Quantum Language and the Migration of Scientific Concepts (The MIT Press) 0262037556, 9780262037556

How highly abstract quantum concepts were represented in language, and how these concepts were later taken up by philoso

613 70 1MB

English Pages 336 [337] Year 2018

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Quantum Language and the Migration of Scientific Concepts (The MIT Press)
 0262037556, 9780262037556

Table of contents :
Title
Copyright
Dedication
Contents
Introduction
1 Experience, Perception, and the Limits of Language
2 The Physics of Visuality, Intuition, and Aesthetics
3 Quantum Paradigms in Literary Criticism
4 New and Post-New Age Appropriations
5 Quantum Versus Nuclear Discourse
Conclusion
Notes
Bibliography
Index

Citation preview

QUANTUM LANGUAGE AND THE MIGRATION OF SCIENTIFIC CONCEPTS

QUA N T UM L A NG UAG E A ND THE MI GRATI O N OF S C IE NT IF IC C ONCEP TS

JENNIFER BURWELL

The MIT Press Cambridge, Massachusetts London, England

© 2018 Massachusetts Institute of Technology All rights reserved. No part of this book may be reproduced in any form by any electronic or mechanical means (including photocopying, recording, or information storage and retrieval) without permission in writing from the publisher. This book was set in Bembo Std by Toppan Best-set Premedia Limited. Printed and bound in the United States of America. Library of Congress Cataloging-in-Publication Data Names: Burwell, Jennifer, author. Title: Quantum language and the migration of scientific concepts / Jennifer Burwell. Description: Cambridge, Massachusetts ; London, England : The MIT Press, [2018] | Includes bibliographical references and index. Identifiers: LCCN 2017028080| ISBN 9780262037556 (hardcover ; alk. paper) Subjects: LCSH: Quantum theory--Philosophy. | Physics-Philosophy. | Quantum theory in literature. Classification: LCC QC174.13 .B874 2018 | DDC 530.1201/4--dc23 LC record available at https://lccn.loc.gov/2017028080 10

9

8

7

6

5

4

3

2

1

for Ella and Maeve

CONTENTS

INTRODUCTION

1

1

EXPERIENCE, PERCEPTION, AND THE LIMITS OF LANGUAGE

2

THE PHYSICS OF VISUALITY, INTUITION, AND AESTHETICS

3

QUANTUM PARADIGMS IN LITERARY CRITICISM

4

NEW AND POST-NEW AGE APPROPRIATIONS

5

QUANTUM VERSUS NUCLEAR DISCOURSE

CONCLUSION NOTES 265 BIBLIOGRAPHY INDEX 317

255

297

169 211

125

27 69

INTRODUCTION

Almost a century after its main principles were established, quantum physics remains one of the most conceptually elusive theoretical paradigms in science—so elusive that even its original architects were confounded by the results that their calculations produced. It also remains one of the most figuratively allusive paradigms, a fact that cannot be separated from the baffling nature of its principles. The tenets of quantum physics—and the strange phenomena that they describe—originate in and are expressed most precisely by highly abstract algebraic equations. The main challenge posed by quantum phenomena does not lie, however, in its mathematics; it lies instead in how these phenomena in their very nature strain the limits of comprehension and representations: electrons that behave sometimes like particles and sometimes like waves; atomic systems that exist simultaneously in all possible states—until they are observed; electrons that reveal their position only if their momentum remains a mystery; and particles that exist at great distances from one another but appear to “know” and respond to what each other is doing. The counterintuitive nature of quantum theory—and especially its curious relationship to everyday experience and representation in language— is the engine for a set of key questions that I address in this book. Why, for instance, did these mathematically driven concepts compel their founders to spend so much time reflecting upon ontological, epistemological, and linguistic concerns? What do quantum phenomena, considered in the context of the founders’ reflection and debates on how to describe them, reveal about the relationship between everyday experience, perception, and language? How—and why—do quantum concepts get taken up again fifty years after their formulation in cultural contexts far removed from their origins in physics? What, for example, made

2

Introduction

quantum theory appeal to advocates of Eastern mysticism starting in the 1970s, literary critics starting in the 1980s, and contemporary hawkers of distant healing and get-rich schemes starting in the 1990s? What is revealed about the agendas and priorities of these later iterations in the ways that they interpret and misinterpret the original quantum concepts? Finally, how can a comparison of the quantum discourse that emerged from these iterations and the “nuclear discourse” that emerged around the development and detonation of the atomic bomb illuminate both to greater advantage? As I progress through the book, working out and through these questions, I offer the reader a tale of conflict, manipulation, personal profit, and mythmaking. In chapters 1 and 2, I focus on core concepts that Austrian physicist Erwin Schrödinger, Danish physicist Niels Bohr, and German physicist Werner Heisenberg developed: complementarity, indeterminacy, wave mechanics, and matrix mechanics. Each of these men made other significant contributions to the field of physics that I do not discuss in these chapters: Heisenberg made important breakthroughs in hydrodynamics and nuclear physics, Bohr was a central figure in the Manhattan Project that developed the atomic bomb, and Schrödinger contributed in essential ways to the quantum principles of nonlocality and superposition— two concepts that become central in chapters 3 and 4, but are not of major significance to chapters 1 and 2. I have also chosen not to include— except in passing and in how they relate to the theories of Schrödinger, Bohr, and Heisenberg—the fundamental contributions to the quantum interpretation made by other physicists and mathematicians such as Paul Dirac, John Von Neumann,Wolfgang Pauli, and Albert Einstein. A number of factors drove my decision to focus on Schrödinger, Bohr, and Heisenberg. First, these three physicists assembled the key concepts that made quantum physics cohere as a theory—to the point where their names are virtually synonymous with the birth and evolution of quantum physics. Second, the debates that surrounded the quantum interpretation circled for the most part around these three men, with Schrödinger typically on one side, and Bohr and Heisenberg on the other. Third, Schrödinger, Bohr, and Heisenberg all reflected on the questions of visualizability, intuition, and aesthetics with respect to quantum physics— questions that compose the primary concerns of my second chapter. Finally, all three men were exceptional in their extensive engagement

Introduction

3

with quantum physics’ problematic relationship to the constellation of language, experience, and sense perception—a relationship that is central to my analysis in chapter 1, and to my analysis in chapters 3 and 4 of the later uses and abuses of quantum concepts. Schrödinger, Bohr, and Heisenberg were unusually reflective about the relationship between language and the material world. For these three physicists, quantum physics proved to be as much a communicative experiment as it was a material one, and as much about language as it was about mathematics. This preoccupation emerged from the way in which the so-called “quantum interpretation,” whose logic defies our sensebased experience of the material world, poses special challenges to communication. Even to name a quantum phenomenon is already to introduce a metaphorical distortion, a kind of semantic drift that imposes conventional concepts onto utterly novel phenomena. So contentious was the question of language and representation for Schrödinger, Bohr, and Heisenberg that a bitter dispute arose between, for example, Schrödinger, who remained committed to using classical language and concepts, and Heisenberg, who frequently advocated limiting the representation of quantum behavior to the exact mathematics. Understanding why this dispute arose, and why the quantum interpretation proved so resistant to representation, requires some knowledge of the primary concepts of quantum theory that I will summarize below. THE BIRTH OF QUANTUM PHYSICS

If you think you understand quantum mechanics, you don’t understand quantum mechanics. —Richard Feynman

From the time of the ancient Greeks, it was believed that all matter was composed of small, indivisible particles called “atoms.” This belief persisted through Newton’s “corpuscular” theory of light, which he introduced in his 1704 treatise Optics.Thirty-two years earlier, Dutch physicist Christiaan Huygens had established mathematically and experimentally an alternative interpretation—that light was wave-like in nature—but Huygens’s theory was never accepted, and Newton’s particulate atomism won the day. Only when Thomas Young performed his famous

4

Introduction

double-slit experiment in 1803, demonstrating the wave quality of subatomic matter, were the foundations of Newton’s corpuscular model shaken. Using sunlight, Young passed light through a screen with two pinholes in it and recorded the results on a detecting surface placed behind the screen. If light were particulate, one would expect to see a collection of dots the size of the holes show up on the plate. Instead, alternating lighter and darker bands appear, indicating a process of interference that could only be possible if light were wave-like. Young’s experiment eventually led to the overturning of Newton’s model, and the wave theory of light seemed to have prevailed. Young’s wave model endured throughout the nineteenth century. Then, in 1900, Max Planck laid the foundation for quantum theory when he proved that energy could only be emitted in discrete, particle-like units, with each unit making up a “quantum” of energy. Plank’s breakthrough was reinforced by Einstein’s work on the photoelectric effect, and in 1905 Einstein proved that light was formed of discrete, particulate “quantum packets,” later known as photons. If Young had proved that light is wave-like, Planck and Einstein had just as convincingly proved that it is particle-like, and the theory of light seemed to have reached an irresolvable impasse.The situation became even more complicated when, in 1924, French physicist Louis de Broglie demonstrated that not just light, but all subatomic particles behave in some respects like waves and in some respects like particles. This meant that both the wave and particle theories were right: somehow, subatomic phenomena were both waves and particles. The “dual” nature of photons, electrons, neutrons, and other atomic energies challenged received notions of atomic matter as being selfconsistent and unchanging; even more than this, it challenged our basic notion of an entity, which must by its nature be one thing or another. A modification of Young’s double-slit experiment, predicted mathematically in 1925 by Erwin Schrödinger, revealed an even stranger aspect of subatomic matter.1 If both slits are open, but only one single photon is emitted, a remarkable result ensues: a wave pattern of light and dark bands appears on the photographic plate, as if this single particle has traveled through both slits simultaneously and “interfered” with itself. This interference only occurs, however, if no detecting instrument is introduced, and the photon is allowed to proceed “unobserved” toward the screen.

Introduction

5

Once a detecting instrument is introduced, the photon appears to know that it is being watched, at which point it decides to behave as a single photon ought to, producing a particle-like dot pattern. On a subatomic level, then, it appears that it is impossible to separate the act of observation from the behavior of the object, and that observation in fact constitutes the object. The months between early 1925 and the middle of 1927 proved to be determinate for quantum physics, with the field’s foundations— Schrödinger’s wave mechanics, Max Born and Pascual Jordan’s matrix mechanics, Heisenberg’s Uncertainty Principle, Max Born’s theory of probability, and Bohr’s Principle of Complementarity—all emerging during that time. In 1926, Schrödinger published four successive papers that solved for a host of atomic behaviors, and included the wavefunction equation that laid out the principles of his wave mechanical interpretation. Schrödinger’s wave mechanics emphasized the deterministic, wave-like continuity of atomic systems and thus maintained a relation to classical physics, while at the same time accounting for a wide range of quantum behavior. Among its many other contributions, Schrödinger’s equation effectively predicted the interference wave pattern that unobserved particles displayed during the double-slit experiment. Because Schrödinger’s equation retained the classical wave concept, it proved to be more conventionally intuitive and “visualizable,” and thus gained wide appeal among physicists. Meanwhile, in July of 1925, Heisenberg had sent Born a copy of his paper “On the Perceptual Content of Quantum Theoretical Kinematics and Mechanics.” Born and Pascual Jordan showed that Heisenberg’s results could be described in terms of mathematical matrices and the first coherent draft of the quantum interpretation was created.2 In 1926, Heisenberg used the matrix equation to demonstrate that, in any attempt to discern certain conjugate states describing the physical properties of a subatomic “particle” (for example, position and momentum), there exists a fundamental limitation to the accuracy with which these properties can measured, and therefore known, simultaneously. Heisenberg named this limitation the “Uncertainty Principle.” Heisenberg’s Uncertainty Principle described the consequences arising from the basic fact that the quantum world cannot be perceived directly, but only through the use of measuring instruments that

6

Introduction

inevitably interfere with the results. If one were to bombard an electron with a beam of light of a sufficiently short wavelength to determine accurately its position, energy would be unavoidably transferred to the electron, thus giving it a “kick” and making it impossible to discern with any accuracy its momentum. A fundamental condition of ascertaining the electron’s position, then, is that we disturb it in an incalculable way, and are thus prevented from ascertaining its speed and direction. Conversely, measuring an electron with a long wavelength of light avoids disturbing its momentum, but the longer wavelength is not focused enough to identify the electron’s position with any degree of accuracy. The more precisely one knows the position of a particle, then, the less precisely one knows its momentum, and vice versa. This, therefore, is the limitation described by the Uncertainty Principle: nothing is or can be known with any specificity about the state of a subatomic particle prior to the act of measurement; however, there exists no mode of measurement that can accurately reveal its state.This limitation cannot be attributed to a simple lack of technical innovation; it cannot be entirely resolved with less disruptive or more precise measuring instruments.3 Knowledge of both the position and the momentum of a subatomic particle is a fundamental impossibility. Unlike the wave-based formulation of Schrödinger, which retained the notions of determinism and continuity associated with classical physics, Heisenberg’s Uncertainty Principle was essentially a particle-based model that emphasized the discontinuous, nondeterministic properties of atomic matter. Soon after Heisenberg published his paper on matrix mechanics, Schrödinger demonstrated that the matrix equation and the wavefunction equation were mathematically identical. In the same paper, he argued that his model was superior because of its greater visualizability and intuitiveness—a claim that Heisenberg, Born, and Wolfgang Pauli dismissed, even as they were undertaking efforts to make their own theory more visualizable. If Heisenberg’s formulation favored particulate behavior, and Schrödinger’s model favored wave-like behavior, Niels Bohr’s approach attempted to account for both. In September of 1927, after receiving a letter from Heisenberg about his Uncertainty Principle, Bohr became convinced that the Uncertainty Principle was merely one manifestation of a deeper complementarity between particle and wave behavior. Bohr’s

Introduction

7

focus was more on evidence derived from experimental set-ups than on the mathematical formulations of Schrödinger and Heisenberg, particularly where the presence or absence of measurement determined which property—wave-like or particle-like—was demonstrated. Bohr’s focus on experimental results no doubt influenced his essentially pragmatic approach to this duality, as well as his concentration on what he believed could be communicated in language. Like Heisenberg and Schrödinger, Bohr was sensitive to the fact that the concepts of wave and particle were drawn from classical physics, where it was not possible for something to be both a wave and a particle at the same time. Rather than rejecting the classical terms, however, Bohr concluded that, since both the “wave” and “particle” experiments produced equally valid results, and since these outcomes were mutually exclusive, it was necessary to describe the two in complementary fashion in order to arrive at a full description of matter. The origin of Bohr’s Principle of Complementarity is almost always associated exclusively with wave/particle duality; however, to reduce Bohr’s understanding of the notion of complementarity to this single duality is misleading, and disregards the subtle evolution in Bohr’s thinking about the relationship of quantum and classical physics. In fact, the wave/particle relation was one of the lesser dualities with which he was concerned. Over time, Bohr used the word “complementarity” to refer to all manner of paired relationships that he felt expressed reciprocal relations, so that it came to refer to complementary theoretical models, complementary atomic behavior, perceiving subject versus perceived object, and related experimental arrangements.4 Complementarity cannot be reduced to wave/particle duality because it includes the entire “quantum situation,” so that “complementarity” referred to a broad range of overlapping ontological, epistemological, and experiential concerns that occupied Bohr’s thinking. A number of other key concepts formed the Copenhagen Interpretation or the “Copenhagen-Gottingen Interpretation,” so named because the principal players were working primarily either at Bohr’s Institute in Copenhagen or at the University of Göttingen.5 One such concept was Max Born’s theory of probability, which reinterpreted Schrödinger’s “physical” waves as waves of probability—in other words, the probability of finding a particle in a particular state at a particular time. Schrödinger

8

Introduction

was distressed by this reinterpretation of his wavefunction, and Born’s theory contributed to the tension between Schrödinger and those involved in the creation of the quantum interpretation. Also contributing to the tension was the reintroduction of a concept that Bohr had developed earlier—that of quantum jumps. Quantum jumps described the transition between one electron state and another, but could not account for what went on between the two states, a fact that led Schrödinger to describe it as incoherent. Derived first as a mathematical solution to a number of questions surrounding Schrödinger’s wave equation was the concept of “superposition.” According to the theory of superposition, prior to detection by a measuring instrument, a subatomic system can exist simultaneously in all theoretically possible states or configurations. Upon detection, however, the system reduces to a single state, an outcome referred to as the “wavefunction collapse.” Unobserved, the electron behaves according to the wavefunction. Observed, it becomes a particle. Thus, the act of observation actually changes the nature of matter in a fundamental way. This phenomenon is known as the “observer effect.”6 In relation to superposition and wavefunction collapse, Schrödinger posed a paradoxical thought experiment known as “Schrödinger’s cat,” where a cat enclosed in a box with a radioactive nuclear sample is said to be both dead and alive until the lid of the box is opened and the cat is observed. It is commonly believed that Schrödinger developed this thought experiment to illustrate this “observer effect” at the quantum level; however, he was actually pointing out the absurdity of superposition applied on a macrocosmic scale, and, more fundamentally, suggesting the inadequacy of the quantum interpretation as a full explanation of the nature of matter.7 Both the concept of superposition and the concept of the observer effect have produced a number of ontological, epistemological, and philosophical claims. Hugh Everett proposed the foundation for the most famous interpretation of superposition in 1957, in what he called the “universal wavefunction,” which denies the wavefunction collapse and asserts the objective (nonmathematical) reality of Schrödinger’s wavefunction.8 Bryce DeWitt later popularized Everett’s theory in the 1960s and 1970s with the much more appealingly named “Many-Worlds Interpretation.”9 The Many-Worlds Interpretation also rejects the

Introduction

9

wavefunction collapse, instead positing that a superposition of all alternative histories continues to exist. According to DeWitt, everything that could possibly have happened in our past, but did not, has occurred in the past of some other universe or universes. Each potential alternative is thus realized in some other actually existing universe, and there exists a potentially infinite number of universes—a condition also known as the “multiverse.”This interpretation arose despite the fact that the founders of the quantum interpretation took pains to clarify that the interference caused by the observer effect is initiated by the detecting device, and not by the human, subjective act of perceiving the outcome—a clarification that was lost in later invocations of quantum theory in the cultural sphere. The observer effect thus has frequently been translated as an “observer/ participant” interaction, and has been used to support broad claims about the interaction between our consciousness and a “knowing” universe, as a model for participatory political structures, and as an interpretive tool describing the effect of the reader’s consciousness in the reception of literary texts. In 1935, Albert Einstein, Boris Podolsky, and Nathan Rosen wrote a paper containing a thought experiment that was designed to undermine the Uncertainty Principle and demonstrate that quantum physics was an incomplete theory. In the paper, they set out to prove that it actually was possible to measure conjugate states of particles because the measurement of one (which caused the wavefunction collapse into a single state) also caused the instantaneous collapse of the other (which then could be measured independently of the measuring apparatus).10 Instead, Einstein, Podolsky, and Rosen inadvertently generated mathematical proof that, under certain circumstances, quantum mechanics predicts a breakdown of locality.To be more specific, they showed that, in certain situations, the reduction of a property of one particle to a single state (say, a “+” spin) can cause another particle existing at a distance from the first to assume the opposite state (a “–” spin).11 Even though the particles are far apart and not physically connected in any way, they appear to act together as a single system—one particle seems to have the ability to know what the other particle is doing and to respond in a corresponding manner instantaneously (i.e., faster than the speed of light—a theoretical impossibility). The mathematical proof derived in this paper came to be known as the “EPR paradox.” Following the paper’s publication, Schrödinger wrote a

10

Introduction

letter to Einstein about the EPR paradox, or nonlocality, in which he used the term “entanglement” to describe this interaction between two apparently unconnected particles.12 EXPERIENCE, PERCEPTION, AND THE LIMITS OF LANGUAGE

Even more than Einstein’s theory of relativity, quantum physics challenged, at its very foundations, everything that had been known and accepted for centuries about the relation between cause and effect, determinism, temporal and spatial relations, scientific observation, and the reality of the material world itself. Quantum behavior defies fundamental aspects of our experience, which means that any attempt to describe this behavior in language, which derives from perceptions tied to that everyday experience, necessarily strays toward misrepresentation. Given the challenges of representing quantum phenomena—challenges that so preoccupied Schrödinger, Bohr, and Heisenberg’s observations about language—I view my initial explanation of the core concepts in quantum physics as only the first step in a series of iterations. As I move throughout this book, I return again and again to these concepts via a host of different contexts and perspectives, an iterative process wherein I circle back to them repeatedly, correcting obvious misuses, but always aware of the “not quite” that shadows my correction. In chapter 1, I uncover the association between language’s origin in ordinary sense-based experience of the physical world, and the intractability of quantum phenomena to concepts such as entity and causality, as well as to linguistic representation. I demonstrate Schrödinger, Bohr, and Heisenberg’s uniquely reflective approach to the relationship between quantum concepts, experience, perception, language, and the material world, and the extent to which the three men understood the problematic relationship between the quantum realm and the realms of perception, concepts, and language as an inevitable outcome of how each was structured. I show how, for them, quantum physics’ disruption of this linkage between sense perception and concept formation initiated a crisis of representation that undermined any attempt to communicate accurately their findings. To provide an expanded framework for why and in what terms Schrödinger, Bohr, and Heisenberg claimed that it was impossible to

Introduction

11

separate quantum phenomena from questions of perception and language, I draw from Conceptual Metaphor Theory—in particular, its core proposition that there exists a clear correlation between the kinds of concepts human beings are capable of forming and the fact that our “primary metaphors” are acquired as both a natural and culturally determined function of the way the human body interacts with the material environment.13 I focus especially on how language and metaphor derive from our sense/perception-based experience of ourselves as embodied and bounded, with particular spatial orientations, and how we use this understanding to formulate our ideas and our descriptions of objects, which we similarly construct as fixed, delimited entities that are oriented in space. My initial sense that Conceptual Metaphor Theory might prove to be a productive model for my analysis emerged from the striking manner in which this model is prefigured in the terms used by Schrödinger, Bohr, and Heisenberg to articulate the problems surrounding the quantum interpretation—including their concern with how the quantum object fundamentally disrupts our spatial orientation, connected as it is to the sort of bodies that we have and how we interact with our physical environment. The fact that our basic “orientation” to the material world is spatial, with clearly defined boundaries between things, and the fact that we experience the passing of time as continuous motion, cannot be accommodated by quantum theory. My conviction that Conceptual Metaphor Theory provides the best model for my purposes also emerges from how effectively it illuminates key representational and linguistic problems posed by the quantum interpretation. These problems include the fact that quantum phenomena exist only as probabilities or tendencies toward being; that they are essentially discontinuous and cannot be said ever to exist at a particular time and place; and that they transform under the act of observation. When viewed through the lens of Conceptual Metaphor Theory, these sorts of problems—how and why the quantum interpretation violates fundamental aspects of our experience—are brought to the fore in a manner that highlights the problems with language that so preoccupied Schrödinger, Bohr, and Heisenberg. I find Conceptual Metaphor Theory persuasive in its construction of the relationship between embodiment and our conventional perception/ representation of the external world, and very effective in highlighting

12

Introduction

the issues introduced by Schrödinger, Bohr, Heisenberg, and their colleagues. I am also aware that other theories of language, embodiment, and experience—most obviously, poststructuralism and more generally postmodernism—align with how quantum physics problematizes language as an expression of experience, as well as the notion of a bounded self. Some have argued (as I will discuss in chapter 3) that quantum physics is an expression of poststructuralism’s recognition that language is selfreflexive, and of deconstructionism’s undermining of being and presence. One might also argue that postmodernism’s refiguring of the body and identity as permeable and contingent aligns with the way quantum physics undermines the concept of self-consistent “things.”While these theories provoke essential discussions on the nature of language and the subject, the fact remains that in our ordinary, everyday experience of the world, and in our language, the majority of us remain tied to the notion of ourselves as separate and discrete entities. Certainly, the primary interpreters of quantum physics saw the problem of quantum physics in these terms, and it had far-reaching implications for them. Heisenberg, Schrödinger, and Bohr were highly sensitive to the fact that even assigning a name to a quantum phenomenon was already to impose habitual notions onto utterly novel circumstances, and Bohr spoke for both Heisenberg and Schrödinger when he observed that conventional terminology had to be retained “because you haven’t got anything else.” The necessity of using conventional terms such as “wave,” “particle,” “complementary,” “observer,” “uncertainty,” and “entanglement” introduced into quantum concepts what might be called an “originary drift”—originary, because the drift emerges simultaneously with the naming of the concept.While I acknowledge that all language mediates and to some extent constructs reality, I maintain that quantum physics is exceptional in the degree to which the behavior of quantum phenomena, which can be expressed mathematically with a high degree of accuracy and specificity, are especially subject to misrepresentation once they enter language. I draw from Liliane Papin’s argument that Indo-European languages can only poorly represent quantum physics because these languages are based broadly on a structural division between nouns, associated with “thingness” (or entities), and verbs, which denote actions or states of being. This structural bias in language, I argue, means that quantum terminology is not merely a screen that reflects or deflects

Introduction

13

our view of the quantum world; rather, the structure of language precludes the accurate representation of quantum phenomena. Particularly in their later efforts to appeal to a wider audience, Schrödinger, Bohr, and Heisenberg introduced extended conceits that went beyond describing quantum phenomena as entities and cast them in specifically human terms. I explore their (sometimes hesitant and reluctant) rhetorical strategy of using anthropomorphism to describe subatomic behavior, examining the implications of metaphorical conceits that include references to electrons as military formations and probabilities in terms of the collective versus the individual. I conclude chapter 1 with an exploration of how they used metaphors of “journey” and “architecture” to describe the relationship between classical and quantum physics, the role of the scientists, and the scientific project itself. THE PHYSICS OF VISUALITY, INTUITION, AND AESTHETICS

In chapter 2, I demonstrate how Schrödinger, Bohr, and Heisenberg’s respective understanding of what could and could not be said about the quantum world both emerged from and defined each man’s approach to the complex relationship between visuality, intuition, and aesthetics. In the first section I offer an account of the “visualizability (anschaulichkeit) debate” that went on primarily between Schrödinger, who argued in favor of a visualizable model of the material world, and Heisenberg, who defended his abstract, algebraic, and highly “unvisualizable” quantum matrix interpretation. I reject, however, the common assumption that Heisenberg was entirely against a visualizable model, and I show how, despite his public repudiation of the “visualizability imperative,” Heisenberg was in fact preoccupied with how far the quantum interpretation fell short of the demands of visualizability, and how his, Max Born’s, and Wolfgang Pauli’s revisions to the quantum interpretation—including the introduction of quantum jumps and of the Uncertainty Principle for which Heisenberg is most remembered—were in part inspired by his desire to make his model more visualizable. I describe how the struggle over the meaning and value of what constituted an “intuitive” model of the material world generated a similarly heated debate within the atomic physics community. Examining, in the context of the quantum interpretation, the different commentaries and

14

Introduction

rhetorical strategies around the concept of intuition (anschauung), I offer an interpretation of the relationship between the physical world and human understanding, and of what a debate over the proper meaning and use of a concept seemingly unrelated to science can reveal about historical struggles over which scientific theory will prevail. After contextualizing how the concept of intuition arises in relation to the quantum interpretation, I examine Schrödinger’s insistence that any intuitive theory of matter must remain intelligible in future scientific paradigms, as well as the rhetorical strategies that he used to make matrix mechanics appear “unscientific.” I proceed to analyze how Bohr’s “symbolic turn” represents a different relationship to intuition, particularly in the emphasis that he places on the description of the entire quantum situation, which includes experimental arrangements, experimental evidence, observation, and classical concepts. I argue that Bohr sees the classical concepts as symbolic idealizations, rather than as conduits for our absorption and processing of data. To the extent that, in Bohr’s model, we have access to concepts such as wave and particle, then, we can apply a customary form of intuition, but to the extent that these concepts become a function of description rather than reality, we must endeavor to reach beyond our usual ways of knowing. I follow with an account of Heisenberg’s attempt to break entirely with the customary definition of intuition, and I trace the complex logic that Heisenberg pursues in his attempt to rid the concept of intuition entirely of its conventional associations and to redefine it in a manner that would put to rest accusations that the abstract algebra of his matrix mechanics was hopelessly unintuitive.Through close readings of Heisenberg’s writings and interviews from the 1920s to the 1960s, I uncover the rhetorical strategies that he used in an attempt to establish the fact that the unintuitive aspects of his model reflect merely the limitations of the linguistic resources and modes of communication that are currently available. I then analyze how the implicit and explicit aesthetic preferences of Bohr, Heisenberg, and Schrödinger complement the formal aspects of their atomic theories. I show how, where it concerns their aesthetic sensibilities, there is a good deal of accord among Schrödinger, Bohr, and Heisenberg, particularly with respect to two notions: harmony and simplicity. I trace how all three men privilege harmony in particular, and

