-Simon Laplace’s physical astronomy was held to demonstrate the perpetuation of the planets in their orbits, his theory of the creation of the solar system was received not only with enthusiasm, but also with the défiance, the distrust, with which he advanced it.²⁶ According to this theory, a great solar cloud contracted and at its fringes, as it condensed and cooled, it successively formed the planets. The theory thereafter was described as the nebular hypothesis, a term coined by William Whewell to name one of in fact a collection of varied hypotheses dealing with the variety, dynamics, and potentialities of such celestial clouds or nebulae.²⁷ These theories generally argued for the historicity of entire stellar systems and, at times, of all the fine-grained chains of being they might support. But the ahistorical objects of the natural world demanded fields of knowledge that themselves could not be suborned to the field of history and to its repertoire of historical productions.

    This departure point, when history could not easily establish a synthesis of natural knowledge, helps to identify the marks of the genre of more recent scientific universal history—its debts to and contrasts with scientific history of an earlier period, clarifying as well aspects of those past natural histories. The presumptions or explicit arguments that related and countered the physical, biological, and geological sciences as they themselves developed were steeped in and dwelled on the cultural dissonances and consonances discernible in new histories and syntheses.

    That longstanding world or universal historical tradition in turn suggests that modern natural history drew from persistent narrative conventions, themselves preceding the nineteenth century. This persistence will be overstated if such older conventions and narrative models are imagined merely to reappear in the future—if, for example, new epic histories are themselves understood as merely iterations of sacred history. Much historiography basing itself on a widening arc distinguishing natural history and the world history of peoples from at least the seventeenth century on argues compellingly against any such simple iterative view of historical genre. From such a perspective, for centuries preceding the nineteenth century, the dominant framework for Western histories was being stretched almost out of recognition. Earlier voyages of discovery had revealed a natural geography and a distribution of cultures, peoples, and recorded histories that placed new and stringent demands on the old, scripturally sourced accounts. The newer modes of empiricist temper arising over the same centuries argued for new kinds of proof for historical claims.

    Across this period, according to this genealogy, historical knowledge about the world appeared to fragment into different modes, depending on whether rocks and nations were read to confirm scripture or to threaten it. Differences in methods, in temporal and spatial scales, and in the subject matter of natural and civic or human histories became only more multiple and manifest. Human history as a whole, much less the history of states, organizations, populations, or individuals within it, occupied an increasingly small moment of chronological/mechanical time, a time itself posited as an ever smaller fraction of a more quickly aging world. Conversely, the history of the natural world could be excluded from human history. There remained some few overt connections between written history and the history of the universe—as when the annals of world history were plumbed for confirmation of the periodicity or nature of astral events—but the documentary emphasis and methodology of each kind of history only intensified differences attributable to paradigmatic historical and natural scientific methods: the reading of strata and skies versus the reading of texts.²⁸

    The general point of view of such a history is that the seventeenth century nursed a process of sacred historical revision in the production of a history of the world (as for example in the genre of sacred histories of the Earth or in relation to conceptions of and politics toward the other in the world) and that the eighteenth century saw the features of a newer universal natural history break more distinctly from earlier patterns (with the arguments of those such as Jean-Andre Deluc, Georges-Louis Leclerc, Comte de Buffon, Johann Friedrich Blumenbach, and many others).²⁹ The widening, aging, pluralistic world had placed pressures on the natural and scriptural history of the West, pressures that at the same time produced innovations, reinterpretations of sacred accounts, or recalibrations. These provided new models for scientific/naturalistic histories, which appear to be cast out of radically different molds from either sacred or (human) universal histories. That same extended period of time saw the further elaboration of all-embracing universal history as a scholarly genre, a genre itself drawing on the forms and dilemmas of sacred history and embracing the content of natural history. But even by the lights of this clear a genealogy, in those newer molds, the older templates still played a structuring role. Through to the turn of the nineteenth century, such reliances were often overt: Genesis might continue to provide an initial chapter of a world history, or, less conspicuously, scripture could sanction the imagined alliance of the character and history of peoples and places.³⁰

    There is much that is enlightening in the sharp lines of such a genealogy. It helps make clear how, by the early nineteenth century, a cultural ground had been prepared for the widening intellectual foment and subdivision in the years thereafter, giving the intellectual West a multiplicity of candidate cosmologies and so, at the same time, no cosmology. A virtual absence of learned consensus—the absence of an accepted cosmology—lasted for most of the nineteenth and twentieth centuries. There was dissensus between disciplines studying the natural world. Biology and geology at times required an earth that outlived the nourishment that seemed available to sustain it; observational astronomy found stars older than the heavens in which they were embedded. At the same time, the divisions between civil and natural history were stark, not only in content and methodology, but according to the lessons and character they were imagined to instruct and instill. Apostles of humanistic values collided with disciples of the varieties of natural scientific ethos and cultures, a collision ramifying through and constructed by a changing matrix of social class and of political and pedagogical movements. It was not until the latter part of the twentieth century that this period of foment appeared to come to some conclusion, when the chronologies of the material and organic world posited by the different branches of the natural sciences increasingly converged and synthetic accounts, finalizing theories and stories, became possible. This still-present moment, in the recent historical genre of Anthropocene history, likewise has seen attempts to bridge humanistic and scientific historical narratives, promoted as a reconciliation establishing a species truth through a universalizing account.