Introduction

15

how both Bohr and Heisenberg find harmony through “unity in multiplicity”—Heisenberg, in the way that the quantum interpretation imposes order on a set of confusing, because seemingly irreconcilable, quantum phenomena. I argue that Bohr also stresses unification, which is a driving force for how he approaches the natural world, starting with the symbolic unity of classical concepts that defines his Principle of Complementarity, and extending to his belief in the shared interests of fields of knowledge other than his own. I connect Schrödinger’s belief that one must only describe those aspects of the material world that can be recorded to his acceptance of the “blank spaces” of his contemporaneous art. I then demonstrate how Bohr’s complementarity represents symbolically-oriented classical concepts of harmony and reciprocity. Finally, I show how, drawing from the Platonic tradition, Heisenberg locates the foundation of aesthetics in the symmetry, harmony, and truth of mathematics, and that his aesthetics represent a logical extension of Bohr’s. For Heisenberg, I argue, simplicity lies in the purity of mathematical forms, for Bohr in the fundamentals of description that regardless of the theory or context must take on classical form, and for Schrödinger in the spare expression of the object’s core characteristics. In some respects, despite the novelty of their theories, all three men are classical in their sensibilities: each one seeks out the one in the many, the unity in difference, and the essential in, as Heisenberg puts it, “the plethora of details.” Following my analysis in chapters 1 and 2 of how quantum concepts led Schrödinger, Bohr, and Heisenberg to reflect on experience and experiment, perception and observation, metaphor and communication, I investigate when and how the “productive imprecision” of quantum terminology resurfaced to provide an interpretive framework for a broad set of agendas, sensibilities, practices, and worldviews. Because quantum phenomena cannot be rooted in any particular experience, they can take root, albeit imperfectly and temporarily, in any experience. Because quantum phenomena have such an oblique relationship both to language and experience, the terminology used to describe them is exceptionally nomadic, and can be applied to support everything from novel methodologies in literary criticism, to new and post-New Age associations of the quantum with Eastern mysticism, to reconceptualizations of global politics. As I follow this “quantum trail” in chapters 3 and 4, I analyze

16

Introduction

commonalities and differences in the way each new context appropriates, refigures, and redeploys quantum concepts such as complementarity, entanglement, uncertainty, and the observer effect on a macrocosmic scale. Reflecting on the precise form of each new context’s deployment and strategic distortion of quantum concepts allows me at once to loop back and further clarify the original quantum concepts, and to uncover the social, political, economic, and cultural needs and expectations, as well as the various constellations of hopes and anxieties that quantum concepts are conscripted to serve. THE LITERARY TURN

In chapter 3, I turn to the emergence in the 1980s of quantum theory as an analytic tool in literary criticism, when the rigorous and complex methodologies and metaphysics of poststructuralism dominated. The principles of quantum theory, I argue, added credence to a literary criticism that was influenced by poststructuralism’s rejection of Western logos and subject/object duality, the decentering of author and authority, and the disruption of the relationship between signifier and signified. With the emergence in the 1990s of programs in cultural studies, women’s studies, and postcolonial studies, and given the key role that English departments played in the establishment of these programs, applying quantum physics to literary analysis served a critical culture that sought out opportunities for interdisciplinary cross-fertilization. At the same time, using quantum concepts granted “scientific” legitimacy and theoretical sophistication to literary studies and, more importantly, offered a solution to the increasing marginalization of the humanities and the growing economic clout of the sciences. While often associated with poststructuralist methodologies and occasionally with postmodern texts, quantum criticism’s main literary object tends to be modernist literature, with its stream of consciousness, temporal play, challenges to linear logic, and reflections on the limits of knowledge. In this chapter, I analyze how wave/particle duality is enlisted to describe (and essentialize) gender differences in Virginia Woolf ’s novels, and how Bohr’s Principle of Complementarity gets linked to poetic form—an association that I attribute to poetry’s tendency toward local juxtapositions and imagistic language, which encourages the reader to

Introduction

17

alternate between more than one perspective. I follow how the observer effect is introduced to explore the relationships between author, reader, and narrator; how the concept of superposition is used both to support claims about the loose or multifaceted nature of interpretation, and to argue that narratives remain undefined until interpreted by a reader; how the word “uncertainty” in the Uncertainty Principle is reduced to its conventional meaning to support claims about textual indeterminacy and fragmented, open-ended narratives; and finally, how entanglement is used to examine how characters affect one another, even when they are separated in space or time. I identify where, how, and in the service of what interpretive agendas quantum concepts are misapplied, and uncover the specific distortions that this introduces into both the literary text and a particular quantum concept, while at the same time I identify where and how these concepts are applied to good effect. I begin by exploring the advantages and disadvantages of claiming a direct causal relation between the birth of quantum physics and the narrative form, characterization, and thematics of particular modernist authors. I then examine the advantages and disadvantages of the less deterministic notion that at any given historical moment, there exists a circulation of ideas and a general intellectual climate that fosters similar logics across disciplines and cultural formations. Through an analysis of specific critics, I demonstrate that, despite these claims of direct influence or cross-fertilization, one is often hard pressed to find any necessary relationship between quantum concepts and the metaphorical use of them to analyze texts. In many cases, other theoretical models seem to offer equally or more fertile ground for interpretation. More problematic than this, however, are the cases where quantum concepts are used in a misleading way that undermines efforts in the humanities to stake a claim within science studies. Of all of the quantum postulates, the concept of wave/particle duality has proven to be the most popular tool for literary analysis of modernist texts. I examine how the concept emerges in particular around character analysis, usually in the claim that some characters are “particle-like” and others are “wave-like.” I explore why Virginia Woolf ’s novels so overwhelmingly dominate as objects of an analysis based on wave/particle duality, and I situate this phenomenon in the context of gender stereotypes that associate women with waves and men with particles. I also

18

Introduction

critique the use of superposition and wavefunction collapse/observer effect to analyze subject/object relationships, the breakdown of the boundary between subject and object, and the transformation of the object by the perceiving subject. I track how references to the observer effect prove particularly popular where critics are discussing the relationship between author, reader, and narrator. Critics may show, for example, how the narrator “interferes” with the narrative form or with the characters, in the same way that the measuring device is inextricably a part of the observed outcome. I also show how references to superposition—that moment prior to interference where all states exist simultaneously—are used to support claims about the loose or multifaceted nature of interpretation, wherein the reader’s mind ranges freely over the text, never settling on one particular interpretation. In opposition to this, some critics argue that the text/narrative exists in a superposition of states until “observed” (interpreted) by the reader; in this model, the act of reading “collapses” the potentialities of meanings contained in the text. While not as popular as wave/particle duality, the Uncertainty Principle provides one of the richer metaphors for critics interested in narrative form. I show how critics invoke the Uncertainty Principle, for example, to launch claims about textual indeterminateness or ambiguity, or with reference to open-ended, incomplete, or fragmented narratives that achieve no final closure. Not surprisingly, given the kinds of textual dynamics in which they are invested, critics tend to associate the Uncertainty Principle with postmodern rather than modernist literature. In the way that the term is typically used, however, it is often unclear why an appeal to quantum physics is necessary in the first place, and why, for example, a deconstructive analysis—with its privileging of undecidability and epistemological uncertainty—is not sufficient to the task. At the same time, as I show, some critics use the Uncertainty Principle productively, most often when they combine their analysis with an appreciation of the problem of the relationship between language and quantum physics that Schrödinger, Bohr, and Heisenberg confronted, and connect this problem to the inherent metaphoricity of language. I consider entanglement/nonlocality for the way in which it is typically used to examine how characters affect one another. I demonstrate two major misapplications of entanglement: the first associates it with proximity or direct interaction, so that entanglement becomes a kind

Introduction

19

of closeness or intermingling that is characterized by the loss of boundaries. The second uses entanglement to describe a to-and-fro, mutually affecting relationship between characters. While this second approach retains the quantum notion of communication at a distance, I argue, the focus on mutually affecting relationships leads critics to efface the deterministic, one-way causal nature of quantum entanglement. In general, critics misrepresent entanglement as a relationship between self, object, and world, in an effort to apply quantum terminology to what amounts to a situation wherein characters are in some way “tangled up” with one another. I conclude that to avoid the pitfalls of using quantum physics as a model for literary criticism, critics must perform due diligence in understanding quantum concepts, and exercise care in applying quantum metaphors to literary works. They must assess whether quantum concepts are really necessary to their interpretation, which means understanding the literary text with the same thoroughness that they understand the scientific concepts. They must focus on specific dynamics within the text, select the appropriate quantum concept in all its specificity, and avoid sweeping claims about “quantum” critical methodologies or entire categories of literature. If these basic conditions are met, I argue, and if the critic integrates literary analysis with quantum physics in a dynamic and productive manner, the use of quantum concepts can both open up new perspectives on a given literary text, and to expand our understanding and appreciation of quantum theory and its epistemological and linguistic implications. NEW AND POST-NEW AGE APPROPRIATIONS

I begin chapter 4 with a discussion of how atomic theories in general, and quantum theory in particular, have provided material touchstones for the expression of existing or emerging political tendencies and ideals, including the “participatory democracy” said to inhere in quantum logic. I consider the burgeoning field of quantum politics, which defines itself against the liberal individualism of Newton’s era, and attempts to generate a new language that better reflects the global complexities of our time. Advocates of quantum politics include political philosophers or political scientists who, in the spirit of the seventeenth-century

20

Introduction

philosophers who turned to Newton’s atomism to provide founding models for liberal individualism, look to quantum phenomena to provide metaphors for imagining how to resolve social fragmentation, alienation, and cultural conflict. I trace how advocates of quantum politics tend to use the notion of “observer/participant” as a metaphor for a more radical form of democracy or more genuine intercultural engagement, and the “both/and” logic of the Principle of Complementarity to model a culture of negotiation and mutual recognition. My analysis of quantum politics considers the following questions: what conceptual consistencies link the various attempts to import quantum metaphors into the arena of politics? How successful are these models in offering a new form of political engagement? To what extent are they able to envision concrete applications of “quantum politics” to the specific socio-political challenges that we face? In their rejection of conventional medical and analytic discourse and their embracing of Eastern philosophies and practices, few New Age adherents saw value in existing scientific paradigms. The exception, it turned out, could be found in two authors who attempted to integrate Eastern philosophy with the obscure and counter-logical world of quantum physics: Fritjof Capra and Gary Zukav. In this chapter, I examine New Age appropriations of the quantum interpretation as a paradigm for understanding Eastern mysticism. In the course of examining the “quantum mysticism” of Capra and Zukav, I demonstrate how, for example, Capra recasts the observer effect as an interactive union between the material world and the consciousness of the observer. I argue that both Capra’s and Zukav’s descriptions of quantum concepts are extensive and well informed; however, they have a tendency to abandon their attention to scientific accuracy and detail whenever they want to establish a connection between quantum physics, human experience, and Eastern mysticism. In an analysis of Capra’s and Zukav’s rhetorical strategies, I show how, in order to align quantum physics with the insights of Eastern mysticism, Capra and Zukav reinvent the language of quantum physics and sometimes engage in metonymic slippages or apply faulty logic as they collapse the distinction between microscopic and macroscopic behavior. I then turn to the post-New Age quantum healing practitioners and quantum “get-rich” gurus who emerged in the 1980s, counterposing

Introduction

21

these practitioners to New Age quantum philosophy, and demonstrating how the former tend to be individualistic, and to market quantum concepts as aids for personal gain. Focusing primarily on Internet-based quantum healing and quantum enrichment sites, I show how these guided self-help programs invoke complementarity, the observer effect, the Uncertainty Principle, and entanglement to lend scientific legitimacy to wide-ranging claims about the participatory interrelation between consciousness and cosmos, subject and world, mind and body. In the process, I argue, they reconfigure quantum particles into vivified agents whose unique movements and interactions, properly directed, offer a methodology to secure health, happiness, and wealth to consumers. Along with teasing out the manner in which new and post-New Age philosophers, practitioners, and vendors apply quantum concepts to support their claims, I trace the increasing commodification of the quantum, which extends into areas of subjectivity—for example, spirituality—that historically have been considered immune to overt commercialization. This extension of the commodification process is evidenced in the way that quantum methodologies are commercialized and then sold to people as a means of advancing not just their financial interests, but their spiritual well-being as well. I argue that the new economy of the subatomic particles, with the goal of withdrawing into “self-care,” emerges in part from the late-twentieth and early twenty-first century atrophy of the public sphere as a site of interpersonal engagement, just as it reflects a deep contemporary sense of being unmoored that fuels the search for structured guidance as a means toward a renewed sense of control over one’s life. Throughout chapter 4 I demonstrate how the language of scientific advancement in general, and quantum physics in particular, can be adapted to reflect particular social, political, and cultural needs and expectations, as well as specific constellations of hopes and anxieties. From Newton’s liberal individualist atomism to the participatory democracy of quantum politics, and finally to the contemporary commodified and personalized atom of quantum healing and quantum enrichment programs, atomic theories provide a material touchstone for existing societal tendencies. In this sense, models of the atom and its component parts have reflected worldviews just as much as they have reflected scientific

22

Introduction

advances, with existing societal priorities driving the formulation and application of these advances. SPLITTING THE ATOM

In my fifth and final chapter, I use quantum discourse as a backdrop for my analysis of the “nuclear discourse” that began with the first atomic test blast. I argue that the discovery of how to release the tremendous power latent in subatomic particles culminated in a spectacle so massive and so laden with unthinkable implications that it, like quantum physics, defied conventional description. In some respects, quantum physics and the nuclear physics on which I focus in chapter 5 are very different: the quantum phenomena such as indeterminacy, complementarity, the observer effect, and superposition are not strictly relevant in nuclear physics.These phenomena do not play a part in my analysis of nuclearism, and my comparison does not surround the dynamics of particles and waves in the subatomic realm. Instead, I concentrate on the kinds of linguistic and cultural concerns that emerged from and have adhered to interpretations of the quantum world. I base my comparison on the proposition that, in their similarities, these two atomic discourses suggest more universal tendencies: in the problems that they pose for language, the tendency to express the ineffable in mystical and religious terms, the tendency to domesticate and make familiar the inexplicable by developing humanizing, anthropic narratives, and the tendency to use our relation to the material world as a vehicle for articulating social relations. In their differences, I propose, quantum and nuclear discourses call for what might be termed a “complementary” approach: a full understanding of the linkages between human experience, perception, language, and the material world requires moving back and forth between the two. Like quantum phenomena, I argue, the initial blast could not be made operational in narrative because it too lacked an experiential point of reference. If quantum phenomena are inexpressible because they contradict ordinary perception, the atomic blast was inconceivable because it overwhelmed ordinary perception. In other words, the size and overwhelming spectacle of a nuclear explosion threatened to make mute any discursive attempt to establish a connection to individual experience. Like the founders of quantum physics, those who sought to describe the

Introduction

23

nuclear blast reflected in a self-conscious manner upon the poverty of language to communicate what they had witnessed. In the case of nuclearism, however, reflection on the poverty of language with respect to the nuclear blast became a rhetorical strategy in itself, manifested most obviously in the way witnesses turned to the language of the sublime to describe what they experienced in that New Mexico desert. By examining the rhetorical strategies used in speeches, newspapers, pamphlets, civil defense videos, popular culture, and images produced from 1945 to 1955, I demonstrate nuclear discourse’s similarities to and differences from quantum discourse: in, for example, the relationship that it expresses between sense-based experience, language, and the material world; in its relationship to visuality; in its construction of the relationship between subject and object, observer and observed; and in its dominant metaphorical conceits. As I trace the maturation of nuclear discourse, I show how the fundamental distinction between nuclear discourse and quantum discourse lies in nuclearism’s rhetoric of mastery, which can be traced back to the atomic phenomenon that it describes. Opposed to this focus on mastery is how quantum physics is about giving up control over the material world in a number of ways: in the fact that it is not possible to know two aspects (e.g., position and momentum) of the same system simultaneously; in superposition and the observer effect, which prevent us from knowing the state of a subatomic system prior to measurement; and in the “spooky action at a distance” of quantum entanglement. If quantum phenomena are about nature’s power to defeat human efforts, I propose, the splitting of the atom is about humanity’s faith in its power to manipulate nature toward a defined end. Rather than indirect observation of the subatomic world, nuclear physics engages in direct and aggressive action upon it, violently rending the core of the atom through unnatural means. Splitting the atom is not about participating in cosmic forces; it is about taming them. It is not about giving up control to nature; it is about wresting control from nature and bringing it under the command of humanity. The difference between quantum and nuclear discourse, I argue, is also manifested in their respective relationships to the religious and mythic registers. As it moved outward from its origins, for example, quantum physics accrued mystical associations, most of which posited a participatory relationship between humanity and the material world. The mythology of nuclearism, on the other hand, was fixed on a

24

Introduction

divine sense of mastery, of ultimate control over the material world and the promise of immortality offered by peacetime atomic energy. The bomb was not about sharing power with the cosmic forces of the universe; it was about the precise control of those forces. Throughout this chapter, I investigate how the eventual rhetorical drift of atomic discourse differed from that following the introduction of quantum physics.While quantum discourse was apolitical, nuclear rhetoric was explicitly political and collectivity-oriented, either toward humanity in general or American citizens in particular. The unfolding of the quantum world took place within a highly constrained arena, among the very few numbers of people who could even begin to understand, let alone generate, the complex mathematics that described it. The later use of the quantum in popular culture and literary criticism tended to be abstract, and either applied to the individual or to support lone efforts at textual exegesis. On the one hand, where the quantum was politicized— for example, in quantum politics or in critiques of Western logos and dominant forms of subjectivity—the outcomes remained academic and speculative.The result of the splitting of the atom, on the other hand, was public, highly politicized, and global in its implications. Much of the rhetoric around the bomb, then, was similarly politicized and public-oriented. I show how the bomb and its use were managed, recast, and domesticated for the public through rhetorical strategies such as misdirection, euphemism, anthropomorphism, associations with benign popular culture figures and imagery, mythic and religious references, and the carefully crafted and controlled release of images. If quantum discourse stood in the place of a vanishing reality, I argue, nuclear discourse was designed to make reality vanish and to replace it with a convenient fantasy. While in chapter 1 I trace how quantum phenomena were resistant to visualization, in chapter 5 I show that in the case of nuclearism, the visual plays a central role—in the accrual of metaphors, symbols, and tropes, but also in imagery surrounding the bomb itself and the people involved in deploying it. I show how images associated with the bomb suggested that the closer America was to its deployment at Hiroshima and Nagasaki, the greater the need was for detaching from it.This detachment was achieved in part through a connotative rhetorical strategy manifested in the mythic imagery of what I call “nuclear celebrity,” which emerged directly before and for a brief time after the bombs were dropped, and

Introduction

25

through which the bomb’s most immediate and disturbing consequences were managed. From arguments that our mind can act on the world and change our reality, to claims about quantum healing and quantum enrichment, to analyses of narrative, the concepts of uncertainty, complementarity, superposition, and the observer effect have strayed far from their initial context. These diverse mobilizations of quantum concepts are artifacts of their nomadic quality—a kind of plasticity derived from fact that they are not embedded in familiar experience. As a result, these concepts could be appropriated in order to serve a variety of agendas and worldviews. More generally, quantum and nuclear discourse—and the relationship between the two—are so instructive because they occupy a “threshold position” beyond which lies the inexpressible. It is inevitably at the threshold of experience, language, and thought—with “threshold” as both a point of origin and an outside limit—that one discovers the entwined potentialities and limitations that most essentially define both forms of discourse.

1

EXPERIENCE, PERCEPTION, AND THE LIMITS OF LANGUAGE

INTRODUCTION

Scientific work in physics consists in asking questions about nature in the language we possess. —Werner Heisenberg

Historically, Western science has tended to operate under the assumption that language can convey scientific discoveries in an unproblematically referential manner. As Charles Bazerman observes, “over the past centuries, several forces have tended to suppress our consciousness of the rhetorical, communicative and symbolic character of scientific knowledge—thereby suppressing awareness of the role of language in the production of knowledge.”1 The suppression of rhetoric in science, Bazerman argues, is because scientists want to remove or repress “misleading” forms of representation, and to communicate direct empirical knowledge of the material object.2 James Edie similarly writes that “the Western promise of science has been an impersonal, mechanical, third person, deanthropomorphized explanation of reality … a science (logos) [that] seems to liberate us from metaphor.”3 Finally, Liliane Papin writes, “the development of positivist science has been accompanied by a subliminal dependence on a neutral scientific discourse, a sort of ‘degree zero’ of language, as opposed to the highly unreliable language of literature, which can indulge in metaphors and embroidery and can therefore never escape subjectivity, the anathema of scientific endeavor.”4 According to these critics, within the scientific community the use of figurative language is viewed as contaminating, causing meaning to stray and sacrificing truth to distorting flourishes that call into question the

28

Chapter 1

legitimacy of science. Scientists fail to acknowledge the materiality of language and the way it mediates between scientific discovery and understanding. There exists one striking exception to scientists’ apparent faith in the capacity of language to describe the natural world: those who sought to communicate the curious world of quantum physics. In this chapter, I examine how the key players in the development of and disagreements over the quantum interpretation of matter—primarily Erwin Schrödinger, Niels Bohr, and Werner Heisenberg—confronted the problem of communicating the core concepts of a science that disrupts our sense of space and time and, more generally, of being and doing. Existing by way of discontinuities, probabilities, mutually exclusive or multiple simultaneous states, quantum phenomena can only ever be said to be incompletely and intermittently “there.” As a result, the quantum world challenges our conventional assumption that nature consists of discrete, coherent substances that persist over time, and this challenge in turn strikes at our assumptions about the reliability of language. Against the scientists who would ignore the interference of language, those who were involved in the development of quantum theory found that their very object of study forced them to be highly conscious and overtly reflective about the language that they used. They spoke and wrote a great deal about the relationship between their experimental findings, the nature of observation, and particularly the structural limitations of language with respect to quantum physics. Their exceptional attentiveness to questions of representation means that exploring the metaphors they used and the implicit rhetorical strategies they undertook can only tell a small part of their story; to uncover the rest of this story requires exploring the linguistic, epistemological, and philosophical insights that they arrived at during, as well as long after, quantum physics evolved into a coherent theory of matter. First, I draw from the central premises of Conceptual Metaphor Theory in order to demonstrate why quantum atomic behavior is so intractable to conceptualization. Beginning with the work of George Lakoff, Mark Johnson, and Mark Turner, and drawing from the work of Penny Tompkins and James Lawley,Vyvyan Evans, and Liliane Papin, I summarize Conceptual Metaphor Theory’s proposition that embodied sense perception and everyday experience are the ground for “image schemas”

Experience, Perception, and the Limits of Language

29

such as UP-DOWN, OBJECT, and PATH that structure our concepts and our metaphors. At the same time, I demonstrate the striking parallels between how Schrödinger, Bohr, and Heisenberg envisioned our relationship to the material world, and how that relationship is constructed in Conceptual Metaphor Theory. I then turn more specifically to the problem of language. With reference to reflections on epistemology and language made by Heisenberg, Bohr, and Schrödinger, as well as work by contemporary critics in the field of science studies, I consider the “entity-problem” of quantum physics in the context of Liliane Papin’s argument that part of the linguistic challenge introduced by the “new physics” is connected to the fact that quantum behavior cannot be reconciled with noun-based Indo-European languages, since in quantum physics elementary particles “exist” only as a set of relations constituted by change, process, and events. I consider examples where boundedness, in the form of anthropomorphism, enters the later, more philosophical and public-oriented works of Schrödinger, Bohr, and Heisenberg in a manner that retains the emphasis on entity and substance, while adding intention and agency to quantum phenomena. With a particular focus on the tensions between “individual” and “collective” in the anthropic metaphors of Schrödinger and Bohr, I explore how, in order to make subatomic behavior meaningful on a human scale, they incarnate acting, knowing, and reasoning particles to create what reads as a human drama. I conclude this chapter with an analysis of the three physicists’ individual use of journey and architecture metaphors to relate a tale of scientists and science itself as central characters in a drama of discovery and creation. In discussing the journey metaphor, I draw on George Lakoff and Kai Mikkonen’s observations about how the journey metaphor spatializes time, so that time becomes a surface and a path along which the individual travels. I show how in their journey metaphors, all three men express their sometimes conflicted relationship to classical and quantum physics, and how in their architectural metaphors they depict how quantum physics both completely dismantles the bedrock of a theory that has stood for centuries, and how the classical framework exists as a limiting structure into which quantum physics must fit—however uneasily. Because the concepts of quantum physics are so far removed from everyday experience, their translation into language underscores the

30

Chapter 1

productive (as opposed to merely descriptive) capacity of metaphor. It is from this perspective that I examine the conceptual drift that occurred when the mathematical axioms of the quantum interpretation were translated into language, as well as Bohr, Schrödinger, and Heisenberg’s anxieties about the effects of this drift.While all of the primary texts that I consider in this chapter contain passages with highly technical language and difficult mathematics, my main focus is on their observations about perception, ordinary experience, and the nature of the relationship between language and the material world. In mining this subject, I show how the form and terminology of each man’s contribution to quantum physics both grew out of and contributed to his epistemological, philosophical, and linguistic concerns. CLASSICAL VERSUS QUANTUM PHYSICS: OBJECT AND CAUSALITY

All the words or concepts we use to describe ordinary physical objects such as position, velocity, color, size, and so on, become indefinite and problematic if we try to use them of elementary particles. —Werner Heisenberg

According to Conceptual Metaphor Theory, metaphors are created when we conceptualize and talk about abstract ideas in terms of more concrete ones by mapping (identifying a correspondence between) the elements of a concretely based concept, which constitutes the source domain, onto a more abstract one, which constitutes the target domain.5 This process of constructing a metaphor by mapping elements from the source domain onto the target domain allows new elements to be defined in terms that are already familiar to the users.6 Lakoff and Johnson’s most fundamental claim is that metaphor can serve as a vehicle for understanding a concept only by virtue of its connection to experience, and that “no metaphor can ever be comprehended or even adequately represented independently of its experiential basis [emphasis in original].”7 “Primary metaphors,” they argue, are acquired as both a natural and culturally determined function of the way the human body interacts with the material environment.8 Conceptual Metaphor Theory proceeds from the argument that our everyday interaction with and observation of the world around us gives

Experience, Perception, and the Limits of Language

31

rise to the image schemas I cited earlier—those recurring, dynamic, multimodal patterns of perception that emerge naturally from our everyday experience—where “experience” is understood to have basic perceptual, motor, emotional, historical, social, and linguistic dimensions.9 Image schemas organize our perceptual experience, and thus give rise to understanding—to “the way we ‘have a world,’ [and] the way we experience our world as a comprehensible reality.”10 As defined by Mark Johnson, image schemas on the one hand are derived from embodied sensory and perceptual experience; on the other hand, they structure our comprehension of that experience. In The Body in the Mind:The Bodily Basis of Meaning, Imagination, and Reason, Johnson writes, in order for us to have meaningful connected experiences that we can comprehend and reason about, there must be pattern and order to our actions, perceptions, and conceptions. A schema is a recurrent pattern, shape, and regularity in, or of, these ongoing ordering activities. These patterns emerge as meaningful structures for us chiefly at the level of our bodily movements through space, our manipulation of objects, and our perceptual interactions.11

Tayebeh Asgari elaborates: An image schema is a recurring structure within our cognitive processes which establishes patterns of understanding and reasoning. They are directly meaningful (“experiential”/ “embodied”), preconceptual structures, which arise from or are grounded in, human recurrent bodily movements through space, perceptual interactions, and ways of manipulating objects. Image schemas are highly schematic gestalts which capture the structural contours of sensory-motor experience, integrating information from multiple modalities that give rise to them in the first place.12

The recurrent structure of the image schema, then, lies at the heart of our experience, and constitutes a stable structural correlation between our situated embodiment in the world, our sense perception, and the kinds of concepts we are capable of forming. One may already see that a parallel exists between Conceptual Metaphor Theory and the manner in which those involved in the generation of quantum physics viewed the dilemma with which they were faced. In a typical passage, drawn from his 1929 article, “Introduction Survey,” Bohr writes,

32

Chapter 1 We are concerned with the recognition of physical laws that lie outside the domain of our ordinary experience and which present difficulties to our accustomed forms of perception. … In appraising this situation, however, we must not forget that, in spite of their limitations, we can by no means dispense with those forms of perception which color our whole language and in terms of which all experience must ultimately be expressed.13

In his 1938 address given at Kronborg Castle, later published as “Natural Philosophy and Human Cultures,” Bohr said that all description of experience has so far been based on the assumption, already inherent in ordinary conventions of language, that it is possible to distinguish sharply between the behavior of objects and the means of observation. This assumption is not only fully justified by all everyday experience but even constitutes the whole basis of classical physics.14

Like Lakoff, Johnson, and Asgari, Bohr identifies an inherent connection between experience, perception, and language—a connection that forms the cornerstone of concept of the image schema and what it implies—and like Lakoff and Johnson, Bohr believes that any description of the behavior of matter must use terms that come to us from ordinary experience. The difference, from Bohr’s perspective, is that the macrocosm-, perception- and experience-based image schemas that “colour our whole language” are in direct conflict with the microcosmic realm of quantum phenomena.When Lakoff and Johnson write that no metaphor can ever be comprehended or even adequately represented independently of its experiential basis, they are setting out the terms of their interpretive paradigm. When Bohr writes elsewhere that you cannot abandon the classical terms “because you haven’t got anything else,” he is expressing the terms of a dilemma.15 This dilemma would prove to be a major inspiration for his Principle of Complementarity, and one that would plague, for decades, not only Bohr, but also Heisenberg and Schrödinger. Our most basic image schemas relate to our spatial orientation, which arises from the fact that we have bodies of the sort that we have, that they function as they do in our physical environment, and that they include up-down, in-out, front-back, on-off, deep-shallow, and central-peripheral aspects. Vyvyan Evans offers the following example of how the UPDOWN image schema structures our experience:

Experience, Perception, and the Limits of Language

33

gravity ensures that unsupported objects fall to the ground; given the asymmetry of the human vertical axis, we have to stoop to pick up fallen objects, look in one direction (downward) for fallen objects, and in another (upward) for rising objects. In other words, our physiology ensures that our vertical axis, which interacts with gravity, gives rise to meaning as a result of how we interact with our environment.16

In their discussion of entity and substance metaphors, Lakoff and Johnson observe that [o]ur experience of physical objects and substances provides a further basis for understanding—one that goes beyond mere [spatial] orientation. Understanding our experiences in terms of objects and substances allows us to pick out parts of our experience and treat them as discrete entities or substances of a uniform kind. … When things are not clearly discrete or bounded, we still categorize them as such. … Human purposes typically require us to impose artificial boundaries that make physical phenomena discrete just as we are: entities bounded by a surface.17

No spatial orientation is possible without interaction with objects around us, even if that interaction consists only of looking for/at an object, and for this reason, the notion of entities is a central structuring aspect of our understanding, of “the way we are meaningfully situated in our world.”18 The relationship of entities to embodied experience and sense perception also plays a central role in the discourse around the relationship between classical and quantum physics, for this relationship is central to how we apprehend the nature of the material world. The importance of how our bodies interact with the world, and how our senses reveal the nature of this world, informed Newton’s empiricist belief that reality can be known directly through the senses. In the following passage, Newton projects onto the microscopic realm what our embodied experience and sense perception tell us about the macroscopic world: We no other way know the extension of bodies than by our senses, nor do these reach it in all bodies, but because we perceive extension in all that are sensible, therefore, we ascribe it universally to all others also. That abundance of bodies are hard, we learn by experience, and because the hardness of the whole arises from the hardness of the parts, we, therefore, justly infer the hardness of the undivided particles not only of the bodies we feel but of all others.

34

Chapter 1 That all bodies are impenetrable, we gather not from reason, but from sensation.The bodies which we handle we find impenetrable, and thence, conclude impenetrability to be an universal property of all bodies whatsoever. That all bodies are moveable, and endowed with certain powers (which we call … inertia) of persevering in their motion, or in their rest, we only infer from the like properties observed in the bodies which we have seen.19

In this passage, Newton echoes one of the basic suppositions of Conceptual Metaphor Theory: that “we know the extension of our bodies from our senses”—and the fact that “all bodies are impenetrable, we gather not from reason, but from sensation.” Newton’s projection of this causal relation onto the microcosmic realm follows from his assertion that “because we perceive extension in all that are sensible, therefore, we ascribe it universally to all others also”—in other words, to that which, being too small, we cannot perceive through our senses. Newton’s atomic particles possess a definite shape and occupy a particular bounded region of space: they have extension, they (theoretically) can be handled, and they are whole and undivided. Newton’s model here presupposes, not only that there exists continuity between the microscopic and macroscopic world in the composition and behavior of matter, but also that the microscopic world is consistent with the way that, as Lakoff and Johnson put it, we pick out parts of our experience and treat them as discrete entities or substances of a uniform kind. For Newton, the microscopic world, like the macroscopic one, can be understood in terms of objects and substances. Newton thus develops a model of atomic behavior that aligns unproblematically with our intuitive understanding of the world. THE INDEFINITE OBJECT

The problem introduced by quantum physics regarding how we experience ourselves and that which is external to us as bounded and discrete is set down in the limitations dictated by the Uncertainty Principle. As I mentioned in my introduction, the Uncertainty Principle describes the fact that it is not possible to ascertain simultaneously both the position and the momentum of an electron: if we bombard an electron with a beam of light of a sufficiently short wavelength to determine accurately its position, energy is unavoidably transferred to the electron, thus making it impossible to calculate its speed and direction. Conversely, measuring an electron with a long wavelength of light avoids disturbing its speed and

Experience, Perception, and the Limits of Language

35

direction, but the longer wavelength is not focused enough to identify accurately the electron’s position. Nothing can be known about the state of a subatomic particle prior to the act of measurement; however, there exists no type of measurement that can accurately reveal its total state, and thus no way to establish it as a determinate entity. Another artifact that is connected to observer interference, and that also undermines the concept of entity, is wave/particle duality, wherein a single particle appears to act in a wave-like manner if it is not observed, and appears to act in a particle-like manner if a detecting instrument is introduced. In this case, it is similarly impossible to separate the act of observation from the behavior of matter, so that the act of observation serves to constitute the object. As Bohr observes in “Causality and Complementarity,” “[we are not] any longer in a position to speak of the autonomous behavior of a physical object, due to the unavoidable interaction between the object and the measuring instruments.”20 The fact that the instrument cannot be separated from the object means that an unambiguous definition of the state (in spacetime) of a system is no longer possible, and the classical link between spacetime and causality is disrupted. And yet, we are unable to conceive of, describe, or even “observe” anything without first assuming a distinction between the perceiving subject and the perceived object—that is part of our originary perceptual interaction with the material world. Without it, we lack any reference point for even our most basic physical orientation, such as up-down.The result of the interaction between instrument and object is that we arrive at a fundamental impasse. Bohr’s solution is a compromise: one must proceed as if this interaction between the object and measuring instrument does not occur. This, then, is the foundational assumption of Bohr’s Principle of Complementarity: accept the Uncertainty Principle, but accept also that any attempt at description binds us to our most basic understanding that objects in the world are bounded and discrete, and accept that we must proceed as if there is a distinction between ourselves and the objects that we perceive. When Max Born reconceived Schrödinger’s entity-based wave packets as probability waves that refer not to entities but rather to probability distributions, he further destabilized the notion of a temporal-spatial entity, because the probability field identifies not the state of a discrete, bounded object occupying a defined region in space, but rather the range

36

Chapter 1

of the probable behavior of an object at any given time. In his essay “What Is an Elementary Particle?” Schrödinger sums up the situation: “the actual statistical behavior of electrons cannot be illustrated by any simile that represents them by identifiable things.”21 In “Indeterminism in Physics,” Schrödinger elaborates: The object referred to by quantum mechanics in this connection is not a material point in the old sense of the word. A material point in that sense is a thing situated at a given place, whether this place is discoverable or not.While this may be true, [quantum physics] fundamentally disrupts the manner in which spatial orientation is fundamental to the metaphors that we create, connected as they are to the sort of bodies that we have and how they function in our physical environment [emphasis added].22

Quantum subatomic behavior thus challenges our preference to perceive the world in terms of entities extending in continuous fashion in time and space, and Schrödinger’s words again echo Conceptual Metaphor Theory’s assertion that the spatial orientation of our bodies, projected onto the material world, is key to how we conceptualize the world through language. In quantum physics, objects change based on the nature or the presence/absence of the instrument of observation. Because the state of the subatomic object can never be fully determined, this object exists only as a tendency toward the knowledge of being, and a fundamental aspect of our embodied, situational orientation disappears. Along with it disappears the relevance and usefulness of the OBJECT schema, which constitutes a determinate aspect of our understanding of the world in which we live, as well as our ability to communicate the nature of that world. Alluding to how communication is compromised, Heisenberg explains, “there is no way of describing what happens [to an electron] between two consecutive observations. It is of course tempting to say that the electron must have been somewhere between the two observations. … This would be a reasonable argument in classical physics. But in quantum theory it would be a misuse of language which … cannot be justified.”23 Nevertheless, Heisenberg admits, one cannot avoid the notion of “entity.” In his 1958 book Physics and Philosophy, for example, Heisenberg observes that “[o]ur perceptions are not primarily bundles of colors or sounds; what we perceive is already perceived as something, the accent

Experience, Perception, and the Limits of Language

37

here being on the word ‘thing.’”24 As inappropriate and inaccurate as it is in the quantum context, the logic of “thingness” that is embedded into our language remains inescapable. The billiard ball metaphor is frequently used to illustrate Newton’s laws of motion that describe the basic force that one body exerts upon another, and are central to the classical laws of causality. In “Women, Fire, and Dangerous Things,” Lakoff observes that “[t]he concept of causation—prototypical [or classical] causation—is one of the most fundamental of human concepts. It is a concept that people around the world use in thought. It is used spontaneously, automatically, effortlessly, and often. Such concepts are usually coded right into the grammar of languages—either via grammatical constructions or grammatical morphemes.”25 According to Lakoff, evidence that causation is a fundamental aspect of our comprehension lies in the fact that it is understood in terms of a very large cluster of properties. Lakoff further observes that of the ten interactional properties associated with prototypical causation, the billiard ball metaphor possesses the first six properties: 1) there is an agent that does something; 2) there is a patient (also known as a target—a term that I will substitute for “patient” here) that undergoes a change to a new state; 3) the agent comes in contact with the target; 4) part of what the agent does (either the motion or the exercise of will) precedes the change in the target; 5) the agent is the energy source: the target is the energy goal, and there is a transfer of energy from agent to the target; and 6) there is a single definite agent and a single definite target.26 According to Lakoff, [t] he most representative examples of humanly relevant causation have all ten of these properties; that Newton’s model fulfills six of these properties suggests that it scores high on the sort of causation that is most “humanly relevant.”27 That the problem of causality is mentioned so often by Schrödinger, Bohr, Heisenberg,Wolfgang Pauli, and Max Born, and that its radical disruption both preoccupied Heisenberg and played a large part in Schrödinger’s critique of the quantum interpretation, underscores the fact that the classical concept of causality is central to our understanding of the world and our interaction with it. If one of the defining and most discussed aspects of the quantum interpretation is how it undermines the classical concept of causality, how does one account for the fact that

38

Chapter 1

almost all of the properties that Lakoff associates with the billiard ball model of causality are fulfilled in the quantum interpretation: an agent (here, the act of measurement) “does something” to the target (the electron); the target undergoes a change to a new state; part of what the agent does precedes the change in the target. The main condition that is not fulfilled is so inherent to the concept of causality that it is likely to be overlooked: for causality to be upheld, there must “a definite target”— in other words, the target must be a bounded, whole entity. The target is not a fixed object in the quantum interpretation, however, because the act of measuring can only offer a partial description of its behavior at any given time. As far as causality is concerned, the problem that the Uncertainty Principle, probability theory, and wave/particle duality introduce is entity based: there is no clear distinction between the agent (the source that exerts force) and the target (the object upon which force is exerted). The nature of the object is not fixed and bounded, but rather changes in response to the presence or absence of the agent. By virtue of its interaction with the agent, then, the status of the object cannot be fully determined, and it does not, and cannot, achieve the status of the undivided object of Newton’s classical physics. This is not the only quantum precept that defies causality, however; the introduction of Bohr’s earlier notion of quantum jumps into the quantum interpretation in order to solve the problem of electron transitions similarly challenged the laws of causality. To discuss how this happened, I will extend my analysis of the OBJECT schema and consider one that deals more directly with motion: the PATH image schema. According to Conceptual Metaphor Theory, image schemas can be simple or complex. Complex image schemas possess more than one component; the different components can be referred to and analyzed separately. At the same time, these components fit naturally together and function as a coherent whole.28 An example of a complex image schema is the SOURCE-PATH-GOAL schema, which I will use to discuss the problems that quantum jumps introduce to the chain of causality and to extend my discussion of the Uncertainty Principle. As Johnson describes it, the PATH schema includes a “source,” or starting point, and a “goal,” or endpoint. Taking these elements into account, the simple PATH schema extends into the complex image schema

Experience, Perception, and the Limits of Language

39

SOURCE-PATH-GOAL. This schema includes the concepts of both locomotion and transition, and is therefore closely connected to the elements that define the concept of causality in its prototypical form. In every case, the internal structure of this complex schema includes “a sequence of contiguous locations (the PATH) connecting the source with the goal,” so that “if you start at point A and move along a path to point B, then you have passed through all the intermediate points in between [emphasis added]”29 The quantum metaphor that concerns me here—“jumps”—has its basis in the SOURCE-PATH-GOAL schema, in the sense that it has a source, a goal, and a transition from point A to point B. In fact, quantum jumps were included in the quantum interpretation precisely to account for an electron’s transition from one state to another.30 The problem, however, lies in the fact that “quantum jumps” do not contain the central property internal to this image schema: they have no identifiable path, and so by definition do not meet the requirement that the object (the electron) pass through all the intermediate points in between the source and the goal. The starting point can be known, the endpoint can be known, but there can be no knowledge of the path taken in between the source and the goal. This gap in knowledge is what troubled Schrödinger about quantum jumps. In August of 1926, Schrödinger interrupted a holiday to travel to Copenhagen, where he engaged in a series of passionate debates with Bohr over their respective mathematic models.31 During one of these conversations, with Bohr refusing to budge, Schrödinger exclaimed: Surely you realize that the whole idea of quantum jumps is bound to end in nonsense. … The electron is said to jump from one orbit to the next and to emit radiation. Is this jump supposed to be gradual or sudden? [I]f the jump is sudden, Einstein’s idea of light quanta will admittedly lead us to the right wave number, but then we must ask ourselves how precisely the electron behaves during the jump. … And what laws govern its motion during the jump? In other words, the whole idea of quantum jumps is sheer fantasy.32

Schrödinger’s opposition to the notion of quantum jumps, an opposition that he maintained throughout his career, is based here on the fact that path of the electron could not be charted, which meant that quantum jumps violated the laws of motion and causality.33 In one of his many

40

Chapter 1

critiques of matrix mechanics, published the same year as the 1926 Copenhagen debates between himself and Bohr, Schrödinger observes, “it is hardly necessary to emphasize how much more congenial it would be to imagine that at a quantum transition the energy changes over from one form of vibration to another, than to think of a jumping electron. The changing of the vibration form can take place continuously in space and time.”34 Schrödinger’s wave model remains loyal to Newton’s first classical law of motion, which dictates that all bodies (including atomic particles) are endowed with the inertial power of persevering in their motion, a law that he infers from like properties we observe in the bodies that we are able to perceive. Schrödinger’s reference to change that is continuous in space and time preserves the conditions of the classical, prototypical causality characteristic of the PATH image schema: the object must pass through a “sequence of contiguous locations”—through all the intermediate points in between point A and point B. Schrödinger’s mathematics fulfilled these conditions, just as the wave metaphor associated with his math suggests an unbroken path between one vibration and another in an unending wave of continual motion. QUANTUM GRAMMAR

Yet physicists still use metaphors. … What other choice do we have? We must breathe, even in thin air. —Alan P. Lightman

Lakoff writes that the notion of causation is “coded right into the grammar of languages,” but it is not only causation that is coded into language—it is the notion of an identifiable location and state, of the separation between subject and object, and of the separation of one object from another in space and time.The quantum interpretation disrupts all of these primary assumptions, and thus defies at many points the conditions that structure our concept formation and our language. No wonder, then, that Schrödinger, Bohr, and Heisenberg spent so much time talking and writing about what quantum physics revealed about the nature of language, what language revealed about quantum physics, and why any attempt to describe quantum concepts in language was likely to

Experience, Perception, and the Limits of Language

41

end in failure, or at least misrepresentation. In a 1926 letter to Pauli, Heisenberg wrote, “we couldn’t doubt that it [i.e., quantum mechanics] was the correct scheme but even then we didn’t know how to talk about it,” a fact that, he says, “left us in a state of almost complete despair.”35 It is this collision between quantum phenomena and the description of quantum phenomena that I turn to in this section. Of quantum physics, Liliane Papin observes that “never before had scientists been so directly confronted with the limitations of language and made so aware of its sedimented and root metaphors, as well as of its cultural linguistic bounds.”36 While retaining the spirit of Conceptual Metaphor Theory—particularly in her emphasis on the problem of entity—Papin begins with a more specifically grammatical account of the disjuncture between quantum concepts and language. According to Papin, part of the problem for the new physics is embedded in the very structure of Indo-European language. Along the lines of Conceptual Metaphor Theory, she observes that this grammatical structure underlies our immediate and subliminal process of perception.37 Papin goes on to argue that “Indo-European languages are, broadly speaking, based on a structural division between nouns, which are associated with ‘thingness,’ and verbs, which denote movement, action, or state of being.”38 The recognition in quantum theory of the dual character of light as wave and particle, for example, tore apart not only the fundamental correspondence assumed between nouns and specific attributes, but also the basic either-or categorization central to classical science.39 Papin notes that “the descriptive inadequacy of words, nouns in particular, is the rule rather than the exception in the realm of modern physics because in quantum physics elementary particles exist not as separate entities but only as a set of relations. They are constituted by change, process, and events—all phenomena that, in Indo-European language, are ordinarily ascribed to verbs.”40 According to Papin, “mere nouns” become obsolete and deceptive in quantum physics because the concept of event replaces the concept of thingness.41 One might even question the extent to which the notion of an event is appropriate, however, given the fact that subatomic phenomena disrupt the causality and temporal/spatial relations upon which the notion of “events” rests. As Heisenberg observes in Physics and Philosophy, the probability function “does not in itself represent a course of events in

42

Chapter 1

the course of time. It represents a tendency for events and our knowledge of events [emphasis added].”42 The concept of a probability wave, writes Heisenberg, “introduced something standing in the middle between the idea of an event and the actual event, a strange kind of physical reality just in the middle between possibility and reality.”43 A reliance on nominal language containing conceptual metaphors founded on the notion of substance and entity renders misleading any attempt to describe what amounts to potentialities of being, rather than the state of being itself. Heisenberg went so far as to claim that it was inherently impossible to use existing language to describe quantum phenomena. For the way he addresses the roots of the problem, Heisenberg’s observations bear quoting at length: Every description of phenomena, of experiments and their results, rests upon language as the only means of communication. The words of this language represent the concepts of daily life, which in the scientific language of physics may be refined to the concepts of classical physics.These concepts are the only tools for an unambiguous communication about events, about the setting up of experiments, and about their results. If therefore the atomic physicist is asked to give a description of what really happens in his experiments, the words “description” and “really” and “happens” can only refer to the concepts of daily life or of classical physics. As soon as the physicist gave up this basis he would lose the means of unambiguous communication and could not continue in his science. Therefore, any statement about what has “actually happened” is a statement in terms of classical concepts and—because of thermodynamics and of the uncertainty relation—by its very nature incomplete with respect to the details of the atomic events involved.The demand to “describe what happens” in the quantum-theoretical process between two successive observations is a contradiction in adjecto, since the word “describe” refers to the use of the classical concepts, while these concepts cannot be applied in the space between the observations; they can only be applied at the points of observation.44

Because Heisenberg’s Uncertainty Principle determines that the amount of information available about the state of a subatomic particle is inherently limited, the information required to make up an event or “happening” is missing. To put it another way, if we know the main subject of the sentence, we cannot know the verb; if we know the verb, we cannot know the subject.This is why there is an inescapable contradiction in the

Experience, Perception, and the Limits of Language

43

sentence “Describe what really happens in a quantum-theoretical process.” I would add that, in shifting our attention to happenings or events, we are merely creating more nouns that fail to capture the radically probabilistic nature of quantum phenomena, and that even the term “change” suggests being able to identify the movement of something from one state to another, knowledge that the quantum interpretation does not allow for. It is worth noting that the word “nominal” has its root in the Latin world “nominalis”—“of names,” and that the very act of naming is fundamental to a language that possesses a bias toward the nominal.To name something is to create an entity, whether it be an object, a feeling, or even a process. For this reason, the terms “particle” and “wave” have an uneasy status in quantum physics; we cannot avoid naming them as such, even as we know that to do so is to misrepresent the situation, for electrons are neither and both. While Heisenberg insisted in his 1926 letter to Pauli that he considered the proposition that the world is continuous “more than ever as totally unacceptable,” he went on to complain, “But as soon as it [the world] is discontinuous, all words that we apply to the description of facts are so many numbers. What the words ‘wave’ or ‘corpuscle’ mean we know not any more.”45 Even the use of the term “probability waves” is caught up in the distorting reliance on entity-based language. Addressing the misleading nominalization of probabilities as waves, Hanna Pulaczewska writes that the modeling of probabilities as “spreading waves” “allows us to conceive probabilities as ‘sort-of ’ things, capable of being split, propagated, reflected, reshaped, and destroyed.”46 The probability function does not in itself represent a course of events in the course of time, but rather a statement concerning our degree of knowledge about a situation. As Schrödinger notes, in our acts of observation “we have only disturbed or changed the probability of finding this or that value of the velocity if we measure it. The implications as to ‘being’ or ‘having’ are misconceptions, to be blamed on language.”47 Here, language is not merely in conflict with the quantum ambiguity of nature, but is shut down entirely by an absent referent that does not even have the status of a likelihood of being, but rather expresses only a likelihood of being known. As Pulaczewska observes, another misleading term related to our entity-based language is “state.” We talk of physical systems as being in a

44

Chapter 1

certain state, as leaving one state and entering another, or as transitioning from one state to another.48 Because the concept of state is an ontological entity-making metaphor, it “becomes commensurate with the notion of a material particle,” so that states may be described as interacting with each other or with other entities.49 The same problem that I described earlier in relation to events occurs with states: we are, in fact, talking about the probability of states or transitions between states, rather than actual states. Bohr ascribed the unsatisfactory tendency to describe probabilities in terms of the “mechanical pictures of stationary states” to the essential failure in quantum physics of the pictures in space and time on which the description of natural phenomena had hitherto been based.50 In his 1975 article “Development of Concepts in the History of Quantum Theory,” Heisenberg admits that “even when we … accepted the quantum jumps, we did not know what the word ‘state’ could mean.”51 Heisenberg suggested replacing the term “state” altogether with the term “potentiality,” because while it is difficult to conceive of coexistent states, one can conceive of “coexistent potentialities.”52 The advantage of this revision lies in the fact that potentialities can overlap, whereas “states” preserve the misleading, entity-based presumption of mutual exclusivity; however, it remains difficult to conceive of potentialities without entertaining the notion that some “ability” (potentiality) is “inside” or “possessed by” an unnamed entity. In his February 27, 1963 interview with Thomas Kuhn, Heisenberg reflected on the intransigence of language where quantum theory and relativity are concerned, and on the resulting “linguistic impasse” that arises: As soon as you come to velocities, near the velocity of light, then it is not only so that Newtonian physics doesn’t apply, but the point is that you even don’t know what you mean by “velocity.” Well, as you well know, you cannot add two velocities and so on, so just the word “velocity” loses its immediate meaning. That, I think is a very characteristic feature of what I mean by close [sic] system; that is, when you have such a system and you get disagreement with facts, then it means that you can’t use the words any more.You just don’t know how to talk. 53

Heisenberg’s sense of the inadequacy of language led him to be wary of using a linguistic system of representation at all, a wariness he expressed in another of his interviews with Kuhn: “I feel that as a constant warning

Experience, Perception, and the Limits of Language

45

that one must be extremely careful about any words which one used [sic]. So whenever I say sentences about an electron moving in an atom or a wave moving there and there, I always feel unconsciously how dangerous every word is which I use.”54 It is not surprising that Heisenberg consistently expressed misgivings about language: not only did its nature and origins in macrocosmic experience pose a real problem when one attempted to use it to communicate behavior at the quantum level, but also, from the beginning Heisenberg’s quantum interpretation was based on abstract algebra— the primary expression of matrix mechanics, probability theory, and the Uncertainty Principle. In 1930, concerning the fact that words can only describe concepts formed from the experiences of daily life, Heisenberg wrote, “Fortunately, mathematics is not subject to this limitation, and it has been possible to invent a mathematical scheme—the quantum theory—which seems entirely adequate for the treatment of atomic processes.”55 Much later, in his February 22, 1963 interview with Kuhn, Heisenberg recalled: There was, of course, from the very beginning the impression that if we find a consistent mathematical scheme, then it should sooner or later be possible also to avoid the contradictions in the way in which we talk about it. After all, mathematics is meant to be consistent; if mathematics is not consistent, then it’s just wrong.Therefore, as soon as you find a consistent mathematical scheme then you should be able, sooner or later, to find also the right words to talk about it.56

For Heisenberg, questioning the adequacy of words was a genuine epistemological concern—but it was also a way of defending his abstract, algebraic quantum interpretation from Schrödinger’s more empirical and language-oriented wave mechanics. The preceding passage also alludes to another solution that Heisenberg considered, which once again concerned the relationship between mathematics and language. Throughout his career, Heisenberg believed in the primacy of theory over experiment, and the primacy of mathematics over language; once having established the theory, he argued, the experiments would come along that would confirm that theory, just as correct language would follow from the correct form of the mathematics. Heisenberg also approached the language problem from a historicized point of view; he believed that while the axioms of mathematics expressed

46

Chapter 1

an enduring truth, the limitations of (classical) language might be viewed as temporary. Unlike Bohr and Schrödinger, Heisenberg implicitly detached language from its origins in embodied experience and instead suggested that it is a social construction with the potential to evolve. As Kristian Camilleri observes, for Heisenberg “[t]he indispensability of classical concepts originates from the historical fact that we have no other language through which we describe what is given to us in experience [emphasis added].”57 By constructing the limitations of language as a historical artifact, Heisenberg could sidestep the intractable problem of the fundamental incompatibility between the quantum realm and the realm of common-sense experience from which language is drawn. Instead of trying to generate a theoretical or descriptive framework that followed from and could be reconciled with existing language, Heisenberg assigned primacy of place to his theory, and promised that in time language would evolve to fit the mathematical abstractions that he saw as the best and most suitable mode of expression. BOHR AND THE DETERMINATE ROLE OF DESCRIPTION

There is no quantum world. There is only an abstract quantum physical description. It is wrong to think that the task of physics is to find out how nature is. Physics concerns what we can say about nature. —Niels Bohr

While Heisenberg suggested that language could evolve to overcome its deficits—and in the meantime, the more reliable mathematics ought to be the primary mode of expression—Bohr accepted the enduring necessity of using ordinary language and classical terminology. With reference to both causality and the obligation to communicate scientific findings, Bohr challenged the suggestion that a revolutionary new mode of description would or could emerge: The view has been expressed from various sides that some future more radical departure in our mode of description from the concepts adapted to our daily experience would perhaps make it possible to preserve the ideal of causality also in the field of atomic physics. Such an opinion seems to be due to a misapprehension of the situation. For the requirement of communicability of the circumstances and results of experiments implies that we can speak of welldefined experiences only within the framework of ordinary concepts.58