    The new contemporary myth or universal history, as narrated by scientific practitioners or historians, is often characterized as a late chapter in the unfolding of the Enlightenment. These histories emphasize that the origin of things is distant from the Word and Eden. The conclusion of the universal history they assemble is found in species extinction and cosmic death or in an unknown human potential or future human transformation, a conclusion immediately alien to, for example, an idea of reconciliation with divinity according to the dictates of providence. The overall account of the history of knowledge embedded in these modern universal histories instead has the flavor of a human epic. According to it, the falsehoods of the past are finally cast off, giving rise to the contemporary synthetic viewpoint, to a day when historically minded scientists and scientifically competent historians can relate the rise and future fall of humanity, of prehistoric and posthistoric cultures, and human nature itself. Such genealogies tend to posit their own telos to the evolution of knowledge: the composition of the history of the world is approaching the truth that the material history of the world is approaching no truth, no culmination at all.

    The contrast between epistemological and material evolution and its implications can be best brought out and examined by looking at the steps in that genealogy. Scholars from a variety of disciplinary standpoints largely accept that the historical disciplines fragmented as natural and civil histories became more reliant on their different expert knowledges. The current scientific histories written by scientists, historians, and other cultural producers, such as film documentarians, see themselves as effecting reconciliation in the wake of that dissolution.

    But any sense of current reconciliation raises the question of how different kinds of history have continued to rely on each other in deep ways, even across their apparent fragmentation, and how they have been understood as companions to each other. How does a progressive account of successively truer universal history potentially depart from and conform to important aspects of world/universal historical ambition and character? Likewise, how might contemporary universal histories continue to be invested in the sacred traditions they explicitly reject and from whose narrative content they have so long departed? How have histories of the universe and universal histories been brought closer together as a function of both repeated critiques of their universalist character and repeated attempts across the last century and more to stitch them together on the assumption that a reconciliation is necessary? The epic trajectory of knowledge embedded in the scientific epic tends to bracket the historical and conceptual connections between the modern categories of scientific and humanistic universal histories, connections often suggested by those past authors most often celebrated as forebears to modern cosmology and universal history.

    The Enlightenment offers no simple picture or belated starting point for this trajectory. Recent historiography functions as a reminder that there is no simple alignment between secularity and Enlightenment thought, troubling as a result the picture of the divergence of varying historical modes—particularly between sacred and scientific/secular, but by extension, between sacred and universal/world-historical.³¹ But even the central figures adduced within historiographies asserting increasing divergence present by the example of their efforts no simple picture of the distinctions between emergent categories of scientific and other histories.

    In 1791, Kant’s Universal Natural History and Theory of the Heavens was published, not long before Laplace’s System of the World (1796), each putting forward varieties of the nebular hypothesis. Kant’s cosmology had been printed nearly forty years earlier in 1755 and well before his Idea for a Universal History with a Cosmopolitan Purpose in 1784—the same year as his famous essay What Is Enlightenment? and three years into his critical period, inaugurated with the 1781 Critique of Pure Reason.³² In 1755, the young Kant had rhapsodically enlarged the story of everything in The Theory of the Heavens, as it is often abbreviated. He theorized a phoenix of nature, an unending creation taking place across an endless universe. Kant and Laplace addressed the origins of the solar system and (at different points in their works) the formations beyond it: Kant extended the Newtonian system in an extrapolation that Newton would not, at least at times, have held to be legitimate, reasoning with an idea of gravitation as active at all times and places and imagining the death of old worlds feeding the birth of new ones.³³

    Kant’s universal natural historical reflections addressed and reconfigured a natural theological concern: What sustains the world? Without some form of sustenance, the world and the solar system, it seemed, would spend their force and halt. The concern as to what sustains the world system, a concern held more generally, persisted in different form in the natural histories of the following centuries. For those who found the possibility of universal death inconceivable, God provided the ultimate assurance of new energies to come. As a natural theological argument, this was close to circular, since the natural world does not in all respects give signs of its endless perpetuation.