Experience, Perception, and the Limits of Language

47

Bohr is referring here to the conditions that define our manner of communicating—conditions that emerge out of our macrocosmic interaction with the material world. For Bohr, classical concepts such as continuous spacetime and causality were not features of reality, but rather artifacts of language.While Bohr acknowledged the unavoidable interaction between the object and the measuring instrument and the indeterminacy of measurement, he believed that this interaction “in principle cannot be taken into account if these instruments according to their purpose shall allow the unambiguous use of the concepts necessary for the description of experience.”59 This meant that for Bohr, while the Uncertainty Principle and wave/particle duality dictate that the object can never be unambiguous or detached from the apparatus, we must accept the fact that we are bound to speak of it as if it is. The determinate role that description plays for Bohr cannot be overemphasized. The fact that Bohr was most inclined to view the classical concepts in the context of description, rather than in the context of the material world, is summed up in a passage from his “Discussions with Einstein”: “we must realize the unambiguous interpretation of any measurement must essentially be framed in terms of the classical physical theories, and we may say that in this sense the language of Newton and Maxwell will remain the language of physicists for all time.”60 Bohr’s focus was less on the attributes of matter than on the question of how to render intelligible the observed evidence of these attributes, and while he accepted the quantum interpretation and played a key role in its development and canonization, he rejected Heisenberg’s argument that the incompatibility between the attributes of matter and classical concepts constituted an unbreachable impediment to the communication of quantum behavior. On October 26, 1935, Bohr replied to a letter from Schrödinger in which Schrödinger asked Bohr repeatedly why he, an advocate and cofounder of the quantum interpretation, should insist that the classical concepts must be retained. Bohr explained, “my emphasis on the unavoidability of the classical description of the experiment refers in the end to nothing more than the apparently obvious fact that the description of every measuring apparatus basically must contain the arrangement of the apparatus in space and its function in time, if we are to be able to say anything at all about the phenomena.”61

48

Chapter 1

While Bohr rejected Schrödinger’s own “classical” position that the wavefunction represented a new kind of physical “reality”—that it described the actual state of an atomic system—he did believe that these concepts inevitably condition what we can say about that world. In his widely celebrated 1949 article “Discussion with Einstein on Epistemological Problems in Atomic Physics” Bohr writes, it is decisive to recognize that, however far the phenomena transcend the scope of classical physical explanation, the account of all evidence must be expressed in classical terms. The argument is simply that by the word “experiment” we refer to a situation where we can tell others what we have done and what we have learned and that, therefore, the account of the experimental arrangement and of the results of observation must be expressed in unambiguous language with suitable application of the terminology of classical physics [emphasis in original].62

Bohr’s Principle of Complementarity was based on his belief that a full description of nature required that all experimental evidence be taken into account, just as it was based on the fact that experimental outcomes, however contradictory and “unclassical” they might be, must be expressed using language derived from the classical concepts.The problem for Bohr lay in the fact that an unproblematized use of ordinary concepts could not apply at the quantum level, so that the act of communicating these attributes was already an act of miscommunication.63 For Bohr, therefore, description in language could only be seen as an idealization—the only possible way to imagine language as the mode of description while at the same time to acknowledge that it does constitute a direct representation of atomic behavior. Heisenberg remained uneasy about Bohr’s solution, and only reluctantly admitted that while “the first language that emerges from the process of scientific clarification is in theoretical physics usually a mathematical scheme, it remains necessary to translate this mathematical scheme into language for the layperson.”64 Heisenberg was troubled in particular by the “intrinsic uncertainty of the meaning of words” that forces one to “rely on some concepts that are unanalyzed and undefined.”65 On what he saw as the equivocal nature of Bohr’s Principle of Complementarity, Heisenberg wrote, “when this vague and unsystematic use of language leads into difficulties, the physicist has to withdraw into

Experience, Perception, and the Limits of Language

49

the mathematical scheme and its unambiguous correlation with the experimental facts.”66 Heisenberg’s cautionary words on Bohr were lost on many who later invoked the Principle of Complementarity in nonscientific or pseudoscientific contexts. In The Atom in the History of Human Thought, Bernard Pullman identifies the root cause of later distortions of Bohr’s Principle: Complementarity attempts to offer a possible solution to the dilemma of describing atomic phenomena while preserving the use of the ordinary language of physics. … But even from that perspective, the meaning of the word is likely to be misunderstood by the casual reader. The conventional notion of complementarity connotes synergistic congruence—in other words the compatible union of fragments. … The type of complementarity conceived by Bohr, on the contrary, is characterized by a fundamental and irrevocable contradiction between the elements involved.67

Pullman is right to locate the misinterpretation of complementarity in assumptions about “synergistic congruence,” which is born out in later uses of complementarity that forget its symbolic and largely heuristic nature in the quantum context as well as its role as the description of a contradictory alternating relationship.These uses associate it instead with the real, ontological unification of oppositions. Pullman introduces his own distortion, however, when he claims that Bohr’s complementarity was entirely about “irrevocable contradiction”; rather, Bohr sought with complementarity both to express the mutually exclusive either/or nature of the classical concepts when applied to quantum phenomena, and the symbolic unity and harmony of “both/and.” ANTHROPOMORPHISM: THE INDIVIDUAL VERSUS THE COLLECTIVE

Language always includes agency, and agency and intention are frequently impossible to distinguish in language. —Elizabeth Leane

I have shown how, in their debates over and reflections upon the constraints of language, Schrödinger, Bohr, and Heisenberg frequently alluded to what would later become the cornerstone of Conceptual

50

Chapter 1

Metaphor Theory: the relationship between experience, perception, and language. I have also shown that when they talked about the limits of language, the three physicists prefigured Conceptual Metaphor Theory’s proposition that our metaphors derive from the sense of ourselves as discrete and bounded entities. In their later works, Schrödinger and Bohr attempted to address the more philosophical and epistemological implications of quantum physics using language that would be accessible to the layperson. This attempt frequently led them to use humanizing metaphorical conceits that cast subatomic phenomena as familiar objects or characters. At the same time that they used anthropomorphizing metaphors, these men appear to have been acutely aware of the distortions that these conceits inevitably import, and throughout these works they display an ongoing tension between the desire to communicate the concepts and the desire to maintain the integrity of those concepts. This tension leads to a curious process wherein a metaphorical conceit is presented and then immediately put under erasure by the demonstration of its inadequacy. According to Conceptual Metaphor Theory, an enormous number of elements in our conceptual system are also oriented with respect to whether or not they correspond to the properties of the prototypical person.68 Lakoff and Johnson observe, “our most obvious ontological metaphors are those where the physical object is specified as being a person, a process that allows us to make sense of phenomena in the world in human terms that we can understand on the basis of our own motivations, goals, actions, and characteristics … adding intention and agency to the object under consideration.”69 In his 1935 collection of essays, Science, Theory, and Man, Schrödinger uses a military conceit to animate the bending or “turning” of light rays. Schrödinger explains that portions of the wave front “advance” at varying speeds, adding that “the turning is effected in the same way as with a company of soldiers marching in line, who are ordered to ‘right wheel’ by taking steps of varying lengths.”70 He elaborates on this image in the following passage: No word of command as to direction is given, but simply the order that each man must march or run as fast as he can. … After a time it will be noticed that the line of advance, when looked upon as a whole, is not straight, but shows a definite curvature. Now this curved route is precisely the one along which the soldiers reach any place on their way in the shortest possible time. … Although

Experience, Perception, and the Limits of Language

51

this may seem remarkable there is actually nothing strange about it, for after all, by hypothesis, each soldier has done his best to travel as quickly as possible [emphasis in original].71

Schrödinger’s decision to depict light as a cast of soldiers engaged in behavior familiar to us, by virtue of our knowledge of military conventions, effectively translates the counterintuitive behavior of light into meaningful human terms. The utility of the soldier metaphor lies not only in its accessibility, but also in the fact that it aligns with our conventional conception of matter as behaving in identifiable, predictable ways according to logical, meaningful rules: we associate soldiers with orderly behavior and loyalty to strictly enforced codes of behavior.When light is construed collectively as waves it can be said to follow the orderly, predictable logic of classical laws. When Schrödinger figures light as composed of individual soldiers, each following the same directive but to different effect (those further from the center will travel less distance than those closer to the center), Schrödinger’s metaphor also relies on the logic of the individual, and suggests the discontinuities of Heisenberg’s particulate matrix mechanics. Schrödinger’s metaphor thus undoes itself: the unity and continuity of the collective is undermined in the moment that the waves are seen to be composed of separate individuals. While Schrödinger allows the military metaphor to stand, he purposely questions other metaphors that portray a relationship between the individual and the collective. In relation to statistical behavior, he first writes: Electrons for instance, correspond to membership in a club. … Any person eligible to membership in that club represents a well-defined state an electron can take on. If the person is a member, that means there is an electron in that particular state. According to Pauli’s exclusion principle, there can never be more than one electron in a particular state. Our simile renders this by declaring double membership meaningless—as in most clubs it would be.72

The club metaphor effectively captures the mutual exclusivity of electron states; however, it elides the fact that “states” are really overlapping potentialities. Schrödinger qualifies his simile immediately, observing that while states suggest “well-defined individuals,” “it is a fact that electrons are not thus well-defined, nor can they be likened to individuals.”73 While Schrödinger finds the metaphor useful to explain subatomic

52

Chapter 1

behavior, he feels obliged to correct immediately the distortion introduced by the metaphor’s entailments. Schrödinger again contrasts the individual and the collective in explaining the process of entropy: In every physical and chemical process there is a transition from relatively well-ordered conditions among the groups of atoms and molecules to less orderly conditions—in other words, a transition from order to disorder, just as might be expected if each individual member of the mass followed its own way more or less without any plan and no definite law. The exact laws which we observe are “statistical laws.” In each mass phenomenon the laws appear all the more clearly, the greater the number of individuals that cooperate in the phenomenon. … And the statistical laws are even more clearly manifested when the behavior of each individual entity is not strictly determined.74

Here, if the individual particle is rogue, the collective expresses statistical behavior that conforms to identifiable laws and that cannot be said to make definite predictions as to what the individual will do next. Schrödinger warns us, however, to be wary of developing a misconception regarding the individual particles—in particular to avoid the assumption that it is merely the mass “crowding” that makes it difficult to distinguish one particle from another.75 According to this faulty logic, if we could just separate the individual/particle out from the mass then we could “see” it. Schrödinger reminds us, however, that “the elementary particle is not an individual; it cannot be identified, it lacks ‘sameness’ … the unsuspected ‘this’ is not quite properly applicable to, say, an electron, except with caution, in a restricted sense, and sometimes not at all.”76 In other words, it is not the fact that they are crowded together that makes it difficult to isolate the individual particles; as statistical probabilities, particles are really mass points that possess no discrete identity. So important does Schrödinger hold dispelling the misconception that particles are individuals that he concludes his book with yet another warning.This, then, is the final sentence of Science, Theory, and Man: “The point is that they [electrons] are not individuals which could be confused or mistaken one for another. Such statements are meaningless.”77 In the end, Schrödinger seems like a man desperately trying to extract himself from his own metaphors—using them to make concepts meaningful, then compelled to unravel the unwelcome excess of meaning.

Experience, Perception, and the Limits of Language

53

In his 1927 article “The Quantum Postulate and the Recent Developments of Atomic Theory,” Bohr introduces the concept of individuality in relation to the Uncertainty Principle and the limits of measurement: Indeed, the position of an individual at two given moments can be measured with any degree of accuracy; but if, from such measurements, we would calculate the velocity of the individual in the ordinary way, it must be clearly realized that we are dealing with an abstraction, from which no unambiguous information concerning the previous or future behavior of the individual can be obtained.78

While Bohr championed Heisenberg’s discontinuous, particulate theory of atomic matter, and believed that the classical concepts could only be used symbolically, here, he emphasizes the fact that the concept of “individuality” must be seen as an abstraction and acknowledges that the basic conditions for an individual particle’s existence as anything other than a metaphorical abstraction cannot be met. Bohr accepts the fact that these sorts of metaphorical abstractions are indispensable, by virtue of their connection to our ordinary spacetime experience; however, he remains cognizant of the paradoxes introduced by the anthropomorphic bias in language. On the disappearance of causality in the Uncertainty Principle, he observes that “an interesting example of ambiguity in our use of language is provided by the phrase used to express the failure of the causal mode of description, namely, that one speaks of a free choice on the part of nature.”79 Bohr cautions against projecting onto subatomic phenomena the notion of a personified nature exercising “free choice,” because it invites us to forget that we are talking about a field of probabilities of knowledge and not a series of actual possibilities “considered” by the electron.80 Nevertheless, however inaccurate it is with respect to quantum behavior, we remain caught within causal logic; because we conceptualize action as self-caused change, even acknowledging the failure of causality leads to the trap of assigning agency and intention to atomic particles. The use of metaphors that suggest “acting,” “knowing,” and “choicemaking” individual particles renders their behavior meaningful on a human scale. Because quantum subatomic behavior contradicts human experience, however, the anthropomorphizing metaphors distort the nature of this behavior. Rather than eschew figurative language

54

Chapter 1

altogether, Bohr and Schrödinger at one and the same time accept anthropic metaphors and acknowledge their failure. To use metaphors to describe quantum physics appears to require a staged approach: one introduces the metaphor, and then immediately identifies its fundamental deficiency.The real meaning of the quantum phenomenon under consideration is conveyed in this second act where the precise nature of the metaphor’s inadequacy is revealed. This strategy of conveying meaning via the undoing of a metaphorical conceit appears to be so central to communicating quantum phenomena that one is tempted to designate a new conceptual category—the “quantum metaphor”—to represent this process. (DIS)ORIENTING JOURNEYS

“Theory” is a product of displacement, comparison, a certain distance. To theorize, one leaves home. —Georges Van Den Abbeele

As I alluded to in my introduction, one of the contributions of Conceptual Metaphor Theory is the manner in which it offers a detailed narration of how time is typically figured in terms of movement in space.This TIME AS SPACE image schema emerges from the fact that space is a fundamental source for our orientation in the world and for our metaphors, whereas time is a quality so abstract that it escapes direct perception. As Lakoff observes in “The Contemporary Theory of Metaphor,” “[i]n our visual systems, we have detectors for motion and detectors for objects/locations, but we do not have detectors for time. Thus it makes sense that time should be understood in terms of things and motion.”81 In Making Truth: Metaphor in Science,Theodore Brown expands upon this observation to argue that the “abstract quantity, time, is conceptualized in terms of entities that we deal with in everyday life”—that, metaphorically speaking, time is an entity, a “thing,” with future times occupying a position in front of the observer, and past times occupying a position behind the observer.82 Two related premises are embedded here: that the temporal metaphors we create relate to our embodied experience, and that time becomes movement or location in space. In the quantum realm, however, interaction between space and time is disrupted; for example, the notion

Experience, Perception, and the Limits of Language

55

of quantum jumps from one electron state to another leaves us without knowledge of the “landscape” that the electron passes through in order to reach a new state, and the Uncertainty Principle dictates that position in space and velocity in time cannot be determined simultaneously, so that to be oriented to the object in space is to become disoriented from it in time. A major consequence of this conceptual disjunction is a disruption in our ability to comprehend “time passing” in spatial terms, so that a primary marker vanishes for how we experience, conceptualize, and understand time. Because how we inhabit science does not occur on a microscopic scale, but rather on a macroscopic scale, it can be represented in spatial terms; for example, as one’s position relative to, and passage through, the “field” of science. As Kai Mikkonen observes, “the notion of travel is prone to give identity and narrativity to a series of events since it ‘humanizes’ the experience of time and space.”83 According to Mikkonen, “the travel metaphor is … not only a way to think about narrative; it also provides one with the means to think through narrative.”84 This second phrase, “think through narrative,” has a double meaning: it implies both the act of interpreting a narrative—thinking it through—and the act of using narrative as a device for “thinking.” As scientists, Schrödinger, Heisenberg, and Bohr all depict, in ways that reflect their unique sensibilities, the temporal axis of scientific progress and discovery as movement or travel through space. In their more philosophical work, Bohr and Schrödinger attempt to make meaningful the process of scientific discovery by casting themselves as travelers on a journey. Focusing on the second meaning of “thinking through narrative,” I will now examine how the travel metaphor provides for these men a device for “thinking about thinking”—that is, a way of narrativizing and thus humanizing the practice of science. Elaborating on his comments, Mikkonen states that “the experiential frame of journey and travel is marked by the subjective, human scale of space and time.This marking includes … the orientation provided by the traveling individual and his or her experiencing point of view, and the structuring of time as a spatial surface that is covered and created by a path through it.”85 Both Bohr and Schrödinger draw on the LIFE IS A JOURNEY image schema to universalize their experience of what it means to advance a

56

Chapter 1

new theory of the material world. For them, developing new theories and new concepts constitutes walking along unfamiliar paths, which introduces the potential for false steps. Classical concepts, however, constitute the domesticated terrain upon which physicists walk, while novel concepts constitute the tools through which the physicist navigates unfamiliar terrain. In his 1928 article “The Quantum Postulate and the Recent Development of Atomic Theory,” Bohr observes how “in a field like this where we are wandering on new paths and have to rely upon our own judgment in order to escape from the pitfalls surrounding us on all sides, we have perhaps more occasion than ever at every step to be remindful of the work of the old masters who have prepared the ground and furnished us with the tools.”86 As his source domain, Bohr uses the familiar metaphor of scientific discovery as a journey and scientists as travelers, in order to appeal to the reader’s customary perceptual orientation. Bohr’s metaphor here contains within it a tension, however; on the one hand, he writes of “wandering on new paths”; on the other hand, he writes of being “remindful of the old masters” who have “prepared the ground” and “furnished us with the tools.” The fact that he describes himself and his colleagues as “wandering” suggests that they lack a clear sense of direction, and the reference to “pitfalls” suggests that these new paths are treacherous, and that a false step may entrap them in wrong thinking. Bohr’s sense of wandering, and of a perhaps disorienting newness, is not surprising, given the rapid pace at which the quantum interpretation was advancing at the time that he wrote this passage. Early in 1925, Heisenberg had calculated the mathematics that represented the first iteration of the quantum interpretation. That same year, Max Born and Pascual Jordan devised a way to translate Heisenberg’s mathematics into matrices, and the first coherent quantum interpretation emerged. Then, in 1926, Schrödinger published four papers in quick succession in which he detailed his wave mechanics, presenting a major challenge for matrix mechanics and the quantum interpretation. Max Born then derived his probability theory from Schrödinger’s wave equation—a reinterpretation of wave mechanics that Schrödinger vehemently opposed. In March of 1926, while studying with Bohr, Heisenberg developed the

Experience, Perception, and the Limits of Language

57

mathematical foundations of his Uncertainty Principle, and in 1927 Bohr devised his own Principle of Complementarity. Bohr had been very impressed by Schrödinger’s wave mechanics, but he was also deeply invested in the quantum interpretation, major portions of which were being developed at his own Institute in Copenhagen.With all of this activity, Bohr found himself at the center of the most fertile and tumultuous time in atomic physics; it is understandable that he might have felt somewhat disoriented in the face of such fast-paced and radical change—especially with the dispute raging around Schrödinger’s wave mechanics versus Heisenberg, Born, and Pascual Jordan’s matrix mechanics, and with the quantum interpretation not yet fully set. This was, after all, around the same time that Heisenberg admitted to falling into despair over the complications thrown up by the quantum interpretation and his inability to master its implications. The danger of falling into wrong thinking must indeed have seemed a possibility to someone as constitutionally cautious as Bohr. At the same time, Bohr’s own Principle of Complementarity was founded on his conviction that a description of atomic behavior could only be expressed in classical terms—in other words, in the tools of the masters. It’s no wonder that Bohr seems to be torn between following the new path, and standing upon the ground of classical physics, and that the travel metaphor he developed does not quite cohere. Like Bohr, Schrödinger represents the development of quantum theory as a journey. In 1935, while living in London and having just fled the Nazis, Schrödinger wrote an essay titled “Science, Art, and Play.” At one point in the essay, Schrödinger writes, “In experimental science facts of the greatest importance are rarely discovered accidentally; more frequently new ideas point the way towards them. The ideas which form the background of the individual sciences have an internal interconnection.”87 Schrödinger’s reference to ideas as “pointing the way” suggests that ideas are like signposts, while facts constitute entities that are discovered along the way. Schrödinger’s construction of facts as entities accords with his belief that science ought to be grounded in evidence, and that theory—which proved sufficient for Heisenberg and he accorded the status of facts—could only for Schrödinger provide direction to that which must subsequently be empirically verified.

58

Chapter 1

In his 1952 article “Are There Quantum Jumps? Part I,” Schrödinger returns to the same theme when he observes that A theoretical science, unaware that those of its constructs considered relevant and momentous are destined eventually to be framed in concepts and words that have a grip on the educated community and become part and parcel of the general world picture—a theoretical science, I say, where this is forgotten, and where the initiated continue musing to each other in terms that are, at best, understood by a small group of close fellow travelers, will necessarily be cut off from the rest of cultural mankind; in the long run it is bound to atrophy and ossify, however virulently esoteric chat may continue within its joyfully isolated group of experts.88

Schrödinger uses the travel metaphor here to emphasize a point that became somewhat of a refrain for him: science must be expansive in its orientation, and scientists must not fall into a state where they cut themselves off from the interested public that exists beyond their narrow field. In fact, one of Schrödinger’s recurring criticisms of quantum theory was that it cleaved so strongly to its abstract algebra and focused so intently on generating ever-new theoretical constructs to bolster the quantum interpretation that it became isolated, not just from other theories in the present, but from a tradition that reached back into the past and that ought to continue forward into the future. Schrödinger offers up a similar journey when he describes the process of conveying the ideas of the new physics to the layperson as “the overcoming of distances in order to promote communication and understanding,” and observes: As a rule the way to the masses is long and less direct and in certain rare cases it may appear as though a complete barrier existed. However, we would ask that the right to exist should be acknowledged even for these distant blossoms on the tree of knowledge; our reason being that they must first fertilize each other in order that other branches shall be able to bear such obvious fruits, palpable to the entire community.89

If, for Schrödinger, the scientist who is engaged in his work is a forwardlooking explorer, the scientist as communicator should aspire to span the sometimes-difficult terrain that separates him from nonscientists. Here, rather than representing the scientist as looking forward toward new discoveries, he positions the scientist as one who surveys the surrounding

Experience, Perception, and the Limits of Language

59

landscape to assess how best to reach nonscientists. Schrödinger’s goal here is not finally to describe the forward-moving trajectory of new discoveries, but rather to consider the communicative process itself, which involves the difficult task of traveling a long and indirect path along which one may encounter almost insurmountable challenges. For Schrödinger, these challenges are as much about communicating the scientific principles as they are about discovering them, which suggests that for him describing these discoveries in accessible terms is at least as important as formulating a scientific theory. Heisenberg, too, ventures into the travel metaphor, although, characteristically, he fixes his attention less on how to maintain our bearings and more on how they are undermined. In a chapter from Physics and Beyond tellingly titled “Fresh Fields,” Heisenberg recalls the reservations that he had concerning Schrödinger’s introduction of wave mechanics— more specifically, the way in which wave mechanics so blatantly contradicted the quantum interpretation. Heisenberg begins the chapter in the following way: If I were asked what was Christopher Columbus’ greatest achievement in discovering America, my answer would not be that he took advantage of the spherical shape of the earth to get to India by the western route—this idea had occurred to others before him—or that he prepared his expedition meticulously and rigged his ships most expertly—that, too, others could have done equally well. His most remarkable feat was the decision to leave the known regions of the world and to sail westward, far beyond the point from which his provisions could have got him back home again. In science, too, it is impossible to open up new territory unless one is prepared to leave the safe anchorage of established doctrine and run the risk of a hazardous leap forward. … [W]hen it comes to entering new territory, the very structure of scientific thought may have to be changed, and that is far more than most men are prepared to do.90

Heisenberg’s use of the journey metaphor is consistent with the extent to which his matrix mechanics breaks entirely with classical concepts, which are replaced by discontinuities, uncertainty, and the collapse of the distance between observer and observed. Heisenberg uses this metaphor to criticize the manner in which Schrödinger’s model clung to the classical concepts—the way that it failed to leave the “known regions” and “safe anchorage” of existing scientific notions. Just as Schrödinger criticizes

60

Chapter 1

Heisenberg and his colleagues for being too quick to embrace the new at the expense of communication and continuity, so does Heisenberg criticize Schrödinger for clinging onto what Heisenberg believes to be outmoded ways of thinking. In his 1958 article “The Representation of Nature in Contemporary Physics,” Heisenberg offers a lengthy journey metaphor that correlates the challenges facing his contemporary society with those presented by the new physics. Here, Heisenberg positions quantum physics as the more concrete source domain, from which he draws characteristics to elucidate the more abstract spirit of his time, the latter representing the target domain. In describing how this new social situation forces us “to confront ourselves,” Heisenberg writes: With the seemingly unlimited expansion of his material might, man finds himself in the position of a captain whose ship has been so securely built of iron and steel that the needle of his compass no longer points to the north, but only toward the ship’s mass of iron. With such a ship no destination can be reached; it will move aimlessly and be subject in addition to winds and ocean currents. But let us remember the state of affairs of modern physics: the danger only exists so long as the captain is unaware that his compass does not respond to the earth’s magnetic forces. The moment the situation is recognized, the danger can be considered as half removed. For the captain who does not want to travel in circles but desires to reach a known—or unknown— destination will find ways and means for determining the orientation of his ship.91

For Heisenberg, the object of science in quantum physics is no longer nature, but rather nature exposed to our modes of questioning; what we encounter is no longer the material world, but rather our knowledge of the material world. In similar fashion, the expansion of technology means that we live in a world so completely transformed by humans that whatever we are doing, what we encounter is not nature (to which humanity has hitherto opposed itself), but rather structures that we have made. In other words, what we encounter is ourselves.92 Heisenberg argues here that society must recognize that the old ways of orienting ourselves are outmoded; for example, our association of technological advancement with progress (in the form of increasing mastery over nature) does not hold when the threat to humankind is no longer a potentially hostile natural world, but rather our own potentially hostile use of technology

Experience, Perception, and the Limits of Language

61

against one another. Just as quantum physics accepts that scientific knowledge is now defined by how we interact with our “ways of knowing” and is thus self-reflexive, so must humanity accept that it is defined by how humans interact with one another. Of the three men, Heisenberg’s conceptual contributions to quantum physics and his preference for abstract mathematics most thoroughly repudiates classical concepts and experience-based modes of representation. Heisenberg’s matrix mechanics tosses out all the old tools, sacrificing much that we take for granted about the material world and radically undermining our conventional means of orienting our world in terms of the opposition between observing subject and observed object. It makes sense, then, that his metaphorical journey is a more fraught one, and the path to success much less clear. If Bohr and Schrödinger encounter “pitfalls” and “barriers” along the way, they at least appear to have faith that their path advances them in a forward direction or, in Schrödinger’s case, that they can traverse difficult terrain. All Heisenberg offers both the world of physics and his society is a clear-eyed view of our current disorientation, and the promise that at some point, if we are willing to face the loss of our traditional signposts, we will one day find the means to reorient ourselves. FOUNDATIONS AND FRAMEWORKS

Bohr, Heisenberg, and Schrödinger also use architectural metaphors to narrate the relationship between classical and quantum physics, making use of both the THINKING IS BUILDING/FORMING/SHAPING and THE COMPATIBILITY OF IDEAS IS THE COMPATIBILITY OF SHAPES image schemas. The first of these schemas surrounds the presentation of quantum physics as a threat to the “foundation” laid down by the collective concepts of classical physics. In describing the relationship between the findings of the new physics and classical physics, for example, Bohr writes, “the great extension of our experience in recent years has brought to light the insufficiency of our simple mechanical conceptions and, as a consequence, has shaken the foundations on which the customary interpretation of observations was based, thus throwing new light on old philosophical problems.”93

62

Chapter 1

This is true, says Bohr, of both the revolutionary “space-time continuum” introduced by relativity, and the rejection of causality in quantum physics.94 Later, Bohr observes that Planck’s discovery that the quantum of action “forms[s] the background of the ordering of our experience … and has brought about a complete revision of the foundations underlying our description of natural phenomena,” and then again, that “the new knowledge has shaken the foundations underlying the building up of concepts, on which not only the classical description of physics rests but also our ordinary mode of thinking [emphasis in original].”95 Here, thinking is not building upon something, but rather a form of “undermining,” a process that tends to occur when the subject concerns wide-reaching concepts such as spacetime, or far-reaching concepts such as the philosophical questions to which Bohr refers. Heisenberg similarly uses the foundation metaphor to depict the relationship between quantum and classical physics when he writes, “Planck must have realized that his formula touched the foundation of our description of nature, and that these foundations would one day start to move from their traditional present location toward a new and as yet unknown position of stability.”96 Reflecting on how the quantum interpretation replaced facts about the objective world of physics with possibilities or potentialities, Heisenberg writes, “Einstein was not prepared to let us do what, to him, amounted to pulling the ground from under his feet.”97 Schrödinger in turn observes, this time in reference to the determinism of classical physics, that “it was said and is said that ‘all the foundations would be lost, that without a determinist background our view of nature would become wholly chaotic.’”98 In all of these cases, a temporal relation is once again cast as a spatial relation, wherein the existing foundation is associated with the past, and the traditional with the already established. All three men associate quantum physics’ threat to the existing classical foundation with a radical conceptual shift; rather than seeing quantum physics as strengthening or extending the bedrock of a theory that has stood for centuries, they see it as completely dismantling it. Predictably, it is Heisenberg who, in his call for building a new foundation in an entirely new location, accepts as inevitable the period of instability that will follow from abolishing the classical concepts. From the foundation of a building emerges the frame upon which all else will hang, and it is in relation to the metaphor of framework that the

Experience, Perception, and the Limits of Language

63

second image schema, THE COMPATIBILITY OF IDEAS IS THE COMPATIBILITY OF SHAPES, emerges. Both Bohr and Heisenberg use the metaphor of frame or framework, as they do with foundation, to describe a relationship between classical, sense-based physical concepts and quantum physics. Architectural references to “frameworks” differ from those that refer to foundations, however, and framework metaphors are more often associated with modifications than with beginning anew. This less radical transformation appears in a passage that I quoted earlier in this chapter, where Bohr acknowledges “[w]e must, in fact, realize that the unambiguous interpretation of any measurement must be essentially framed in terms of the classical physical theories, and we may say that in this sense the language of Newton and Maxwell will remain the language of physicists for all time.”99 For Bohr, embracing complementarity—whether it be between two apparently mutually exclusive physical theories, or between mutually exclusive experimental results within one area of physics—is to expand the frame of reference. It is not to reject the other theory or experimental results, but rather to unite them by “provid[ing] a frame wide enough to embrace the account of fundamental regularities of nature.”100 In typical fashion, Bohr defines his goal in broad terms, and sets out to reconcile and render consistent apparent incompatibilities, rather than let them stand as inherently discordant. In his 1955 piece “Atoms and Human Knowledge,” Bohr writes of the endeavor “[t]o achieve a harmonious comprehension of ever wider aspects of our situation, recognizing that no experience is definable without a logical frame and that any apparent disharmony can be removed only by an appropriate widening of the conceptual framework.”101 The resource that Bohr has in mind is language; however, his goal is not to transcend existing language, for he sees everyday language as the only resource available for communicating quantum behavior: Indeed, the development of atomic physics has taught us how, without leaving common language, it is possible to create a framework sufficiently wide for an exhaustive description of new experience. In this connection, it is imperative to realize that in every account of physical experience one must describe both experimental conditions and observations by the same means of communication as one used in classical physics.102

64

Chapter 1

Bohr’s approach is to preserve the classical scaffolding, but to adopt a more expansive orientation to it, so that the classical concepts are retained, but now stand in a new relationship to one another. Bohr admits that, in the quantum interpretation, classical concepts such as wave and particle are fundamentally incompatible—that it is not possible to “occupy” these concepts simultaneously—yet he also believes that they remain indispensable to a full description of nature. Similarly, he seeks to reconcile theories of matter that appear inconsistent with the quantum interpretation. It is from this perspective that Bohr gives an account of the relationship between relativity and quantum physics: “no content can be grasped without a formal frame and … any form, however useful it has hitherto proved, may be found to be too narrow to comprehend new experience.”103 At several points in his career, Bohr emphasized the necessity of “widening the frame,” without dismantling it, and he does so again here when he advocates a complementary approach that encompasses all of the experimental evidence. Since tearing down the framework of classical concepts is not possible, the only alternative is to keep it, but to revise one’s understanding of the manner in which the parts are oriented toward one another—to revise one’s viewpoint. Like the foundation metaphor, the framework metaphor suggests an antagonistic relationship between the old and the new theory—however, with a slightly different emphasis. If the “foundation” of the classical concepts must be destroyed to make way for a new conceptual base, the classical “framework” appears to exist more as a limiting structure into which quantum physics cannot “fit.”Thus, while the existing foundation can be replaced, the existing framework seem to frustrate attempts to generate new axioms.While this may appear contradictory, it holds together if one considers the fact that when the authors make reference to foundations, they tend to focus on quantum theory’s sweeping challenge to fundamental truths such as spacetime, causality, and determinism. When they speak of frames, on the other hand, they tend to be writing about individual physical laws.When they use frameworks, they are acknowledging that there exists much within the Newtonian laws of physics that cannot be jettisoned, but that resists seamless integration with the quantum model. Instead of attempting to replace such key aspects of the classical conceptual framework that remain indispensable—either in practical

Experience, Perception, and the Limits of Language

65

applications or in terms of their explanatory value—sometimes a compromise must be fashioned, most notably in Bohr’s Principle of Complementarity, which attempts to construct a new relationship between the classical terms instead of replacing them. A curious artifact emerges from the journey versus architecture metaphors as Bohr and Schrödinger use them: while journey metaphors tend to express an enabling and generally supportive relationship between the old and the new physics (or between the physicist and the layperson), architecture metaphors tend to construct this relationship as antagonistic. It stands to reason that journey metaphors are associated with connection—they are more relational, continuous, and progressive in nature—one can begin with the old ideas and move forward to new terrain, or begin alone and move across the landscape toward others at different locations. In many cases, one travels—at least to begin with—along a path that has been cleared by those who came before. Radical paradigm shifts, however, require entirely new conceptual structures; where the old must be preserved, it can result in ill-fitting additions or limiting and partial modifications to the old architecture. Not surprisingly, given his more forceful and generalized rejection of the classical concepts, Heisenberg is somewhat of an outlier here: both his journey and his architectural metaphors—including both foundation and framework metaphors—suggest a more radical break with existing models. In fact, Heisenberg’s ship metaphor is as much architectural as it is about travel, and it is in the architectural component of the metaphorical conceit that he expresses the need to break with the old concepts: only by recognizing the misleading influence of these concepts can the captain hope to discover a new method that allows him to break free from this influence, or at least to discover that the old method can no longer serve him. Heisenberg’s framework metaphor is also more radical than the others’—he alters the conventional raw material of description in language, and replaces it with mathematics. Whether it’s Schrödinger, Bohr, or Heisenberg, however, their respective use of the journey and architecture metaphors reveals the manner in which they not only understand the complex relationship between the old and the new, but also the manner in which they believe this relationship ought to unfold.

66

Chapter 1 CONCLUSION

Historically, quantum physics has proved to be as much a communicative experiment as a material one, and as much about language as about mathematics. Because the paradigm used by Conceptual Metaphor Theory so closely accords with the conceptual paradigms of Schrödinger, Bohr, and Heisenberg, I have used it as a blueprint for analyzing how they viewed the relationship between embodiment, sense perception, and quantum physics, and to explain the challenges that they understood to be inherent to the use of language in communicating quantum behavior. I have discussed how, while classical physics was entity based, the concept of an entity does not hold up in quantum physics, where the probable knowledge of things replaces the observation of things, and tendencies toward being replace being itself. A statistical probability cannot be placed and has no clear boundaries or an exact trajectory; any effort to describe it must confront the bias of Indo-European language. I have argued that even if one turns to terms such as “events” or “happenings”—alternative, more “verb-based” words that are offered as more appropriate to quantum phenomena—new, agency-driven things come into existence.While I acknowledged the fact that quantum physics exceeds the limits of nominal language, I questioned Papin’s association of quantum phenomena with the more verb-based concept of events by reflecting on Heisenberg’s own observations that the concept of an event or happening cannot be sustained where quantum physics is concerned, given the fact that quantum physics undermines the causality and temporal/spatial relations upon which the notion of an event or a happening is based. I have contended that, however unsuitable noun-based language may be to quantum physics, nominal language and the relationship between the macrocosmic and the microcosmic realms are inescapable, a fact demonstrated by the way that noun-based concepts such as “wave,” “particle,” and “states” remain trapped in its logic. I have traced the conceptual relationship between the classical physics of Newton, and early twentieth-century quantum physics, drawing from several image schemas—defined in Conceptual Metaphor Theory as patterns of perception that emerge from our everyday experience. I considered first the OBJECT image schema to illustrate how, in his conceptualization of our interaction with the material world, Newton linked

Experience, Perception, and the Limits of Language

67

embodied experience, entity-based sense perception, and the concepts that we are capable of forming, and how he believed that these associations extended to the microcosmic atomic realm. I then contrasted Newton’s object-based classical physics to Heisenberg’s Uncertainty Principle, arguing that the Uncertainty Principle is so conceptually unintuitive because, in the behavior that it describes, the concept of an “entity” fails. I used the SOURCE-PATH-GOAL image schema to explore the concept of quantum jumps which, I argued, is so challenging because it precludes knowledge of the “path” that an electron takes, and thus leaves a gap between the “source” and the “goal.” I concluded that we cannot fully apprehend quantum jumps because the notion of quantum jumps precludes the continuous existence of objects that we assume to be a feature of our external environment. Both Schrödinger and Bohr valued continuity in their atomic theories—Schrödinger, with his continuity-based wave mechanics, and Bohr with his harmonizing Principle of Complementarity, and both, in their retention of classical terms and classical and quantum concepts, also valued a sense of cohesion between different theories of the material world, the past and the present, and scientists and laypeople. Continuity between the macrocosmic and the microcosmic proved difficult to sustain, however. The problem that these two physicists encountered lay in the fact that language derives from our macrocosmic sense perception and the embodied experience, while quantum events, in their disruption of spacetime and causality, defy everyday sense perception. As a result, efforts to use anthropomorphizing metaphors are bound to reflect a conceptual drift. Heisenberg, with his emphasis on abstract algebra over experiencebased evidence, remained warier of language and often appeared to view the inevitable conceptual drift that language imported as an unacceptable consequence of straying from more exact mathematics. In chapter 2, I demonstrate how Schrödinger, Bohr, and Heisenberg’s understanding of the relationship between experiment/experience, observation/perception, and the macroscopic versus microscopic realm had a profound effect on their respective opinions about the relative value of visualizability, the role of intuition in science, and their aesthetic preferences. The value of visualizability would prove to be a central point of dissent between Schrödinger and Heisenberg in particular (with Wolfgang Pauli both inspiring and contributing to Heisenberg’s

68

Chapter 1

position). I pursue my examination of science rhetoric when I consider a second battle between Schrödinger and Heisenberg, this time over the relative importance of presenting science in an intuitive form, and I explore in particular the rhetorical strategies that Heisenberg used to justify his matrix model as intuitive. In discussing Bohr, I focus on how his symbolic approach to nature implies a unique approach to the relationship between science and intuition, as well as the “aesthetics of science.” Throughout, I show how each man’s approach to each of these concerns emerged from and influenced the characteristic principles that each applied in interpreting the material world.

2 THE PHYSICS OF VISUALITY, INTUITION, AND AESTHETICS

INTRODUCTION

In The Conceptual Development of Quantum Physics, Max Jammer observes that “certain philosophical ideas of the late nineteenth century not only prepared the intellectual climate for, but contributed decisively to, the formation of the new conceptions of the modern quantum theory”— specifically, “contingentism, existentialism, pragmatism, and logical empiricism,” which, argues Jammer “united in rejecting causality though on different grounds” and “prepared, so to speak, the philosophical background for modern quantum mechanics.”1 Regarding this same rejection of causality in physics in Weimar Germany—a rejection central to the quantum interpretation—Paul Forman argues that “substantive problems in atomic physics played only a secondary role in the genesis of this acausal persuasion, [and] that the most important factor was the socialintellectual pressure exerted upon the physicists as members of the German academic community.”2 Along these same lines, in an interview with Einstein in 1932, James Gardner Murphy said, “scientists live in the world just like other people. … They cannot escape the influence of the milieu in which they live. And that milieu at the present time is characterized largely by a struggle to get rid of the causal chain in which the world has entangled itself.”3 In all of these cases, the development of quantum physics is attributed to the zeitgeist of the time, with the critics adopting a “hypodermic” model wherein intellectual currents have a direct influence on scientific discovery. The ascendancy of quantum physics might also be attributed to the fact that throughout the 1920s and 1930s, Niels Bohr’s Institute for Theoretical Physics in Copenhagen, along with Göttingen University,

70

Chapter 2

rose to be the two main powerhouses in theoretical physics, and it was out of these two institutions that quantum physics was born. The sheer concentration of intellect and the vast academic influence commanded by these two establishments was unprecedented, and their close alliance only served to enhance the impact that they had on the world of physics. This kind of coordinated effort contributed to the fact that the quantum interpretation advanced by the Copenhagen-Göttingen group became the canonical interpretation of atomic behavior. The Copenhagen-Göttingen group did have some challengers, represented most notably in the debates between Einstein and Bohr, which, as Olival Freire writes, “produced images and text featuring the two giants quarreling [that] are now iconic in the culture of Physics.”4 Still, as Freire also points out, the “almost unchallenged monocracy of the Copenhagen School” persisted up until the 1950s.5 Neither the historic collaboration between the greatest luminaries of physics, nor the influence of the dominant intellectual currents of the time tell the whole story, however—a story that is best told as a series of arguments about the seemingly unscientific qualities that a theory had to possess for it to be intelligible, or even imaginable. By 1925, Born and Jordan’s matrix mechanics had emerged as the most promising model for describing the behavior of atomic matter, and represented a major coup for Heisenberg, Pauli, and Born’s quantum interpretation. A significant challenge was about to arise, however, in early December of 1926, when Schrödinger published an equally groundbreaking paper: “An Undulatory Theory of the Mechanics of Atoms and Molecules.”6 In this paper Schrödinger first presented his wavefunction equation and laid out the principles of his wave mechanical interpretation. Suddenly, a persuasive and more accessible challenge to matrix mechanics had come onto the scene. The two formulations were on one level fundamentally at odds with one another; while wave mechanics preserved the classical vision of matter as continuous over time and space, matrix mechanics defied this conventional understanding of time and space and was instead founded on a discontinuous theory of matter. Heisenberg, Born, and Pauli found themselves at a distinct disadvantage, but like Schrödinger they remained convinced that their interpretation would prove to be definitive. What followed was a bitter struggle over which model would prevail, a struggle that revolved around two characteristics where Schrödinger’s model

The Physics of Visuality, Intuition, and Aesthetics

71

appeared to have a clear advantage: the ease with which its depiction of atomic matter could be visualized, and its responsiveness to ordinary intuition. In chapter 1, I examined how quantum theory inspired, from its inception, investigations into the nature, function, and origins of language— particularly as it relates to the relationship between human perceptual experience and the limitations of language. Even more fundamental than language is our capacity for visualization, which plays a key role in orienting us in space and, following from this, our initial concept formation. Any theory of the material world that is visualizable will appeal to our basic sense perception; any theory that is unvisualizable will be received as foreign and remote. The term most often used by those involved in the visualization debates was anschaulichkeit—typically translated as “visualizability.” In this chapter, I use anschaulichkeit and “visualizability” interchangeably; however, I frequently pair anschaulichkeit with the word “imperative”—as in, “the anschaulichkeit imperative”—for, in the quantum debates and in critical analyses of these debates, the term anschaulichkeit always appears in tandem with the question of whether visualizability was and is a precondition for any legitimate model of the physical world. In the first section of this chapter, Anschaulichkeit: The Visualization Debate, I offer an account of the visualization debate, but my main purpose is to challenge the common critical assumption that Heisenberg entirely dismissed the imperative that an atomic theory be anschaulichkeit, or that he saw this imperative as altogether backward-looking. Instead, I argue that despite his public repudiation of visualizability, Heisenberg’s preoccupation with how far the quantum interpretation fell short of the demands of anschaulichkeit was a central motivation for how he revised matrix mechanics—including the introduction of quantum jumps and the Uncertainty Principle for which he is most remembered. In a similarly historicized account, I extend and revise Mara Beller’s argument that key quantum postulates emerged from efforts to make the quantum interpretation more anschaulichkeit. I focus in particular on the tension between the private efforts of Heisenberg, Pauli, and Born to generate a more anschaulichkeit version of matrix mechanics—efforts that included Born’s distortionary co-optation of Schrödinger’s wavefunction in order to create the theory of probability—and their public efforts to discredit

72

Chapter 2

Schrödinger by dismissing anschaulichkeit as irrelevant and anachronistic. Throughout, I foreground the rhetorical nature of the anschaulichkeit debate, focusing on how the concept was variously invoked, accommodated, dismissed, and massaged. I conclude that it is out of these rhetorical battles around the concept of anschaulichkeit, as well as the piecemeal concessions made to the anschaulichkeit imperative, that quantum theory was born. In my second section, Anschauung: The Birth and Rebirth of a Concept, I describe how the struggle over the meaning and value of anschauung—typically translated as “intuition”—generated a similarly heated debate within the community of atomic physics and, like the concept of anschaulichkeit, proved to be a preoccupation of Heisenberg’s, and to a lesser extent Pauli’s. Like anschaulichkeit, anschauung was key to the struggle between Schrödinger and the quantum camp represented by Bohr, Pauli, Born, Dirac, and Heisenberg, and as happened with visualizability, these men wrote, spoke, and disagreed vehemently over how intuitive a legitimate theory of atomic matter had to be. I begin by arguing that it is vital to distinguish between the “anschaulichkeit battle” and the “anschauung battle” in order to understand how the terms were deployed by the physicists themselves. Only by distinguishing between the two concepts, as they circulate in the quantum debates, is it possible to appreciate how and why Pauli and Heisenberg’s strategy for addressing the anschauung imperative was so very different from how they addressed the anschaulichkeit imperative. Examining separately the different rhetorical strategies and epistemological commentaries around these two terms also enables me to offer a subtler interpretation of the relationship between the physical world and human experience, and a deeper understanding of what a debate over the proper meaning and use of a concept seemingly unrelated to science can reveal about the historical struggles over which scientific theory will prevail. I also examine anschauung in the context of Schrödinger’s wave model, which rests on the assumption of continuity between the microscopic behavior of atomic matter and our macrocosmic embodied experience of time and space, and which thus possesses the advantage of being more intuitive, as we typically experience intuition. I then connect Schrödinger’s emphasis on the continuity between the macrocosmic and microcosmic world to his emphasis on the continuity of scientific concepts across

The Physics of Visuality, Intuition, and Aesthetics

73

history—particularly his insistence that any intuitive theory of matter must remain intelligible across future scientific paradigms. Because Schrödinger’s model is the most open to customary intuition, I argue, he found himself able to devote his efforts to discrediting matrix mechanics, and he did so by employing a series of rhetorical strategies designed to make matrix mechanics appear unscientific. I proceed to analyze how Bohr’s “symbolic turn” represents a different relationship to anschauung, particularly in the emphasis that he places on the description of the entire “quantum situation,” which includes experimental arrangements, experimental evidence, observation, and the classical concepts. I point out that while the classical concepts remain central to Bohr’s interpretation, he does not retain Schrödinger’s notion of a straightforward relationship between concept and reality, nor does he envision the possibility, at least with respect to the quantum situation, of a direct link between our experience of the material world and the way we employ concepts to understand that world—a link upon which intuitiveness is founded. Instead, Bohr sees the classical concepts as symbolic idealizations, arguing that we must consider our language to be referencing the notion of a wave or particle rather than a “real” wave or particle, and that we must consider the observation of quantum behavior to be itself an idealized concept rather than a conduit for our absorption and processing of data.To the extent that in Bohr’s model we have access to concepts such as wave and particle, I conclude, we can apply customary anschauung, but to the extent that these concepts become a function of description rather than reality, we must reach beyond our usual ways of understanding intuition. Finally, I analyze Heisenberg, Pauli, and Born’s attempt to break entirely with the customary definition of anschauung, and I trace the sometimes-byzantine logic that Heisenberg follows in his effort to rid the concept of anschauung entirely of its conventional associations, and to redefine it in a manner that would put to rest accusations that his abstract matrix mechanics was hopelessly unanschauung.Through close readings of Heisenberg’s writing and interviews from the 1920s through to the 1960s, I unpack the discursive strategies that he uses to argue that the unintuitive aspects of his model reflect merely the limitations of the linguistic resources and modes of communication that are currently available. I conclude by considering Heisenberg’s discontinuity-based proposition

74

Chapter 2

that the customary definition of anschauung is historically contingent, and thus that the meaning of anschauung has the potential to evolve over time. In the final section of this chapter, Judgments of Taste and the Aesthetics of Physics, I explore how the aesthetic preferences and sensibilities of Schrödinger, Bohr, and Heisenberg parallel their theoretical and epistemological methodologies and priorities. Despite their differences, I argue, all three men express an aesthetic sensibility that accords the highest value to harmony and simplicity. Along these lines, I contend that Schrödinger’s praise for a form of art that eschews ornamentation and tolerates “blank spaces on the canvas” is in fact an implicit critique of the unnecessary elements that he believed were forced over time into the matrix model. I then argue that Bohr’s aesthetic sensibility is similarly based on continuity and the unification of knowledge and that while Bohr may have expressed fascination over the multiple perspectives offered by cubist art, he ultimately is motivated by a logic of reconciliation that seeks harmony in difference. Finally, I turn to Heisenberg who, of the three men, spent the most time reflecting upon the relationship between knowledge and aesthetics. I trace the manner in which Heisenberg draws from the Pythagorean and Platonic traditions in order to establish simultaneously the beauty and truth of mathematics, and how he further elevates mathematics by praising its unifying nature. ANSCHAULICHKEIT: THE VISUALIZATION DEBATE

Within a few months the Göttingen physicists had replaced the goal of an unvisualizable deterministic theory with a visualizable statistical one. —Mara Beller

In “Visualization Lost and Regained,” Arthur I. Miller observes: The path to the quantum theory of 1927 was not an orderly progression from visualizable models to a mathematical formalism whose description of matter and phenomena in the atomic domain defied visualization in the ordinary sense of the word. Rather, the situation was closer to the one recalled by Werner Heisenberg where the physicists experienced despair and helplessness because of their loss of visualization.7

The Physics of Visuality, Intuition, and Aesthetics

75

In reality, the “despair” to which Miller refers was felt primarily by Heisenberg, and emanated from the following: Heisenberg’s mathematics abandoned the foundational link between sense-based experience, visualizability, and representation, yet Heisenberg also recognized the indispensability of the visualizable concepts such as “wave” and “state” that were more closely associated with Schrödinger’s model.8 In Scientific Explanation and Atomic Physics, Edward MacKinnon argues that the struggle over whose theory would prevail was a result of Schrödinger’s need to compete with Heisenberg. MacKinnon is right to identify competition as a primary factor in the battle between matrix and wave mechanics; however, he gets the relationship between the two men and their theories wrong. Schrödinger’s wave mechanics theory was not a reaction to Heisenberg’s early matrix mechanics. Schrödinger initially had no reason to feel either despair or insecurity. His wave mechanics model, which was based on Louis de Broglie’s 1924 hypothesis that all matter had wave-like qualities, retained a close relationship to the classical concepts and was therefore easily visualizable.9 As Henk de Regt puts it,“in classical physics, Anschaulichkeit is unproblematic because all classical theories are visualisable, in the sense that they are formulated in the same spatio-temporal framework as our visual experiences.”10 Schrödinger’s model retained this spatio-temporal framework; Heisenberg’s did not. In fact, it was the success of Schrödinger’s wave mechanics—which initially, and for some time, was received enthusiastically by the physics community—that posed a threat to Heisenberg’s matrix mechanics model and that left him in a tactically defensive position.11 The publication of Schrödinger’s wave mechanics model was also the driving force behind Heisenberg’s contradictory efforts to, on the one hand, develop a more visually friendly version of his matrix mechanics, and, on the other hand, publically discredit the “vulgar” visuality of Schrödinger’s wave mechanics. Of those critics who explore the visualization debate that occurred prior to the canonization of the quantum interpretation, most oppose the final version of the quantum interpretation of Schrödinger’s visualizable wave mechanics, and represent the supporters of the quantum interpretation as persistently disdainful of the argument that a theory of the material world must be visualizable. In “Spacetime Visualisation and the Intelligibility of Physical Theories,” for example, de Regt refers back to Pauli’s dismissiveness of anschaulichkeit as expressed in a letter that Pauli

76

Chapter 2

sent Bohr on December 12, 1924, where Pauli wrote, “I consider this certain—despite our good friend Kramers and his colourful picture books—‘and the children, they love to listen.’ Even though the demand of these children for Anschaulichkeit is partly a legitimate and a healthy one, still this demand should never count in physics as an argument for the retention of fixed conceptual systems.”12 According to de Regt, Pauli’s near-contempt for the requirement of visualizability had a significant impact on Heisenberg’s own rejection of visualization. In similar fashion, Paul Forman associates Heisenberg’s matrix mechanics with a rejection of visualizability, arguing that “the program behind matrix mechanics was to create an atomic dynamic free of any … pictorial, elements,” and he casts Heisenberg’s matrix theory as the culmination of his “emphatic commitment to give up pictorial atomic models.”13 Miller further entrenches this origin story when he asserts that Heisenberg’s insights, “driven by mathematical formalism, actually relied on the rejection of visualization as a meaningful category,” so that “the loss of visualizability turned out to be an essential prerequisite for Werner Heisenberg’s formulation of the new quantum mechanics in the 1925 publication” (of “Quantum-theoretical Reinterpretation of Kinematic and Mechanical Relations”).14 Part of the reason that Miller forges an essential link between quantum theory and unvisualizability concerns his focus on Heisenberg’s 1925 paper, which did not represent the mature form of matrix mechanics, and lacked the concessions to visualizability contained in the more mature iteration of matrix mechanics. Heisenberg’s rhetorical strategy was indeed to reject visualizability, in an effort to discredit Schrödinger’s wave mechanics and promote his matrix mechanics. Like de Regt, however, Miller posits as too essential a link between matrix mechanics and the absence of visualizability, and thus misses the important concessions to visualizability that were made in the later, more robust quantum interpretation. An even more essentializing account of the relative extent to which Heisenberg, Schrödinger, and Bohr’s individual approaches were visualizable associates their respective models with the innate cognitive abilities of each man, who was designated as either a “visual thinker” or a “nonvisual thinker.” Such an account is offered by de Regt, Miller, and Walter J. Moore. According to Miller, “[w]hereas for Bohr the loss of visualization

The Physics of Visuality, Intuition, and Aesthetics

77

was painful, Heisenberg found it to be congenial to his nonvisual mode of thinking. … The loss of visualization must have been especially difficult for Bohr whose essays are filled with visual words—for example, picture [Bild], visual ideas [Vorstellungen vor Augen], mechanical models.”15 With reference to Bohr, Einstein, and Poincaré, Miller states, “in their research the boundaries between disciplines are often dissolved and they proceed neither deductively through logic nor inductively through the exclusive use of empirical data, but by visual thinking and aesthetics.”16 At yet another point, Miller writes: Bohr, by late 1926, had accepted the duality in the quantum theory and its reflection in nature as the complete wave-particle duality, even though the wave aspect of matter had not yet been definitely established experimentally. For Bohr the wave-particle duality was the “central point in the whole story,” because it permitted him to use visual thinking once again, that is, to play with pictures of waves and particles.17

The problem here lies in the fact that Miller’s take on Bohr is too reductive; Bohr was not merely “playing with pictures of waves and particles,” and in fact Bohr saw any classical concepts as symbolic idealizations— alternating concepts necessary for a full description of experimental outcomes, but possessing of themselves no relationship to the sense perception associated with visualization. Like Miller, de Regt distinguishes between visualizers and nonvisualizers, and he suggests that the tendency to think one way or another is born of an innate quality: [M]any theoretical physicists have developed a familiarity with, and intuition for, the behaviour of the solutions of the equations they use, which enables them to have a feeling for the qualitative behaviour of the described systems without invoking picturable physical mechanisms. Physicists like Pauli, Heisenberg and Dirac clearly possessed this ability, and Pauli could persistently reject visualisable models because his capacity for abstract reasoning was so well developed.18

Throughout his article, de Regt reveals a clear preference for abstract, nonvisual thinking over more “concrete” visual thinking, and according to de Regt, Pauli’s rejection of visualizability as a legitimating factor for any model of the physical emerged not from his desire to promote matrix mechanics, but rather from a special mental aptitude. Moore similarly