    Three decades later, his Idea of Universal History grounded future political hopes in the natural character of human interaction present from the state of nature. These different conceptions and treatments of universal histories, natural and human/political, suggest points of both divergence and connection between different synthetic histories, natural, universal, and, through his treatment of each individually, the sacred tradition. The divergence might be grounded in the disunity of Kant as a historical personality whose intellectual convictions change over time. But there are also the persistent themes and sources prompting the different works and theses circulating within a decade of each other that connect them through and beyond the biographical personality signing the same name to them. How did the matrix of these histories—all universal in different senses of that word—remain in dialogue, even as they adopted their own generic frames?³⁴

    In that generic and critical work, Kant provided an ongoing resource and target for considerations on how natural and universal history were integrated in the conceptual relations of the sciences, each to each. The sciences in broad terms could be distinguished and made possible by elucidating the grounds on which knowledge is possible—on a knowledge that structures experience, rather than being informed by it. In his view, to ground knowledge, and so to clarify what makes the sciences possible and their proper relations to each other, constrains the ambitions of both a natural history and a universal history.³⁵

    But Kant’s own speculations on the heavens at times connected more immediately to other historical initiatives, just as his universal history also connected more immediately to the natural sciences. In The Theory of the Heavens, Kant tied his cosmological reflections to the massive universal historical project of the day, the multiauthored, multivolume Universal History from the earliest Time to the Present, compiled from Original Authors, and illustrated with Maps, Cuts, Notes, Chronological and other Tables,³⁶ looking to it as a justification for the propriety and sense of his natural historical labors: "I will therefore quote the Authors of the Universal History, when they say: ‘However, we cannot but think the essay of the philosopher who endeavoured to account for the formation of the world in a certain time from a rude matter, by the sole continuation of a motion once impressed, and reduced to a few simple and general laws; or of others, who since attempted the same, with more applause, from the original properties of matter, with which it was endued at its creation, is far from being criminal or injurious to GOD, as some have imagined, that it is rather giving a more sublime idea of His infinite wisdom."³⁷

    These universal histories had themselves been extensive, multigenerational, collective projects—and as in this quote, themselves relied on potential naturalistic cosmogonies and sacred sources. Roughly three-quarters of a century prior to Kant’s effort, Bishop Bossuet, who had been tutor to the eldest son of Louis XIV, addressed the dauphin in a universal history rooted in the mosaic cosmogonic account and the biblical peoples, a chronicle up to the time of Charlemagne.³⁸ Voltaire built on and critiqued Bossuet, attempting to recount the histories and periods of nations that his predecessor ignored or did not reach.³⁹ At roughly the same time, the English Arabist George Sale, along with colleagues at Oxford and Cambridge, contributed to the production of a many-volume universal history; their joint work attempted to canvass the recorded histories of every nation, but was also grounded in a biblical cosmogony. It was the translation of this work by Siegmund Jacob Baumgarten that Kant apparently examined in 1755.⁴⁰ While these ambitious chronicles were being written, themselves across intergenerational, multilingual networks, natural historians and philosophers such as Buffon reworked the origin accounts in which the universal histories were rooted, ultimately suggesting more extended, naturalistic cosmogonies and anthropogenies.

    Tying his efforts to these, Kant’s early Weltgeschichte advocated a view of the world that saw in gravity a universal and timeless force that, in forces of attraction and repulsion (by more contemporary lights, a term for the more fictive centrifugal force), formed the galaxies out of matter distributed everywhere. He attempted to demonstrate his views by making sense of a set of observations concerning the orbits of the planets, their shared directions of revolution, the eccentricity of their nearly circular orbits, their masses, and so on, which in many ways were shared with the later, partly tentative reflections of Laplace. In aid of this demonstration, he argued for a view of the fixed stars as moving and for the hazy nebulae in the heavens as their own as yet unresolved and unknown star systems. He found in these hypotheses a way to unify contrary positions, the one holding with Newton that there was no way to imagine the present configuration of the solar system to have emerged without God’s intervention, the other suggesting a mechanistic view of the world explaining the many coincidences of the system by appeal to a natural origin in time. He presented his speculations as tied to the strongest proofs of the existence of God.⁴¹ If this was an antinomy of sorts that he posited and resolved, thirty years later, he would address antinomies of origin in starkly different fashion.

    The older Kant provided an Enlightenment-era vision of the way in which such a scientific history, however compelling, might never amount to knowledge, confronting fears that long predated those of Carl Sagan that a possible anthropomorphism structures any cosmogony. In this later critical work, Kant believed himself to have proven that the ontological proof of God’s existence is impossible. And on the way to that proof, he argued that a full history of the universe that would imagine itself to know the origins of everything also is impossible. The restriction against telling absolute origins, as he explicated it, is a constraint on the powers of reason itself.

    These constraints were announced with the publication of his 1781 Critique of Pure Reason.⁴² The tone of the exuberant Theory of the Heavens was gone. But however much the tone of his later works differed from his earlier cosmogony, they revealed abiding concerns dating from his first natural historical efforts. Regardless of how much the older and younger views and styles differed and evolved over time, it would be as much an oversimplification to treat the older Kant’s work as simply contradicting that of his younger self as to treat the elder’s work as a mere extension of the younger.⁴³

    The First

    Enjoying the preview?
    Page 1 of 1