78

Chapter 2

essentializes the opposition between visual and nonvisual thinkers, although he departs from Miller in positioning Bohr as a nonvisual thinker. As his reference point, Moore uses the 1926 Copenhagen debates between Bohr and Schrödinger. Based on his analysis of these debates, Moore concludes that “a basic reason for the failure of Bohr and Schrödinger to communicate more effectively [during these debates] was that their minds belonged to two different categories” with respect to the capacity for visualization.19 Moore proceeds to argue that “[i]n the analysis of human personalities devised by Francis Galton, Schrödinger was a ‘visualizer’ and Bohr was a ‘non visualizer,’ one thought in terms of images and the other in terms of abstractions, and it is virtually impossible for such twain to agree in any kind of discussion.”20 Moore’s point of reference is highly questionable—he fails to mention the fact that Galton was the originator and conceptual father of eugenics, and that Galton constructs the opposition between visualizers and nonvisualizers using that framework. In Inquiries into Human Faculty and Its Development, for example, Galton argues that “an over-ready perception of sharp mental pictures is antagonistic to the acquirement of habits of highlygeneralised and abstract thought, especially when the steps of reasoning are carried on by words as symbols, and that if the faculty of seeing the pictures was ever possessed by men who think hard, it is very apt to be lost.”21 He also expressed the conviction that visualizers are genetically inferior to nonvisualizers, which he supports in part by anecdotal evidence that scientists are more nonvisual, while the working class, women, and the majority of the “barbarian” races tend to be more concretely visual and are less able to construct abstract mental images.22 None of these assessments are correct, largely because they reduce each man’s theory of atomic behavior to either a clearly defined conceptual preference or an inherent cognitive ability. In opposing visualizers and nonvisualizers, Miller, de Regt, and Moore ignore the fact that Schrödinger’s commitment to visualizability was primarily epistemological and philosophical—he believed in the more visualizable classical concepts because he felt they were more intelligible and because he was committed to a theory of nature that could be communicated outside of mathematics. As to the assertion by both Moore and de Regt that Schrödinger possessed a more limited capacity for abstract thinking, one need only consider the fact that it was Schrödinger who demonstrated

The Physics of Visuality, Intuition, and Aesthetics

79

the mathematical equivalency of wave and matrix mechanics.23 Similarly, to position Bohr as either a visual or nonvisual thinker ignores his nuanced approach to experimental results, observation and description, and epistemology. Miller, de Regt, and Moore also miss the fact that Heisenberg’s relationship to visualizability was full of contradictions: he variously dismissed the argument that visualizability was an essential aspect of any description of nature, expressed anxiety about his theory’s lack of visualizability, and sought ways to reintroduce it in his later formulation of matrix mechanics. WAVE VERSUS MATRIX MECHANICS: THE ANSCHAULICHKEIT BATTLE

I am a professional theoretical physicist and I would like to make a clean theory. And when I look at quantum mechanics I see a dirty theory. —John Bell

In “Matrix Theory before Schrödinger: Philosophy, Problems, Consequences,” Mara Beller offers a very different origin story for the quantum interpretation, a more nuanced account of its relationship to visualizability, and a more historicized description of the role that Schrödinger’s wave mechanics played in launching Heisenberg, Pauli, and Born on the somewhat circuitous path that led to the final version of matrix mechanics. As she progresses through her argument, Beller exposes the extent to which Heisenberg’s uneasiness with the unvisualizability of his original formulation of matrix mechanics, and his subsequent revisions to it, emerged in response to the formal coherence and success of Schrödinger’s much more visualizable wave mechanics. In the process, Beller reveals that the defining aspects of quantum theory owe their existence to Heisenberg’s stubborn determination to give a coherent account of atomic behavior that could rival Schrödinger’s highly successful and widely accepted wave mechanics.24 On July 9, 1925, Heisenberg shared with Max Born what he called his “crazy paper,” in which he outlined a theory of observables that allowed for the existence of only select, mathematically permissible stationary states.25 This theory represented Heisenberg’s first effort to replace “the explanatory mode of continuous classical mechanics with a discrete descriptive approach that emphasized discontinuities” and that, as a result,

80

Chapter 2

represented a radical departure from the anschaulichkeit of classical physics.26 According to Beller, Heisenberg’s original formulation of matrix mechanics “not only implied that it was meaningless to measure motion as movement in time, but also that it was impossible to ascribe position to an electron at a given instant.”27 The rejection of concrete spacetime locators meant that matrix theory was entirely formal, and offered no way of giving a physical interpretation of subatomic behavior. Beller states that the elimination of the classical, continuum-based picture of the physical world “was aimed at … dispensing with visualizable models that relied on continuous space-time pictures.”28 Beller then observes that this elimination of spacetime pictures had “momentous consequences”: “carried to its logical conclusion … it meant eliminating the concept of a particle, or ‘thinghood,’ from the atomic domain [because] the concept of ‘thing,’ or particle, presupposes the ability to locate it in a definite point in space.”29 One of the formative moments for the quantum interpretation occurred not during these first formulations of matrix mechanics, but rather later, in January of 1926, when Schrödinger published his wave mechanics model in the highly regarded Annalen der Physik.30 Within a few months, Schrödinger had published three successive papers that solved for a host of atomic behaviors, including transitions between the stationary states of electrons. Schrödinger had provided a consistent model for the relationship between microscopic and macroscopic perspectives while maintaining continuity with classical physics and its visualizability. Of particular importance for the future development of Heisenberg’s quantum mechanical theory was Schrödinger’s publication of his third paper (“On the Relationship of the Heisenberg-Born-Jordan Quantum Mechanics to Mine”) in May 1926. In a footnote tacked on at the beginning of the paper, Schrödinger writes,“[m]y theory was inspired by L. de Broglie and by short but incomplete remarks by A. Einstein. No genetic relation whatever with Heisenberg is known to me. I knew of this theory, of course, but felt discouraged not to say repelled, by the methods of transcendental algebra, which appeared very difficult to me and by the lack of Anschaulichkeit.”31 Schrödinger proceeds to prove in this paper, step by step, that wave and matrix mechanics are mathematically equivalent—that, from “the formal mathematical endpoint, one might well speak of the identity of the two

The Physics of Visuality, Intuition, and Aesthetics

81

theories.”32 Schrödinger then sets out to demonstrate that it is only logical to replace Heisenberg’s abstract and unvisualizable discontinuum theory, along with its “peculiar calculating rules,” with his mathematically isomorphic but “perfectly clear” wave mechanics.33 Suddenly it seemed as though all of the hard work that Heisenberg, Born, Pauli, and Dirac had devoted to the development of matrix mechanics was for naught. A much later 1963 interview with Thomas Kuhn attests to the fact that the publication and wide acceptance of Schrödinger’s wave theory made Heisenberg anxious. Speaking to Kuhn, Heisenberg admits, I was so much afraid that by means of the Schrödinger mathematical scheme, a new interpretation of the thing would be brought in. Just because the [matrix] interpretation was not perfectly clear at that time, I was very much afraid that now entirely wrong ideas could enter into the thing and actually have entered. Schrödinger, as you know, wanted to throw all the quantum jumps away and to say that there is no quantization, it’s just all wave pictures and so on. I was so much afraid for this interpretation that I tried to avoid it from the very beginning.34

Heisenberg’s response at the time to Schrödinger’s attack on his “transcendent algebra” was swift and vehement: In his June 8, 1926 he issued the following retort: “The more I reflect on the physical portion of Schrödinger’s theory the more disgusting I find it. Just imagine the rotating electron whose charge is distributed over the entire space with axes in 4 or 5 dimensions. What Schrödinger writes on the visualizability of his theory … I consider trash.”35 Heisenberg’s dismissal of the value that Schrödinger placed on visualizability appears to prove that he thoroughly rejected the value of anschaulichkeit. As Beller demonstrates, however, the opposite is true; it was Schrödinger’s publications that provided the motivation for Heisenberg to strengthen his ties to visualizability. Directly after Schrödinger’s flurry of publications, the Copenhagen-Göttingen group began introducing a more visually friendly spacetime dimension by embracing a particulate model of electrons. The introduction of the more visualizable, particulate model is evident in the three defining postulates of quantum theory: Born’s probability theory, the resurrection of Bohr’s earlier theory of “quantum jumps,” and Heisenberg’s Uncertainty Principle.

82

Chapter 2

Part of the Copenhagen-Göttingen group’s anxiety grew out of the fact that they needed to respond to Schrödinger’s straightforward account of stationary states. In his 1926 paper “On the Quantum Mechanics of Collisions,” Born claimed to have found a solution to the stationary state transition problem. In the 1926 article in which Born introduces probability, Born begins with a summary of the wave-based formula that de Broglie used to calculate wavelength frequencies, then writes, “If one translates this result into terms of particles, only one interpretation is possible.”36 Born’s solution was to turn on its head Schrödinger’s understanding of “waves” as physical manifestations, and by using Schrödinger’s own mathematical schema to reinterpret these waves as undefined fields that reflected the statistical likelihood that a particle would be found to be in a particular stationary state. Ironically, then, Born felt that Schrödinger’s wave model was best suited for his investigation of describing collisions, and even attested that “exactly for this reason I might regard it as the deepest formulation of the quantum laws.”37 However, Born’s use of Schrödinger and de Broglie’s wave model as a point of departure did not reflect, as Arthur I. Miller claims, a situation in which Born’s “desire for visualization led him to enlarge his aesthetic from the one held currently by Bohr and Heisenberg to a viewpoint in 1926 in which particles are guided by waves from Schrödinger’s theory.”38 Miller’s focus on Born’s point of departure (wave mechanics) leads him to ignore the import of what Born proceeds to do to wave mechanics—namely, his strategic transformation of Schrödinger’s physical theory into an abstract theory of fields of probability. Born’s reinterpretation had tremendous strategic value: it both gutted the central, physical, visualizable aspect of Schrödinger’s wave theory, and provided the Göttingen group with an account of stationary states that was consistent with matrix mechanics. The invention of a visualizable particle was a necessary condition of probability theory, and the cornerstone of Born’s coup for the quantum interpretation, and despite the fact that the probability theory precluded “seeing” precisely where an electron was, it retained a relation to visualizable space because, as Beller argues, it “presupposed that an electron does indeed occupy a definite position in space, and provided the formulae for estimating that position’s probability.”39

The Physics of Visuality, Intuition, and Aesthetics

83

Schrödinger was understandably distressed by Born’s radical reinterpretation of the wavefunction, as well as by Heisenberg, Bohr, Hans Kramers, and John Slater’s support of probability waves. In his 1953 paper “The Significance of Wave Mechanics,” presented in Paris as part of the commemoration of de Broglie’s sixtieth birthday, Schrödinger stated that both he and de Broglie had been “shocked and disappointed that this transcendental, almost psychical interpretation of the wave phenomena had become the almost universally accepted dogma.”40 Schrödinger insisted that the wavefunction represented not an abstract estimation of the likelihood of a particle being “either here or there,” but rather a continuous physical region wherein “standing waves” with discrete frequencies emerged as perceptible phenomena.41 Throughout his career, Schrödinger took pains to emphasize that his wavefunction ought to be understood in terms of directly observable features—“physical densities in frequencies within an entire wave field— somewhat like the white crests in a choppy sea.”42 Schrödinger argued that the difference between defining the wavefunction as a perceptible frequency, and defining the wavefunction as a purely formal probability expressed “the momentous difference between ‘both-and’ (et-et) and ‘either-or’ (aut-aut).” According to Schrödinger, “if you accept the current probability view (aut-aut) in quantum mechanics,” then “the singleevent observation becomes comparatively easy to tackle, but all the rest of [classical] physics (unfashionable at the moment) is lost to sight.”43 Here, Schrödinger critiques the newfangled theory for being short-sighted and opportunistic. Beller’s argument that probability “presupposed that an electron does indeed occupy a definite position in space” requires a point of clarification. Schrödinger’s insistence on the material reality of his “standing waves” reminds us that while probability introduced the visualizable notion of a particle, the particle that probability implies must be seen as more symbolic and less “entity-like” than the standing wave that Schrödinger had in mind. In the same 1963 interview with Kuhn to which I referred before, Heisenberg muses about the nature of reality with respect to probability theory: What Bohr, Kramers, and Slater did was to establish the probability as a kind of reality, that is, that the probability is not something about just counting numbers … but that this expectation itself is like something real. … [N]ow

84

Chapter 2 the idea is that there is a wave. But this wave is not the reality. This wave is a probability—this wave is a tendency. … And it was the tendency that a fact should happen.That was the striking thing about it, you know, this new invention of a possibility which was a reality in some way but not a real reality—a half reality.44

The particle of probability theory does not have the same material reality in time and space as Schrödinger’s standing wave. Rather, the particle implied here is a “fact” waiting to happen—it possesses the potential to be seen, and it is this last phrase that captures how Heisenberg’s “half reality” ought to be understood in the context of visualization. The Copenhagen-Göttingen group still needed to account for the transitions between states, something that Schrödinger’s frequency-based standing waves had already done in a physically grounded, visualizable, and persuasive manner. Heisenberg wanted to offer an account of transitions between states that both challenged Schrödinger’s frequency model, and confirmed the Copenhagen-Göttingen probability interpretation of the wavefunction. He found it in Bohr’s earlier theory of quantum jumps, where electrons moved discontinuously from one stationary state to another. In his November 1926 paper, “Fluctuation Phenomena and Quantum Mechanics,” Heisenberg set out to demonstrate not only that probability emerged naturally from quantum mechanics (thus fully detaching probability theory from its origins in Schrödinger’s wave theory), but also that “a probability interpretation emerges naturally and can be understood only if there are quantum jumps.”45 Schrödinger found Heisenberg’s reintroduction of quantum jumps to be arbitrary and full of gaps and contradictions. He expressed his objections to quantum jumps most fully in the 1926 Copenhagen debates with Bohr. Toward the end of Heisenberg’s account of the debates in Physics and Beyond, he describes Schrödinger exclaiming, “[i]f all this damned quantum jumping were really here to stay, I should be sorry I ever got involved with quantum theory.”46 In part, Schrödinger objected to quantum jumps on the grounds that the isolated particles of quantum jumps remain purely hypothetical because they cannot be detected experimentally. In “Are There Quantum Jumps? Part II,” for example, Schrödinger writes, “we are not experimenting with single particles, any more than we can raise Ichthyosauria in the zoo. We are scrutinising records of events long after they have happened. … We can never reproduce the same

The Physics of Visuality, Intuition, and Aesthetics

85

single-particle-event under planned varied conditions; and this is the typical procedure of the experimenter [emphasis in original].”47 Although Schrödinger objects to the hypothetical nature of quantum jumps, his reference to them as “single-particle-events” reinforces the fact that quantum jumps were part of the movement over to a more particulate model. Of Schrödinger’s objections, Beller observes, “No wonder that Schrodinger was furious about the reintroduction of ‘these damned quantum jumps.’ Such offbeat concepts seemed alien not only to his theory but to the matrix approach, originally constructed so as to dispense with such illustrative notions.”48 While the record of quantum jumps cannot coincide, in space and time, with the “event,” the internal logic of the quantum jump aligned seamlessly with the new, particulate, and more visualizable matrix mechanics. On February 23, 1927, Heisenberg wrote a letter to Pauli from Copenhagen in which he first described his theory of the Uncertainty Principle.49 While the Uncertainty Principle introduces an irresolvable indeterminacy in the measurement of either the position or momentum of an electron, as Beller points out, it nevertheless “does not limit infinite precision of position when momentum remains undetermined.” If conjugate states such as position and momentum cannot be measured simultaneously (as in classical physics), it remains the fact that, as Bohr observed, the state of an electron can be definitively identified in either space or time. In “The Quantum Postulate and the Recent Development of Atomic Theory,” published in 1928, Bohr expresses in clear terms the fact that the Uncertainty Principle, while founded on indeterminacy, followed the logic of particles: the coordinates of a particle can be measured with any desired degree of accuracy by using, for example, an optical instrument, provided radiation of sufficiently short wave-length is used for illumination, just as “[t]he momentum of a particle … can be determined with any desired degree of accuracy …, provided the wave-length of the radiation is so large that the effect of recoil can be neglected.

This ability to determine position and momentum, even if only separately, is also predicated on a particulate model, offering a “mental picture of electrons.”50

86

Chapter 2

As with probability, it is necessary to consider precisely what the concept of a particle means, and how the Uncertainty Principle can be seen as supporting a particulate model. As I stated in chapter 1, because the Uncertainty Principle dictates that the state of the object can never be fully determined, it also dictates that this object can only possess a tendency toward being—a tendency that challenges the logic of embodied “entities.” How, then, can the Uncertainty Principle be seen as anything but a challenge to the notion of a “thing”? Well, once an electron can be either “here” or “there,” we return to the language of entities, and once we are trapped in the language of entities, we are inevitably trapped in “mental pictures.” To be sure, the Uncertainty Principle undermines our conventional notion of the object and reveals the inadequacy of an entitybased conceptual framework, and Heisenberg’s electron is less substantial, less tied to our embodied experience of the macrocosmic world than Schrödinger’s standing wave.The problem as I see it here is more rhetorical than conceptual. On the one hand, Heisenberg readily admitted that the language of entity was inescapable. In Physics and Philosophy, Heisenberg observes that “[o]ur perceptions are not primarily bundles of colors or sounds; what we perceive is already perceived as something, the accent here being on the word ‘thing.’”51 On the other hand, rhetorically he disavowed the anschaulichkeit that is the inevitable outcome of naming a particle as such. In his papers, letters, and statements, Schrödinger consistently demonstrated that epistemological and philosophical concerns were key to his commitment to visualizability, while Heisenberg’s comments and his revisions to matrix mechanics consistently were more strategic and reactionary, and inspired by his desire to “win out” over Schrödinger. While Schrödinger never diverged from his philosophically and epistemologically driven commitment to maintaining a physical, visualizable, and sense-based picture of time and space that could be expressed via classical concepts, Heisenberg demonstrated a more inconsistent, contradictory, and fraught relationship to the question of classical spacetime and visualization. This fraught relationship is suggested in the fact that while Schrödinger’s wave mechanics emerged in its more or less complete form in a matter of months, Heisenberg’s matrix mechanics developed, somewhat circuitously, over a number of years. To a significant degree, the slow progression of matrix mechanics was caused by Heisenberg and

The Physics of Visuality, Intuition, and Aesthetics

87

the Copenhagen-Göttingen group’s successive efforts to respond convincingly and coherently to the successes of Schrödinger’s wavefunction in order to establish the superiority of matrix mechanics. Schrödinger’s elaborations of wave mechanics in a quick succession of papers were just that—elaborations; Heisenberg’s and others’ revisions to matrix mechanics in response to Schrödinger’s wave mechanics fundamentally changed the structure of and premises informing the quantum interpretation. Heisenberg’s expression of contempt for the “disgusting” physical aspects of Schrödinger’s wave mechanics and his dismissal of the visualizable aspect of Schrödinger’s wave mechanics as “trash” again appear largely rhetorical—given Heisenberg’s repeatedly expressed anxiety about the nonvisualizability of his theory, followed by his shift from the early, unvisualizable iteration of matrix mechanics to a more visually friendly model. BOHR: OBSERVATION, OBJECT, AND OBJECTIVITY

Bohr’s stature was instrumental in legitimizing the quantum interpretation within the larger physics community. Certainly, Bohr both accepted and helped consolidate key postulates developed by the CopenhagenGöttingen group, including the discontinuous theory of matter that relied in part on his concept of quantum jumps. As far as the concept of anschaulichkeit is concerned, however, Bohr’s position is ambiguous and indirect. By 1925, Bohr appeared to have succumbed to the same “despair” that Heisenberg had expressed to Pauli over how the quantum postulates emptied of meaning terms such as “state,” “wave,” and “corpuscle,” and appeared to abandon hope that the newly described attributes of atomic matter could be visualized. After a discussion with Einstein in 1925, Bohr stated, “[i]n my opinion the possibility of obtaining a spacetime picture based on our usual concepts becomes ever more hopeless,” and in his 1925 article “Atomic Theory and Mechanics,” Bohr lamented that the quantum behavior of atomic matter represented “an essential failure of the pictures in space and time on which the description of natural phenomena has hitherto been based.”52 Bohr again seemed to give up on visualizability when, during the Copenhagen debates with Schrödinger, he accepted the “gaps” in our knowledge of electron behavior between quantum jumps, but stated that their unvisualizability did not preclude their existence. Rather, Bohr argued, “[i]t only proves that we

88

Chapter 2

cannot imagine them, that the representational concepts with which we describe events in daily life and experiments in classical physics are inadequate when it comes to describing quantum jumps. Nor should we be surprised to find it so, seeing that the processes involved are not the objects of direct experience [emphasis added].”53 Here, Bohr links mental pictures to everyday experience, and argues that the unproblematic link between image and object in classical physics does not extend to quantum phenomena because they occupy a different realm than that of everyday, macrocosmic perception. Like Heisenberg, Bohr rejected Schrödinger’s argument that any legitimate theory of atomic matter must be consistent with our embodied, sensate, everyday experience. Like Heisenberg, he believed that the uncertainty relations introduced an unavoidable ambiguity that overturned the classical assumption that the senses deliver a full and accurate visualizable record of atomic behavior. For Bohr, the uncertainty relations constituted “a direct expression of the absolute limitation of the applicability of visualizable conceptions in the description of nature.”54 With his suggestion here that visualisability is of limited relevance where quantum behavior is concerned, Bohr appears to fall into the same category as Heisenberg—both men seem to qualify as “nonvisual” thinkers. Bohr also shared much ground with Schrödinger, however, and Bohr’s extended reflections on the apparent materiality of observation, object, and experimental set-up—and the importance he ascribed to communicability and thus the necessary use of classical concepts—seemed to align with Schrödinger’s concerns. In the end, however, Bohr differed from both Schrödinger and Heisenberg in one important respect. Both Schrödinger and Heisenberg sought a theoretical model that could be reduced to one coherent set of concepts; both anticipated, or at least hoped, that their theory would eventually provide a complete picture of the behavior of matter. Heisenberg gave primacy to the algebra of his matrix mechanics, and then predicted that experimental evidence would eventually validate this theory. Schrödinger gave precedence to his physical wave mechanics, and argued that a theory must be compatible with experimental evidence. Both were concerned with the mechanics of atomic behavior. If for Schrödinger, and for Heisenberg, the goal was to develop a coherent a model of what happens inside the atom, for Bohr the goal was to discover how the conditions

The Physics of Visuality, Intuition, and Aesthetics

89

determined by sense perception define how we observe what happens inside the atom, and this is the context in which his relationship to visualizability must be seen. Bohr’s primary distinction lay in his philosophical approach, and he was driven in part by a desire to determine what the quantum postulates revealed about the nature of human perception and the limits that it imposed on observation. In “The Quantum Postulate and the Recent Developments in Atomic Physics,” Bohr states that the problematic relationship between the quantum postulates and observation originates in the fact that “[u]ltimately every observation can of course be reduced to our sense perceptions [emphasis in original].”55 As Dugald R. Murdoch points out, for Bohr, the primary aim of physics is to help us make sense of our perceptual experience. He does not doubt that our experience is of an independently existing physical world, but he denies that the main aim of physics is the comprehension of the imperceptible structure of that world for its own sake.We construct theories of the microstructure of the physical world in order better to comprehend the world in which we move and have our being—the macroscopic world.56

Bohr’s focus on perceptual experience had a very specific context, however: the observation of experimental results. Bohr notes, “[i]n tracing observations back to our sensations, once more regard has to be taken to the quantum postulate in connexion [sic] with the perception of the agency of observation, be it through its direct action upon the eye or by means of suitable auxiliaries such as photographic plates.”57 Bohr’s statements place his position on quantum jumps in a different context than Schrödinger’s; for Bohr, the gap in our knowledge concerning the transition between quantum jumps reveals not the limitation of quantum theory, but rather the limitations imposed by the conditions under which we are able to perceive traces left by objects. In his 1958 essay “Quantum Physics and Philosophy: Causality and Complementarity,” Bohr associates these conditions with the experimental set-up, an association that is central to his approach: it is also essential to remember that all unambiguous information concerning atomic objects is derived from the permanent marks … left on the bodies which define the experimental conditions. … The irreversible amplification effects on

90

Chapter 2 which the recording of the presence of atomic objects rests rather remind us of the essential irreversibility inherent in the very concept of observation [emphasis in original].58

Bohr argued that the experimental apparatus must be treated as a classical object occupying the macrocosmic domain of experience (one that “amplified” quantum effects to a macrocosmic scale), and that the resulting observations (of experimental evidence) can only be expressed using the limited terminology of a classical physics based on our customary, everyday, sense-based experience in the macroscopic world. Bohr observed in his 1937 paper “Causality and Complementarity,”“it is no longer possible sharply to distinguish between the autonomous behavior of a physical object and its inevitable interaction with other bodies serving as measuring instruments, the direct consideration of which is excluded by the very nature of the concept of observation in itself.”59 According to Bohr, a sharp distinction between experimental arrangement and perceived object inheres in the very definition of observation, and is in turn a precondition for concept formation. The uncertainty relations entail the failure of sense perception, defy observation, and block concept formation—all of which require a differentiation between subject and object. For Bohr, the necessity for a complementary description of atomic behavior emerged from the manner in which atomic behavior is revealed to our senses through the process of observation. Because he accepted the inescapable role that sense perception plays in our conceptualization of matter—both as a precondition of observation, and as the ground of classical physics—Bohr adopted a complementary approach to the description of matter. This approach represents Bohr’s effort to reconcile the inherently classical nature of observation, which requires a distinction between the object and the agency (apparatus) of measurement, and the fact that the interaction between apparatus and quantum systems does not allow for this distinction. Since both the experimental apparatus and observation can only be approached classically, a full description of nature requires that observations gained from different experimental arrangements be viewed in complementary fashion. Initially, matrix mechanics was broadly criticized for its abstract algebra, and outside of the Copenhagen-Göttingen circle, it was not favorably received, a situation that was no doubt exacerbated by the dense and

The Physics of Visuality, Intuition, and Aesthetics

91

often impenetrable writing style of Bohr. Schrödinger’s wave mechanics, in contrast, had wide appeal, precisely because in its physicality and continuity with ordinary experience it was more accessible and visualizable. As Miller observes, “wave mechanics delighted the more continuumbased portion of the physics community, who were intent on preserving pictures and clinging to a classical realism.” In 1926, for example, Dutch physicist Hendrik Lorentz wrote Schrödinger to say, “if I had to choose between your wave mechanics and the matrix mechanics, I would give preference to the former, owing to its greater visualizability.”60 Miller points out that even Max Born expressed a preference for Schrödinger’s wave mechanics because it allowed for the use of “conventional ideas of space and time in which events take place in a completely normal manner, that is, the possibility of visualization.”61 Thirty years after the breakthroughs that established quantum physics, Schrödinger still insisted, “I do not believe that we can be satisfied in the long run with the answer which I once received in conversation with a young physicist of outstanding genius [Paul Dirac]: ‘Beware of forming models or pictures at all!’”62 Later in the same book, Schrödinger concludes, “physics takes its start from everyday experience, which it continues by more subtle means. … Discoveries in physics cannot in themselves—so I believe—have the authority of forcing us to put an end to the habit of picturing the physical world as a reality.”63 While he continued to acknowledge the difficulties in visualizing quantum phenomena, he never gave up on his argument that his more visualizable wave model was superior to Heisenberg’s matrix algebra, and continued to maintain that experience and visualization were the ground from which physics should emerge. ANSCHAUUNG: THE BIRTH AND REBIRTH OF A CONCEPT

Schrödinger described quantum mechanics as a formal theory, of frightening, even repulsive un-intuitiveness. —Werner Heisenberg

Like the concept of anschaulichkeit, the concept of anschauung has special significance in the context of quantum theory, first, for those involved in the theory’s development, and then for those writing about it. Anschauung

92

Chapter 2

is usually translated as “intuition”; however, it can also be translated as “actual,” “the product of direct experience,” “sense-impressions,” “contemplation,”“view”“opinion,” and “notion.” In some cases the translation “intuition” is rejected because it is considered to be too vague or immaterial—particularly when it is used in the context of the quantum debates. In an editorial commentary titled “What Does Anschauung Mean?,” published by The Monist in 1892, the author finds “intuition” unsatisfactory both because it is too abstract, and because it suggests a kind of “mysticism or supernaturalism” that exceeds the German meaning of the word. The editor argues that the German word “anschauung” is vernacular in origin, and suggests an immediateness and directness not expressed in the English word “intuition.”64 As an alternative, he offers the Saxon-derived phrase “at-sight,” which better captures the direct experience and immediate “sense-impression” associated in particular with the Kantian notion of empirical intuition.65 The definition of anschauung as “at-sight” suggests a strong connection between anschauung and anschaulichkeit, a connection further emphasized by Arthur I. Miller in his analysis of the quantum interpretation. In his introduction to “Visualization Lost and Regained: The Genesis of the Quantum Theory in the Period 1913–1927,” Miller writes that his use of anschauung “will refer to intuition through the pictures constructed from previous visualizations of physical processes in the world of perceptions” because, he argues, this definition “best fits its intended meaning in the period 1925–27 by Bohr and Pauli, among others.”66 Care must be taken, however, not to associate anschauung too closely with anschaulichkeit, for the way that anschauung became associated with—or disassociated from— the different atomic theories involves very different sets of dynamics than did the anschaulichkeit debates. While favoring the conventional translation of anschauung as “intuition,” then, I argue that its meaning varies according to the context in which the concept is invoked. For this reason, I have found a comparative analysis of the different contexts to be most productive—one that distinguishes between how Schrödinger invokes the conventional notion of anschauung in his defense of his wave theory, how Heisenberg’s transforms it in his defense of his matrix mechanics, and how Bohr’s Principle of Complementarity draws on both the conventional notion of anschauung and suggests one that is more abstracted from direct experience.This comparative analysis reveals how the ways in

The Physics of Visuality, Intuition, and Aesthetics

93

which Schrödinger, Heisenberg, and Bohr characterize anschauung both reflect and reveal the structural, epistemological, philosophical, and ontological principles attached to their theories of atomic matter. SCHRÖDINGER: IN DEFENSE OF CUSTOMARY ANSCHAUUNG

Of natural objects, whose observed behavior one might treat, one sets up representation—based on the experimental data in one’s possession but without handcuffing the intuitive imagination. —Erwin Schrödinger

Because Schrödinger’s wave theory was physically grounded in our embodied experience of matter as continuous in time and space, his model is most consistent with customary intuition. Similarly, because Schrödinger’s model was continuous with both macrocosmic experience and classical concepts, and because his theory was widely accepted within the physics community, he did not have to challenge the fundamental assumptions informing existing physical theory, and so his main focus was on demonstrating that the Copenhagen-Göttingen model was untenable. For Schrödinger, quantum theory not only left gaps in our knowledge of the material world, it also relinquished the responsibility that scientists had to communicate with those outside the scientific community, at the same time that it detached itself from any connection to the history and future of scientific discovery. While the members of the CopenhagenGöttingen group were busy trying to resolve the internal contradictions of their theory, Schrödinger was searching for ways to discredit a theory of atomic matter that he believed was fundamentally flawed. In his 1928 paper “On the Relation Between the Quantum Mechanics of Heisenberg, Born, and Jordan, and that of Schrödinger,” Schrödinger defends wave mechanics with an argument that is both epistemological and rhetorical. First, he demonstrates the mathematical equivalence of wave and matrix mechanics—the fact that they “agree with one another with regard to the known facts.”67 Schrödinger proceeds to demonstrate that, mathematically, any problem that matrix mechanics can claim to have solved can be solved with more clarity using the classically and physically grounded wave equation.68 Since, observes Schrödinger, “the two theories are completely equivalent from the mathematical point of view, it can only be a question of the subordinate point of convenience.”69

94

Chapter 2

Given the fact that Schrödinger has already demonstrated mathematically that wave mechanics offers a clearer picture of atomic behavior and thus has already established that, in Henk de Regt’s words, wave mechanics is “heuristically more powerful,” it is clear which theory Schrödinger considers more “convenient.”70 Schrödinger then makes what might be considered a more ontological argument, with the claim that his theory is superior because it represents a continuity between the macrocosmic realm of experiment and the microcosmic realm of atomic matter. Schrödinger begins this portion of his argument by asserting that “the problems which the course of development of atomic dynamics brings up for consideration are presented to us by experimental physics in an eminently intuitive [anschauung] form.”71 He details a number of these problems, and then argues that, in order to solve them, it is necessary only to survey “clearly the transition between the macroscopic, perceptual mechanics and the micro-mechanics of the atom,” wherein “[m]icro-mechanics appears as a refinement of macromechanics.”72 Here, Schrödinger places special emphasis on the continuum from the macrocosmic to the microcosmic, a prerequisite for any classically intuitive theory of matter. Because the macroscopic world of experiment is “eminently intuitive,” and atomic behavior is presented via and in accordance with this macroscopic world, it too is eminently intuitive. Schrödinger proceeds to challenge the way in which the matrix model “attacks” anschauung, and in this portion of his argument, he adopts a rhetorical strategy that attaches to matrix mechanics words that are outside of the scientific register, in order to suggest that the founders of matrix mechanics have been motivated by agendas that are distinctly unscientific. This strategy is evident in the following passage, where Schrödinger writes, “it seems extraordinarily difficult to tackle problems of the above kind … as long as we feel obliged on epistemological grounds to repress intuition [anschauung zu unterdrücken] in atomic dynamics, and transition probabilities, energy levels, etc. [emphasis added].”73 Schrödinger distances matrix mechanics from properly scientific methodology by using the word “unterdrücken” which, along with “repress,” can also be translated as “oppress” or “quash,” and seems more to be in the moral than the scientific register.

The Physics of Visuality, Intuition, and Aesthetics

95

Schrödinger also observes, “there might perhaps appear to be a certain superiority in the matrix representation because, through its stifling of intuition (vollkommenen Unanschaulichkeit), it does not tempt (nicht dazu verleitet) us to form space-time pictures (bilder) of atomic processes, which must perhaps remain uncontrollable.”74 The moral register is clearly expressed in the term verleitet (“tempt” or “entice”), and is introduced again when Schrödinger contrasts the arbitrary and special “fleshly clothing” (fleischliche Umkleidung) of his theory to “the bare matrix skeleton” (kahlen Matrizenskeletts).75 These two passages suggest, with irony, that supporters of matrix mechanics see something “degenerate” and “excessive” about wave mechanics that is not shared by the Spartan and pristine matrix mechanics. Schrödinger’s irony here is also best captured when the phrase “vollkommenen Unanschaulichkeit” from the first of these two passages is translated more literally as “perfectly unvisualizable.” This translation suggests that we are not tempted because matrix mechanics has reached the pinnacle of inaccessibility and detachment. Finally, Schrödinger writes that matrix mechanics would indeed be superior from an epistemological point of view if it did not have to “pander to the need for visualizability” [anschaulichkeit fröhnende].76 The term “pander” suggests, again ironically, that matrix mechanics would somehow be “lowering” itself by acknowledging the importance of visualizability, when of course Schrödinger has repeatedly made clear, here and elsewhere, that visualizability and the openness to conventional intuition are key to any legitimate model. Because Schrödinger’s wave theory of matter has its physical grounding in our embodied experience of matter as continuous in time and space, assuming an unbroken line between the macrocosmic realm, the microcosmic realm, and classical concepts, it is available to the customary intuition that enables us to understand our world and our place within it. Assuming the consistency between his theory and our experience, he focuses his attention on the incongruities of the matrix model, and he both launches a conceptual critique and applies rhetorical strategies to undermine its legitimacy.While not averse to language play in making his point, his criticisms are genuine: he believes that quantum theory not only leaves unacceptable gaps in our knowledge of the material world, but also hinders our ability to apply our given faculties with the goal of understanding the material world. Of Schrödinger, Bohr, Heisenberg,

96

Chapter 2

Pauli, and Born, Schrödinger stands alone in the extent to which he associates anschauung with anschaulichkeit, an association consistent with the reciprocity that Schrödinger sees between anschaulichkeit and any properly accessible model of atomic behavior. One cannot ascribe Schrödinger’s view to the other men, however, because to do so fails to appreciate the very individual way in which all of these men approached the notion of anschauung. BOHR, ANSCHAUUNG, AND THE SYMBOLIC TURN

Thoughts without content are empty, intuitions without concepts blind. —Immanuel Kant

To understand Bohr’s relationship to anschauung requires approaching from a number of perspectives the singular manner in which he positions classical concepts. Bohr’s thoughts on the quantum interpretation are distinguished by their range, which spans his analyses of the interaction of macrocosmic and microscopic realms; his reflections on the connection between ordinary experience, human perception, and the material world; his theories about the conditions that define description; and finally, his philosophical reflections on the origins and status of scientific concepts. Bohr’s relationship to classical concepts must therefore be considered in relation to a cluster of determinate elements that defined the quantum situation: experimental set-up, material object, observation, definition, and description. Bohr’s notion of complementarity represented an effort to reconcile the intuitive classical concepts with the unintuitive experimental evidence. According to Bohr, at every stage—equipment, perception/ observation, the ontological status of the object—we cannot escape a classical relationship to matter; in Michael Cuffaro’s words, “[t]he experimental apparatus (a voltmeter, say) is always a piece of classical equipment which communicates classical information about what we assume to be (using classical criteria) an independently existing object [emphasis in original].”77 For Bohr, the instruments of measurement were inherently bound to the terms and concepts associated with the macrocosmic realm of ordinary experience and sense perception, because the instruments themselves were part of that realm.

The Physics of Visuality, Intuition, and Aesthetics

97

At the same time, Bohr did not share Schrödinger’s belief that there was a direct correspondence between reality and the representation of reality; rather, Bohr insisted that “an independent reality in the ordinary physical sense can neither be ascribed to the phenomena nor to the agencies of observation,” and he had no special commitment to the classical concepts beyond their symbolic content.78 Bohr’s Principle of Complementarity constitutes the “unification” of classical concepts, but his stress on the “logical compatibility of apparently contradictory laws” emphasizes the fact that this union is no longer between nature and its corresponding concept, but rather between two concepts of nature—in this case, the concept of a “particle” and the concept of a “wave.”79 To see the world from Bohr’s perspective is to understand that the atomic domain cannot be open to customary anschauung in an unproblematic or straightforward way, because in this domain the classical concepts signify in abstract relation to one another, and knowledge of quantum behavior is already mediated by a process of symbolic representation. It is from this restricted perspective that Bohr’s Principle of Complementarity must be understood, and it is in this restricted sense that one must approach his relationship to anschauung. Where the quantum interpretation is concerned, Bohr believed that it is not just the classical concepts that must be understood as symbolic idealizations, but also the concept of observation itself. Bohr’s thoughts on the interrelation of classical concepts, the concept of observation, and the act of description are summed up briefly in the following passage: “The very nature of the quantum theory thus forces us to regard the spacetime co-ordination and the claim of causality, the union of which characterises the classical theories [including spacetime], as complementary but exclusive features of the description, symbolising the idealisation of observation and definition respectively [emphasis in original].80

Bohr’s understanding of the relationship between observation, classical concepts, definition, and description is difficult to parse, and he is not always consistent in how he represents this relationship. It is clear, however, that, in this equation, it is the act of description that imposes constraints upon the other elements, and that observation and classical concepts are symbolic artifacts of how description functions where quantum behavior is concerned. To understand the classical concepts as idealizations within this context is not terribly difficult, but to understand the

98

Chapter 2

concept of observation itself as an idealization is more challenging. For Bohr, observation must be seen as an idealization because there exists no clear separation between the act of observation and its object, since the object transforms based on the observational situation. Therefore, any conventional notion of observation as the detached perception of an objective, stable state can, in the quantum context, only be understood as a construction from which our linguistic and conceptual system cannot escape. Bohr was convinced of the practical and inescapable fact that any experimental situation must be framed in classical terms if it is to be communicated at all. Only through the fullness and definition (as delineation) of complementarity can we transcend ambiguity and open nature up to the customary anschauung that requires, at every level, the separation of object from subject. Bohr saw his Principle of Complementarity not merely as a tool for the representation of quantum behavior, but also as a commentary on the interpretative challenges posed by quantum theory itself. On the difficulty that arises from the fact that language refers to our ordinary perception, for example, Bohr writes, “In the quantum theory we meet this difficulty at once in the question of the inevitability of the feature of irrationality characterising the quantum postulate I hope, however, that the idea of complementarity is suited to characterise the situation [emphasis added].”81 Elsewhere, he writes, “In the last resort an artificial word like ‘complementarity’ which does not belong to our daily concepts serves only briefly to remind us of the epistemological situation here encountered, which at least in physics is of an entirely novel character.”82 Bohr’s use of the word “characterise” indicates that he is speaking about how the very notion of complementarity serves to portray the quantum “predicament,” as does his reference to the “epistemological situation.” Finally, in a 1933 address that Bohr gave to a group of scientists on the benefit of light therapy, he observed that “the notion of complementarity serve[s] to symbolise the fundamental limitation, met with in atomic physics, of our ingrained idea of phenomena as existing independently of the means by which they are observed.”83 The notion of “complementarity” is a commentary upon, a symbolization of, the problem arising out of the unintuitive relationship between experimental arrangement, evidence, and observation. Complementarity

The Physics of Visuality, Intuition, and Aesthetics

99

is not simply a heuristic solution to the representational problem introduced by indeterminacy—it is the thematization of a novel and “irrational” situation which, in Bohr’s opinion, is worthy of consideration in its own right. The applicability of customary intuition to Bohr’s thinking is often more implicit than explicit, and follows a progression of conceptual abstraction from evidence to symbolic representation, a progression that reflects the evolution of his thinking over time as he responded to experimental outcomes—both real and hypothetical. The place of intuition in Bohr’s schematic approach to the quantum phenomena can only be fully understood if one considers the fact that Bohr’s intention was never to develop a coherent and persuasive theory of atomic matter, but rather to communicate the epistemological, heuristic, ontological, and linguistic implications introduced by the quantum interpretation. Driving these considerations was his desire to facilitate the communication of apparently contradictory and unintuitive quantum effects, which in turn led him to reflect upon the way ordinary experience, human perception, and the material world are implicated in one another, and to consider the limits of language when the goal is arriving at a full and accessible description of atomic behavior. HEISENBERG: REDEFINING INTUITION

To change the very concept of a category is to change not only our concept of the mind, but also our understanding of the world. —George Lakoff

With Bohr, we see the loss of the immediate relationship to the material world that characterizes Schrödinger’s position relative to anschauung, with key components such as the classical concepts and the notion of observation taking on the status of symbolic idealizations, and with an emphasis on epistemological concerns. If Bohr’s relationship to anschauung was implicit, Heisenberg wrote and spoke at some length directly about the relationship between anschauung and the quantum interpretation. Bohr accepted the main tenets of the quantum interpretation, but it was Heisenberg who was most invested in the unintuitive matrix model, and this investment led him to break almost entirely from the customary notion of anschauung.While both Bohr and Schrödinger anticipated some

100

Chapter 2

degree of evolution in our language and in our concept formation, Heisenberg’s formal model hung on this evolution, just as it hung on establishing the primacy of mathematics over more perception-based experimental evidence as the intuitive point of entry to the microcosmic realm. This was not an easy task, however, and like his revisions to matrix mechanics, his attempts to redefine anschauung sometimes appear forced, and he often seems to engage more in rhetorical maneuvers than in the pursuit of a coherent understanding of the concept. As I observed in my introduction to this chapter, Heisenberg’s strategy for resolving the problem of anschauung was of a different order than his attempt to resolve the problem of the anschaulichkeit imperative, and this difference is primarily due to the fact that the concept of anschauung is more foundational to human understanding than the concept of anschaulichkeit. While visualizability is a quality that is attached to the object, intuition concerns what we can understand about an object.Thus the two attributes have a different order of relationship between subject and object. Because the attribute of visualizability has a more concrete attachment to objects and is more closely related to sense perception, it is measured by the degree to which an object is available to the senses. Intuition’s conceptual and metaphorical origin is not in the direct apprehension of an object—in the object’s external location or movement in space and time—but rather in the nature of understanding. Intuition thus has a more abstract relationship to perception; it is more sensibility than sense perception and, occupying the same category as reason, more faculty than attribute. In the case of visualizability it makes sense to change the description of the object, for the attributes of an object are more malleable than a sensibility—particularly a sensibility that is not subject to the application of rational thought in the same manner as an object.To make intuition subject to rational thought, then, is to change its place in the order of human understanding, to transform the sense of the term itself rather than to refigure it with respect to an external object. Because intuition represents a capacity for understanding the world, the incompatibility between matrix mechanics and the customary definition of intuition constituted a more intractable problem for Heisenberg than did the definition of what counted as visualizable. Heisenberg faced the troublesome fact that, unlike visualizable elements (such as the particulate elements

The Physics of Visuality, Intuition, and Aesthetics

101

added to the matrix interpretation) the concept of intuition could not be inserted “mechanically” into matrix mechanics. Heisenberg’s main strategy for addressing the accusation that quantum theory was unintuitive, and thus inferior to a classically derived theory, was to reposition, rhetorically, the relationship between theory and experiment, so that a theory became correct by virtue of its internal coherence rather than by virtue of its consistency with experimental evidence. If a coherent theory was all that was needed, then entirely hypothetical experiments could be imagined that would confirm the theory, there would be no disjuncture between experimental evidence and the quantum interpretation, and the quantum interpretation would “feel” more intuitive. Of Heisenberg’s strategy to invert the canonical relationship between experiment and theory, as well as his claim that a theory need only be free of contradiction to be intuitive, Paul Forman observes: “This redefinition equated ‘intuitive’ to ‘satisfactory’ in a strictly positivist sense … [p]erhaps the best index of the wantonness of this solution is that in their colleagues’ eyes Heisenberg and Born (who joined in this redefinition) had simply inverted the ordinary, accepted signification of the word” (anschaulich).84 In this inverted model, the representation of quantum behavior no longer follows from experimental evidence; rather, the “intuitive” mathematical abstractions of matrix mechanics precede and predict experimental outcomes, and these hypothetical outcomes are intuitive because they conform to predictions built into the mathematical model. Heisenberg’s colleague and teacher Arnold Sommerfeld sums up Heisenberg’s strategy in the following manner: “From an intuitive theory in this sense one ought thus demand only that it be without contradiction with itself, and that it permit prediction, without ambiguity, of the results of all imaginable experiments in its domain.”85 Since observed evidence no longer structures the theory, the extent to which theoretical models are both persuasive and intuitive is no longer evaluated with reference to their physical content. This revision promised a significant payoff for Heisenberg, because it meant that supporters of Schrödinger’s more physical wave mechanics no longer had claim to superiority over the abstract algebra of matrix mechanics, and any existing experiment that did not confirm a theory could be dismissed as incorrect.

102

Chapter 2

Heisenberg’s attempt to resolve the tension between theory and experiment is demonstrated variously in his 1927 paper “The Actual Content of Quantum Theoretical Kinematics and Mechanics,” his 1962–1963 interviews with Thomas Kuhn, and his 1958 collection of lectures, Physics and Philosophy: The Revolution of Modern Science.86 “The Actual Content”—the paper that famously introduced matrix mechanics—represents Heisenberg’s first concerted attempt to redefine anschauung. In fact, Heisenberg begins his paper with an attempt at this redefinition, suggesting that even at this early stage he was preoccupied with the accusations that his theory was unintuitive: We believe to understand a theory intuitively, if in all simple cases we can qualitatively imagine the theory’s experimental consequences and if we have simultaneously realized that the application of the theory excludes internal contradictions. For instance: we believe to understand Einstein’s concept of a finite three-dimensional space intuitively, because we can imagine the experimental consequences of this concept without contradictions.87

In this passage, Heisenberg links both theoretical coherence and hypothetical experiment; imagined experimental outcomes cannot contradict theory, because the internal coherence of the theory guarantees that any imagined outcome will be internally coherent as well. Heisenberg risks disaffirming his claim on the following page, however, when he asserts that quantum theory is more legitimate than classical theory because it is derived from “relations between concrete, experimentally derived values,” meaning that the origins of matrix mathematics in experimental evidence demonstrate that the mathematics need not be revised. Heisenberg’s apparent retreat here from his inversion of theory and experiment suggests that he remained cautious about the this strategy, particularly since he has just admitted that “quantum mechanics is still full of internal contradictions [innerer widersprüche].88 Heisenberg quickly regains his footing, however, when he claims, “the fact that a revision of the [classical] kinematic and mechanic concepts is required seems to follow immediately from the basic equations of quantum mechanics.” Here again, the fact that the necessary revision to classical concepts “seems to follow immediately” from mathematical values suggests that if there is conflict between mathematics and the experimentally derived classical concepts, it is the classical concepts that need to be modified.

The Physics of Visuality, Intuition, and Aesthetics

103

About halfway through the paper, Heisenberg expands on his inversion of the relationship between theory and experiment, when he speculates that “[a]ccording to the intuitive [anschaulich] interpretation of quantum theory attempted here, the points in time at which transitions— the ‘quantum jumps’—occur should be experimentally determinable in a concrete manner.”89 Heisenberg then sums up his argument with the claim that “since … we can qualitatively conceive of the theory’s experimental consequences, in all simple cases, we shall no longer have to view quantum mechanics as not intuitive [anschaulich] or abstract.”90 If it appears here that “experiment” is the final arbiter of the correctness of the mathematics, it is still the theory that drives what experiments can be imagined—experiments that need not ever be performed. The primacy that Heisenberg assigns theory is not merely implicit in his arguments, however; he also states his opinion quite explicitly during an interview with Thomas Kuhn in 1963: “one felt that by making the probability become some kind of reality, you get hold of something which is there. … [T]he main point was that the probability itself was something real. It was not only in the mind of the people, but it was something in nature. … Nobody wanted to lose this concept even if the result of the experiment of Bohr and Bothe-Geiger came out contrary to this idea.”91 Heisenberg restates this opinion in a second passage, in his 1974 recollection of his first conversation he had with Einstein: “[o]nly when you have the complete theory can you say what can be observed. … So we did convince ourselves, that we now had a scheme by which we could predict every experiment.We did not doubt that nature would also follow these predictions.”92 In the first passage, Heisenberg implicitly admits to the rhetorical strategy of first asserting the reality of quantum concepts (“making the probability become some kind of reality”), and then acting as though its status as “reality” is a priori (“It was not only in the mind of the people, but it was something in nature”). Heisenberg goes on to make his final point— that the theory is “the requirement for experimental verification.” In the second passage, the inversion of the relationship between quantum theory and experiments becomes complete: not only will the experiments follow the theory and be subject to its laws, but mathematical schemes are legitimized by nature itself, and can only arise in the imagination if they are already inherently true.

104

Chapter 2

Any lingering doubt surrounding the relationship of theory to experiment in Heisenberg’s model vanishes when he recalls, in 1974, the turning point in his perspective after a conversation he had with Einstein in 1922: “he [Einstein] explained that what can be observed is really determined by the theory. He said, you cannot first know what can be observed, but you must first know a theory, or produce a theory, and then you can define what can be observed.”93 Heisenberg then recalls “We had asked, given the situations in nature like the orbit in a cloud chamber, how can it be described with a mathematical scheme?”94 A problem arises, however, from the fact that the concept of an orbit presupposes knowledge of both position and velocity, something that is not possible according to the quantum interpretation. Heisenberg then describes the epiphany that provided the solution: “could it not be the other way around? Namely, could it not be true that nature only allows for such situations which can be described with a mathematical scheme? Up to that moment, we had asked the opposite question. … But by turning it around, one could at once see that now it’s possible, if I say nature only allows such situations as can be described with a mathematical scheme.”95 After having come to this realization, Heisenberg says, he proceeded to worked out the calculations through which he was able to revise the representation of the atomic behavior that was assumed to produce the observed evidence, and thus lay the foundations for the Uncertainty Principle, which Heisenberg felt “established the much-needed bridge between … observations and the mathematics of quantum mechanics.”96 Reflecting once more on his thoughts after the conversation with Einstein, Heisenberg says, True, it had still to be proved that any experiment whatsoever was bound to set up situations satisfying the uncertainty principle, but this struck me as a plausible a priori, since the processes involved in the experiment or the observation had necessarily to satisfy the laws of quantum mechanics. On this presupposition, experiments are unlikely to produce situations that do not accord with quantum mechanics.97

Here, Heisenberg goes far beyond anticipating that the theory will be verified in the future via experimental evidence; instead, he makes the more radical claim that experiments must and will conform to the

The Physics of Visuality, Intuition, and Aesthetics

105

Uncertainty Principle, which constituted the cornerstone of the quantum interpretation. Heisenberg’s formalism, along with dismissal of empirical evidence as the ground for an intuitive theory, appears to have led to the very sort of disconnection and loss of intelligibility that Schrödinger predicted. In fact, Heisenberg criticized what he felt was the overextension of his own abstract mathematical formalism when, in his 1930 publication, The Physical Principles of the Quantum Theory—only three years after the fundamental principle of quantum theory had been established—he lamented that “even today a physicist more often has a kind of faith in the correctness of the new principles than a clear understanding of them.”98 The final culmination of the association of Heisenberg’s notion of intuition with mathematical formalism can be found in Henk de Regt’s “Spacetime Visualisation and the Intelligibility of Physical Theories.” In this article de Regt arrives, through a series of logical abstractions, at the conclusion that an intuitive grasp of quantum physics can be achieved merely through repeated exposure to its mathematics, and that this intuitive understanding requires no direct or indirect connection to experiment, or to the physical world at all. Immediately following his identification of Pauli, Heisenberg, and Dirac’s “well developed” “capacity for abstract reasoning,” de Regt argues that “[o]ne can get accustomed to the mathematical concepts used in a theory or to an initially weird physical interpretation of a theory, to such an extent that it becomes possible to reason in the associated terms in an intuitive manner.”99 De Regt thus claims that one can intuit a physical situation without even possessing the skill to solve for the equations that are already an abstract expression of this situation. Not only does de Regt sever intuition from any relationship to the physical world, what he calls his “familiarity-view” of understanding extends Heisenberg’s mathematical formalism to the point where mere repeated exposure to “mathematical concepts” satisfies the conditions for the deep understanding implied in the term “intuition.” To be sure, one can agree with de Regt’s argument that a student of quantum physics would be capable of memorizing an equation, and that this student would be capable of passing on the variables of the equation without ever understanding them. Not surprisingly, however, defining intuitive knowledge as the acclimatization to a concept, and replacing physically based

106

Chapter 2

understanding with mathematically based explanation assumes, somewhat confoundingly, that ignorance of the physical behavior of matter is acceptable in the field of physics. De Regt’s reference to an “initially weird physical interpretation of a theory” represents a further distortion of the manner in which Heisenberg privileged theory over experiment, for to accept his logic here is see the physical world as a simulacrum—an abstract representation. Although de Regt’s conclusions represent the logical extension of Heisenberg’s mathematical formalism, Heisenberg would surely have rejected the terms of de Regt’s argument. In the first place, while Heisenberg may have expressed rhetorically his anticipation of a time when the measure of intuition would be taken according to the conceptual coherence of theory and mathematical constructs—rather than according to everyday experience—his postulates were conceptually grounded in the outcomes of experimental arrangements. Heisenberg’s Uncertainty Principle, for example, was derived from predictions about the physical interaction between matter and measuring instrument, and Heisenberg accepted that, where the unavoidable description of quantum events in language was concerned, we must accept that the communication must take place in the space and time of classical, sense-based concepts. When de Regt constructs his “non-visual” approach to physics, he abandons Heisenberg’s acceptance that classical language was at times necessary. In the end, though, while it seems certain that Heisenberg would have found de Regt’s argument a little too formal, it may be viewed as the final outcome of Heisenberg’s effort to detach intuition from everyday experience, which culminates in a situation where the physical world is fading into irrelevance. Heisenberg’s effort to legitimize quantum theory by redefining intuition is also illustrated in his claims about the historicity of intuition. Following the discontinuous logic of matrix mechanics, Heisenberg and Pauli departed altogether from the historically minded commitment to the continuity of knowledge production that Schrödinger saw as an obligation of scientific discovery. Instead, Heisenberg and Pauli argued that scientific concepts would inevitably evolve, and with them, the faculty of intuition. According to Heisenberg, any contradiction between conventional concept formation and quantum theory was the product of a contingent and historically situated (rather than essential) restriction. As

The Physics of Visuality, Intuition, and Aesthetics

107

Heisenberg saw it, the use of classical concepts such as space and time, while currently a requirement of language (if not of mathematics), reflect a present limitation in our mode of thinking, rather than a set of a priori conditions. In Physics and Philosophy, Heisenberg similarly observes,“since they [the classical concepts] represent for the time being the final result of the development of human thought in the past, even in the very remote past; they may even be inherited and are in any case indispensable tools for doing scientific work in our time. In this sense they can be practically a priori. But further limitations of their applicability may be found in the future.”100 In both passages, Heisenberg uses the word “practically,” not in the sense of “almost” or “nearly,” but rather in the sense of “pragmatically,” allowing that while these concepts may be considered a priori, this is because they are currently the only ones that we have.101 The phrases “for the time being,” “in our time,” and “may be found in the future” similarly refer to his belief that intuitive classical concepts such as space, time, and causality are, in Kristian Camilleri’s words, “historically contingent, yet indispensable in our time, because we have no other language through which we can describe and conceive of the interaction between ‘object’ and ‘measuring device.’”102 Heisenberg rejects Schrödinger’s argument that an unbroken continuity from the past to the future is the prerequisite for conceptual intelligibility. Instead, Heisenberg suggests that continuity with the past leads to conceptual anachronisms, and he anticipates a future point where classical concepts will be less applicable and a new form of intuition will emerge that accords with the quantum interpretation. In his February 1963 interview with Kuhn, Heisenberg makes a similar point when he reflects back on the early days of quantum theory: Well, I quite see that you look at this problem from the historical point of view, and I certainly would agree that I do not know how people will talk about these problems say 1000 years from now. It may be that there the language and the concepts have changed so much that they would not use the Newtonian concepts at all anymore. They would just have different words which fulfill all the functions of the words in Newtonian physics which we now have.103

In the same interview, Heisenberg predicts the same evolution of the parameters of understanding:

108

Chapter 2

We had seen that we could, possibly, understand nature, but at the same time we had to learn that the word “understand” means something different from what we had believed it would mean in the earlier times. So it is a whole change of the attitude about what we can do and cannot do. “Understanding” means suddenly just predicting experiments. “Understanding” does not mean something like in the old classical physics. That change took place already in the very first stage of quantum mechanics, I would say in [19]24 already.104

Heisenberg’s vision of a future break with existing language and concepts, illustrated in the preceding passages, is consistent with the discontinuity of Heisenberg’s matrix mechanics, in the same manner that Schrödinger’s emphasis on historical and conceptual continuity is consistent with his wave mechanics. Like the concept of anschaulichkeit, the concept of anschauung had and has special significance in the context of quantum theory—both for the creators of the new physics and for those who have sought to interpret that creation. While I have shown how the manner in which they addressed the concepts of anschaulichkeit and anschauung was directly related to the theoretical investments of Schrödinger, Bohr, and Heisenberg, I have rejected the tendency to associate anschauung too closely with anschaulichkeit, arguing that the two concepts are of a different order, were approached differently by the key players, and ought to be analyzed separately. A comparative analysis—of the different notions of anschaulichkeit and anschauung, and of the distinctive manner in which Schrödinger, Bohr, and Heisenberg approached each concept—reveals not only their respective theories of atomic matter, but also the structural, epistemological, and philosophical principles that inform them, and the rhetorical strategies that they used to defend them. JUDGMENTS OF TASTE AND THE AESTHETICS OF PHYSICS

The final harmony we visually admire is but the product of the rules of harmony that have governed its makings. —Paul Weiss

Where it concerns their aesthetic sensibilities, there is a good deal of accord between Schrödinger, Bohr, and Heisenberg, expressed in an affinity that rests upon two notions: harmony and simplicity. All three men

The Physics of Visuality, Intuition, and Aesthetics

109

privilege harmony in particular—even and especially Heisenberg, who up until this point I have largely associated with disjuncture and discontinuity. Both Bohr and Heisenberg find harmony through “unity in multiplicity”—Heisenberg, in the way that a few fundamental mathematical axioms unite the manifold mathematical expressions, and the way the quantum interpretation imposes order on a set of confusing, because seemingly irreconcilable, quantum phenomena. Bohr also stresses unification, and while Bohr’s relationship to this unification is more undefined and inferential, it is clearly a driving force for how he approaches the natural world, starting with the symbolic unity of classical concepts that defines his Principle of Complementarity, and extending to his belief in the shared interests between fields of knowledge other than his own. For Heisenberg, simplicity lies in the purity of mathematical forms, for Bohr, in the fundamentals of description that, regardless of the theory or context, must take on classical form, and for Schrödinger, in the spare expression of the object’s core characteristics. In some respects, despite the novelty of their theories, all three men are classical in their sensibilities: each one seeks out the one in the many, the unity in difference, and the essential in, as Heisenberg puts it, “the plethora of details.” SCHRÖDINGER: SCIENCE UNADORNED

If the criterion of usefulness be thoroughly carried out it will evolve its own kind of beauty. —Erwin Schrödinger

In his 1957 book Science, Theory, and Man, Schrödinger writes, “I need not here speak of the quality of the pleasures derived from pure knowledge; those who have experienced it will know that it contains a strong aesthetic element and is closely related to that derived from the contemplation of a work of art.”105 Schrödinger’s relationship to art was more than contemplative, however, for he saw a very specific connection between the art and the science of his day, and the formal qualities that he appreciated in the arts reflected those formal qualities he felt ought to define a theory of the material world. In a section of the book titled “Simplicity and Purposefulness in the Arts and Crafts,” Schrödinger writes:

110

Chapter 2

In art, for instance, a dominant idea is that of simplicity and purposefulness —reine Sachlichkeit, to use a German expression—and in all our crafts the same thought rules. … [O]ur artists will be satisfied to catch the main features of their sitter and they will consider all attempts at decoration or careful painting of accessories as a hindrance to the main purpose, which is to present the character of the sitter as expressed in its main features. Behind all our craftsmanship there is the very same will to purposefulness. … Everything is banished which does not contribute to the main purpose in view. We feel that we do not want any ornamentation that would not be in harmony with the keynote of practical usefulness. … [W]e are convinced that if the criterion of usefulness be thoroughly carried out it will evolve its own type of beauty.106

Schrödinger’s emphasis on spareness and functionality in art seems straightforward enough and, in his emphasis on “practical usefulness,” one can perhaps already see hints of an approaching analogy to the scientific approach. When he does turn to science, however, the analogy takes a strange turn, one that reveals something more than attentiveness to “purpose” and “usefulness.” Schrödinger continues on to observe: Now, there is something similar in our science. We are beginning to make a point of constructing our picture of the physical universe in such a way as to represent only the facts that can be actually verified through experiment and we eschew as far as possible all voluntary theories or assumptions.We want no ornamental accessories. Just as we are no longer afraid of bare surfaces in our furniture and our dwelling rooms, so in our scientific picture of the external world we do not try to fill up the empty spaces.We try to exclude everything that in principle cannot be the object of experimental observation.107

Twice, Schrödinger makes a reference here to the importance of including only that which can be verified by experiment. His emphasis on experimental evidence here can easily be interpreted as a repudiation of matrix mechanics, where mathematics drives what most often remain hypothetical experiments. Schrodinger’s criticism of ornamentation can also be interpreted as a critique of matrix mechanics. One might at first consider mathematics as the form most devoid of ornamentation and most austere—the form that dispenses with all that is extraneous and external to it. Mathematics appears to admit no cumbersome experimental apparatus, no possibility of observing what is not there, and to require for itself only internal harmony. Schrödinger’s attack on mere theoretical

The Physics of Visuality, Intuition, and Aesthetics

111

constructions is even more explicit when he continues on to observe,“[a] nd we think it is better to leave our feeling of incompleteness unsatisfied rather than to introduce mental constructions which cannot by their nature be experimentally controlled or tested for their correspondence to external reality.” The algebra is this “mental construction,” and the abstract matrix mechanics cannot, by its nature, be tested for its “correspondence to external reality.” In fact, Schrödinger is critiquing the very assumption that Heisenberg attempted to overturn with his inversion of theory and experiment—to make the experimental evidence follow the theory, so that the theory becomes the “external reality” against which experiment is tested. On physical systems, Schrödinger writes, “we are requested to refrain from giving any meaning at all to the conception of the actual energy of the system! Our world picture has to remain bare and empty in this respect—we are not afraid of the empty space on the canvas [emphasis in original].”108 One might be tempted, encountering this phrase out of context, to reply that the notion of quantum jumps fulfills just this requirement, and that accepting the inability to know what goes on between two jumps is the very definition of “not being afraid of the empty space on the canvas.” But Schrödinger means just the opposite— rather than accept the illogic of quantum jumps, it would be better to admit that we cannot form a picture of the actual energy of a system at all, including the problem of transitions between states that the notion of quantum jumps was designed to resolve.109 Quantum jumps are another example of mere “mental constructions,” and certainly, from Schrödinger’s perspective, they appeared to be just one part of a piecemeal effort to impose coherence onto the quantum interpretation. For Schrödinger, then, these quantum jumps are merely “ornamental” additions to an already fragmentary formulation of atomic behavior, and thus are not “in harmony with the keynote of practical usefulness.” Schrödinger’s critique of fashionable aspects of quantum physics— something that I introduced earlier when I discussed his thoughts on the “newfangled” theory of probability—also emerges in his thoughts on art and aesthetics, which in turn grow out of his investment in the value of historical continuity in science. In “Are There Quantum Jumps: Part I,” Schrödinger leaves no doubt about his rejection of the current

112

Chapter 2

voguishness in both science and culture, which he believes acts to the detriment of both scientific practice and artistic form: The disregard for historical connectedness, nay, the pride of embarking on new ways of thought, of production and of action, the key endeavor of shaking off, as it were, the indebtedness to our predecessors, are no doubt a general trend of our time. In the fine arts we noticed strong currents quite obviously informed by this vein; we witness its results in modern painting, sculpture, architecture, music and poetry. … But I may say that whenever this trend enters science, it ought to be opposed.110

As he does on the topic of anschaulichkeit, Schrödinger opposes an irresponsible embrace of the new for its failure to recognize the essential value of retaining a sense of congruity between the past and the present. Schrödinger appears, concerning the arts at least, to allow for a break with the past, though he clearly disapproves of this “shaking off ” of “historical connectedness,” and he proceeds to criticize the poetry of the Spaniard Louis de Gongóra (1561–1627) for its neologisms and for how these neologisms contribute to an excess of words that either conceal meaning or confound it altogether. Schrödinger admits that Gongóra poems are “well-sounding,” but objects to the fact that he applies “all his acuity and skill” toward “making it as difficult as possible to the reader to unravel the sense, so that even natives of Castile use extended commentaries to grasp the meaning safely.”111 In the next paragraph, Schrödinger insists that physics must oppose “this general trend in our time,” and that “[t]hough we are entirely the product of historical development, yet it is we who make its continuation and not history that drags us along a predestined trail.”112 With wave mechanics Schrödinger emphasizes connectivity—in his rejection of discrete “particles” in favor of continuity-based “wave-crests,” in his rejection of the notion of quantum jumps, and more generally in his commitment to the classical concepts and their concrete ties the material world. Similarly, connection is a key value expressed in Schrödinger’s understanding of the relationship between nature, history, culture, and art. For Schrödinger, any theory of the material world must be understood as part of a larger cultural and societal milieu. Schrödinger rejects art that includes irrelevancies, and values connection to those members of society whose interests extend beyond the rarified world of quantum mathematics, since results in physics are “destined, eventually, to

The Physics of Visuality, Intuition, and Aesthetics

113

be absorbed into our world picture and to become part and parcel thereof.” According to Schrödinger, physics and art have implications beyond their time, and in relation to the larger questions that define their historical moment. BOHR’S UNIFYING SENSIBILITY

That an object could be several things, could change, could be seen as a face, a limb and a fruitbowl. —Mogens Anderson

In “Physics as Art: The German Tradition and the Symbolic Turn in Philosophy, History of Art and Natural Science in the 1920s,” Catherine Chevalley traces a “symbolic turn” in art, science, mathematics, and philosophy—one that she associates with a movement beginning with Immanuel Kant and progressing through Wilhelm von Humboldt, Hermann von Helmholtz, and Heinrich Hertz, toward an aesthetic judgment wherein human understanding becomes indirect, symbolical, and linguistic. While most of Chevalley’s article is devoted to summarizing how Ernst Cassirer and Erwin Panofsky trace the expansion of the concept of symbol across fields of knowledge and cultural production, toward the end of her article she offers a few thoughts on how Bohr and Heisenberg might be seen as emerging from this tradition, two of which I will examine. “Only if we … allow that language plays a constitutive role in objectivation,” writes Chevalley, “can we arrive at the statement that physics is like art.”113 One of the primary ways in which Chevalley connects Bohr to the German tradition is in relation to his approach to language, and indeed, Chevalley places Bohr squarely within the movement toward a “world of representation” where language as a symbolic system is accorded primary of place. According to Chevalley, Bohr believed that quantum physics raised the following “epistemological paradox”:“On the one hand, our interpretation of the experimental material rested essentially upon the classical concepts, which were refined concepts of ordinary language; on the other hand, objects which were not given in ordinary space and time could not be described with the help of classical concepts and ordinary language.”114 Chevalley’s argument that Bohr’s epistemology grew out of what he saw as a paradox seems fair: one of his primary goals was to address the

114

Chapter 2

fact that, in any quantum description, it is necessary to use the classical concepts derived from everyday experience, while at the same time facing the fact that experimental results at the quantum level do not allow for the direct use of the classical concepts in description. Chevalley goes on to argue that “[t]his paradox, however, persisted only as long as one assigned ontological weight to concepts and words, and disappeared under the auspices of a more sophisticated conception of language.”115 Again, Chevalley is right to point out that Bohr’s resolution was to remove the “ontological weight” of language, and to insist that quantum description could only be the symbolic idealization of classical description. It must be emphasized, however, that Bohr’s attention was on the linguistic problems raised by the quantum situation, and he did not overly concern himself with a general theory of language as nonreferential or detached from perception-based experience; in this sense, his sensibility must be understood as tied directly to the quantum interpretation. Chevalley goes on to write that “in conjunction with a strategy of the ‘multiplication of the points of view,’ Bohr developed an understanding of language as a network of conceptual systems which provide different descriptions of the object.”116 Chevalley’s observation aligns with Bohr’s attraction to cubism, and with his goal of generating a mode of description that could represent more than one perspective. However, Chevalley’s argument that language presented itself, for Bohr, as “a network of conceptual systems” that provide different descriptions of the object is somewhat less accurate. For Bohr, description was tied to a single conceptual system—that of the classical concepts derived from classical observation of classical outcomes produced by classical experimental equipment. Different arrangements of the equipment will produce a different outcome, but the description of quantum behavior does not and cannot, for Bohr, break free from the single system of classical concepts. Furthermore, Bohr departs from both cubism and from Chevalley’s “network” metaphor in his consistent tendency toward binarism, which is expressed in the large number of paired relationships to which Bohr ascribed the term “complementarity” and which he felt expressed reciprocal relations: contrasting experimental outcomes, competing theoretical models (including relativity and quantum mechanics), the perceiving subject and the perceived object, different but related experimental arrangements, and even the mind–body relation.

The Physics of Visuality, Intuition, and Aesthetics

115

Unlike Chevalley, I do not view the movement toward the symbolic as Bohr’s primary aesthetic, although it is true that Bohr sees language—as well as the very notion of observation—as symbolic within the quantum context, just as it is true that his unique sensibility spans the fields of science, epistemology, and philosophy. This sensibility, however, is located primarily in his preference for harmony and the reconciliation of points of view. Bohr’s Principle of Complementarity is precisely about the effort to harmonize two apparently contradictory results, and it is within this context that Bohr’s notion of symbolic description must be understood. Because Bohr wishes to present a unified quantum interpretation without sacrificing the discreteness of quantum phenomena, and because the classical concepts and ordinary language are “all that we have,” Bohr sees no other option but to see these concepts as symbolic forms. For Bohr, the impulse to harmonize comes first, and the symbolic idealization of the classical concepts comes second. Bohr’s harmonizing impulse was not limited to reconciling apparently contradictory results; it also included his desire to harmonize different theories of atomic matter. In “The Quantum Postulate,” for example, he writes,“I should like to make the following general remarks regarding the principles underlying the description of atomic phenomena, which I hope may help to harmonise the different views, apparently so divergent, concerning this subject.”117 In part what Bohr has in mind here is the reconciliation of wave mechanics and matrix mechanics by showing, for example, that the wave theory of superposition applies in one instance, while the particulate quantum of action applies in another. Bohr’s desire to unite theories of the material world that reach beyond the quantum realm also included an attempt to reconcile relativity and quantum theory—two theories that remain infamously incompatible to this day. In his paper responding to the Einstein-Podolsky-Rosen article “Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?” Bohr sets out to demonstrate how his complementarity viewpoint leads to a complete description of quantum behavior. At the end of his paper, Bohr writes: I should still like to emphasize the bearing of the great lesson derived from general relativity theory upon the question of physical reality in the field of quantum theory. In fact, notwithstanding all characteristic differences, the situations we are concerned with in these generalizations of classical theory

116

Chapter 2

present striking analogies which have often been noted. Especially, the singular position of measuring instruments in the account of quantum phenomena, just discussed, appears closely analogous to the well-known necessity in relativity theory of upholding an ordinary description of all measuring processes, including a sharp distinction between space and time coordinates, although the very essence of this theory is the establishment of new physical laws, in the comprehension of which we must renounce the customary separation of space and time ideas.118

Bohr’s conciliatory tone here stands in sharp contrast to the original Einstein-Podolsky-Rosen article’s attempt to discredit quantum theory. However, in order to reconcile the two theories, Bohr has to generalize quite broadly—in this instance, through an analogy built upon the fact that both theories must resort to “ordinary description” in the end. This passage also includes a conciliatory nod to the fact that it was relativity that first encountered the problem of classical spacetime description—a problem central to the discourse around the quantum interpretation— and that the quantum interpretation has in this respect learned a “great lesson” from relativity. Like his Principle of Complementarity, which harmonizes incompatible experimental outcomes, Bohr’s expansive effort here to achieve a rapprochement between relativity and quantum theory demonstrates that he is significantly invested in harmonizing our conceptualization of the material world. Bohr’s impulse to integrate seemingly distinct perspectives extended beyond apparently mutually exclusive theories of matter and into the relationship between different fields of knowledge. In a 1932 address delivered at the opening meeting of the International Congress on Light Therapy, Bohr observed that “atomic physics … forces us to an attitude towards the problem of explanation recalling ancient wisdom, that when searching for harmony in life one must never forget that in the drama of existence we are ourselves both actors and spectators.119 Here, Bohr embeds within a much broader disciplinary context an allusion to the Uncertainty Principle, which describes the observer’s inevitable influence on the observed; at the very same time that we adopt the position of spectators, we also enact a force upon the object. As demonstrated in the following passage, Bohr hoped to establish the fact that our status as both actors and observers transcends any single field of knowledge and extends into other aspects of our existence, including both the study of psychology and our own psychic experience. Bower states that

The Physics of Visuality, Intuition, and Aesthetics

117

quantum theory presents us with a novel situation in physical science, but attention was called to the very close analogy with the situation as regards analysis and synthesis of experience, which we meet in many other fields of human knowledge and interest. As is well known, many of the difficulties in psychology originate in the different placing of the separation lines between object and subject in the analysis of various aspects of psychical experience.120

Bohr makes a similar analogy to the field of biology in his introductory comments when he reflects upon “what significance the results reached in the limited domain of physics may have for our views on the position of living organisms in the realm of natural science,” and what these results might tell us about “the subtle character of the riddles of life.”121 Once again, Bohr seeks points of connection, and in this case his thoughts delve into questions of epistemology and philosophy, attempting to forge unity out of the apparent discontinuities across experience, and correspondence between subjective experience and the material world. In the end, Bohr’s aesthetic is not one of “detachment from,” but rather one of “connection to,” and any abstractions that he incorporates are designed to serve this impulse toward connectivity. Only by idealizing the unavoidable classical concepts can he make them “fit” with quantum experimental evidence, and only by insisting that description be understood as symbolic, where quantum behavior is concerned, can he use the language of the everyday from which he believes there is no escape. Bohr may have been fascinated by the multiple points of view offered in cubism—the fact that “an object could be several things, could change”— but his ultimate investment was not in multiplying points of view, but rather in reconciling them and presenting a picture wherein the different facets form a unified whole. HEISENBERG AND THE MATHEMATICAL SUBLIME

Beauty is truth, truth beauty. —John Keats

Like Bohr, Heisenberg privileges the value of harmony, and like Schrödinger, he values simplicity. Unlike Bohr, Heisenberg refers much more specifically to the aesthetics of beauty, and unlike Schrödinger, Heisenberg’s aim is inevitably to elevate mathematics over empirical science. In a 1971 letter to Einstein, Heisenberg comments on the connection between beauty and mathematics: “You may object that by speaking

118

Chapter 2

of simplicity and beauty I am introducing aesthetic criteria of truth, and I frankly admit that I am strongly attracted by the simplicity and beauty of mathematical schemes which nature presents us. You must have felt this too: the almost frightening simplicity and wholeness of the relationship, which nature suddenly spreads out before us.”122 In a move that recalls Bohr, Heisenberg attempts to foster a connection between relativity and the quantum interpretation by referencing a shared understanding. In Heisenberg’s case, however, the goal is not to forge a collective connection, but rather to set himself and Einstein, as mathematicians, apart from others. In Across the Frontiers, Heisenberg also expresses this tendency toward exclusivity when he observes that although the attempt in quantum theory to order “the bewildering plethora of details” remains largely unintuitive, its completeness and “abstract beauty” will be immediately convincing “to all who could understand and speak such an abstract language.”123 Heisenberg states that he derives his aesthetics of mathematics from the Pythagorean notion of beauty, in which beauty is “an ideal principle of form,” wherein all things in their nature are revealed through numbers.124 According to Heisenberg, Pythagoras’s discovery that the numerical ratio is the source of harmony in music is “one of the most momentous discoveries in the history of mankind,” since it showed that the mathematical relation was also the source of beauty.125 For Heisenberg, the primacy and beauty of numbers is also the primacy of theory over experience, expressed in the conflict between the empiricist, which by careful and scrupulous detailed investigation first furnishes the presuppositions for an understanding of nature, and the theoretician, who creates mathematical pictures whereby he seeks to order and so to understand nature—mathematical pictures that prove themselves, not only by their correct depiction of experience, but also and more especially by their simplicity and beauty, to be the true Ideas underlying the course of nature.

It is no small thing for Heisenberg to tie mathematics to the “true Ideas” underlying nature, for it is here where he brings in Plato’s notion of the “central Idea” in which,“the divine becomes visible and at sight of which the wings of the soul begin to grow.”126 Heisenberg values harmony and beauty particularly insofar as they are expressed through unity. He incorporates an aesthetics of beauty into the

The Physics of Visuality, Intuition, and Aesthetics

119

process of unification in two ways: first, he argues that, while unity and simplicity are not precisely the same (nor, for that matter, are they necessarily tied to beauty), “the fact that in such a theory [Newton’s laws] the many are confronted with the one, that in it the many are unified, itself has the undoubted consequence that we also feel it at the same time to be simple and beautiful.”127 Here, Heisenberg equates unity, simplicity, and beauty, an association that he extends elsewhere when he juxtaposes the phrase “The simple is the seal of the true” (Simplex sigillum veri), and “Beauty is the splendor of truth” (Pulchritudo splendor veritatis).128 The combined sense of these statements may be summed up in the following way: unity equals simplicity, and simplicity equals both beauty and truth—and unity, simplicity, beauty, and truth combine in the mathematical form. While Heisenberg acknowledges here that Newton was instrumental in reinventing a science wherein the many are unified with the one, the unification that Heisenberg has in mind is not finally Newton’s unification of a multifarious material world under a set of formal laws, but rather the unification of individual mathematical expressions under a set of mathematical axioms. The second way in which Heisenberg ties the axioms of mathematics to aesthetics occurs when he observes that the reduction of the many mathematical expressions to a few simple axioms is similar to how the multifarious architectural styles can be reduced to certain “basic forms,” as is expressed, for example, in the “semicircle and rectangle in Romanesque architecture.”129 All variations of architecture emerge from these initial mathematical forms and according to Heisenberg, the variations are already “perceivable in these original forms, even at the outset; otherwise it would be scarcely comprehensible that many gifted artists should have so quickly resolved to pursue these new possibilities.”130 At the same time that Heisenberg returns, with architecture, to a material expression of unity and harmony, he also makes his most a priori argument about mathematical forms in arguing that every material expression is already contained in the archetypal system of numbers. Like Schrödinger, Heisenberg envisions an informing “spirit” behind artistic production; however, while Schrödinger focuses on the spirit of his age—the fact that “the very same will to purposefulness” is behind all our craftsmanship— Heisenberg’s informing spirit is mathematical form.

120

Chapter 2

Heisenberg also creates a hierarchy between the sensate world and the world of form and idea with which mathematics is associated when he observes that “[m]aterial things are the copies, the shadow images, of ideal shapes in the mind; moreover, as we should be tempted to continue nowadays, these ideal shapes are actual because and insofar as they become ‘act’-ive in material events.”131 Again, as he does with intuition, Heisenberg elevates ideas, as the vehicle through which nature is revealed, over empirical, sense-based evidence. Significant to Heisenberg, then, is the fact that Plato distinguishes “with complete clarity a corporeal being accessible to the senses and a purely ideal being apprehensible not by the senses but only through acts of the mind.”132 Heisenberg consistently associates embodied, sense-based experience with a lower order of apprehension, in one instance citing how Plato contrasts the “imperfect shapes of the corporeal with the perfect forms of mathematics.”133 Heisenberg later reinforces this hierarchical distinction in his inclusion of an extended quotation from Kepler, who similarly associates the faculty that perceives what is given to the senses with “the lower region of the soul,” and who argues that “just as the sensorily presented things in the outer world recall to us those which we formerly perceived in the dream, so also the mathematical relations given in sensibility call forth those intelligible archetypes which were already given inwardly beforehand, so that they now shine forth truly and vividly in the soul, where before they were only obscurely present there.”134 With specific reference to the fact that quantum physics was immediately found convincing by virtue of its completeness and abstract beauty, Heisenberg asks, “Why is it that with this shining forth of the beautiful in exact science the great connection becomes recognizable, even before it can be understood in detail and before it can be rationally demonstrated?”135 Heisenberg’s thoughts on the primacy of mathematical abstraction here again recall his strategy of reinventing intuition by inverting experiment and theory; in this case, it appears that inner beauty (or internal coherence) is the vehicle through which true connections are recognized. Since the abstraction is established as beautiful because true, true because beautiful, no empirical demonstration is necessary. In his reinvention of the unity and simplicity of Newtonian physics, Heisenberg bypasses the correspondence in classical physics between the formal expression of an idea and the sense-based perception of

The Physics of Visuality, Intuition, and Aesthetics

121

phenomena in the material world, and locates truth entirely within the “complete and abstract beauty” of the mathematical forms that “restored exact science, under entirely new presuppositions, to that state of harmonious completeness which for a quarter of a century it had lost.”136 Because Heisenberg has already elevated aesthetically-based mathematical truth above sense-based empirical evidence, disconnection from experience becomes a boon rather than an impediment, and can be interpreted as representing a higher order of truth than Newtonian mechanics. Quantum physics, whose intuitive form is mathematics, is thus more purely beautiful, harmonious, and complete precisely because it is free from the any connection to the material world. CONCLUSION

There are two main reasons why I have examined and juxtaposed the notions of anschaulichkeit and anschauung in this chapter. First, the contested meaning and relative importance of both concepts were part of a larger debate about which atomic theory should prevail. Both notions of anschaulichkeit and anschauung were key to the struggle between Schrödinger and the quantum camp represented by Bohr, Pauli, Heisenberg, and Born, and these men wrote and spoke about—and disagreed vehemently over—the role that these notions played in the legitimacy of their competing atomic models. For Schrödinger, who saw a direct relationship between the classical concepts, macrocosmic reality, and microcosmic reality, the connection to anschaulichkeit was clear, and customary anschauung was sufficient in order to understand his model. For Bohr, who retained the classical concepts but accorded them a symbolic value where description of quantum phenomena were concerned, the notion of anschaulichkeit was more abstract, access to customary intuition was more problematic, and the relationship between the classical concepts, description, and observation is more difficult to apprehend. Bohr is similarly more challenging than Heisenberg, who undertakes the more obvious attempt to revise altogether what counted as anschauung. My second reason has to do with the fact that the two terms are often paired, particularly by critics examining both the historical development of quantum theory and its theoretical foundations. Part of my goal in this chapter has been to challenge a too-close association of anschaulichkeit

122

Chapter 2

with anschauung, an association that runs from instances where anschaulichkeit is unproblematically translated as “intuition,” to related cases where an argument concerning one concept hangs on its interchangeability with the other. I have argued that it is vital to distinguish between the two concepts in order to understand how the terms were deployed by the physicists themselves. Schrödinger’s approach is the most straightforward and materially grounded of the three men, while Bohr’s is at times difficult to penetrate, both because of his abstruse manner of expression (which led to frustration on both his part and the part of those who studied him), and because of the great number of variables that he integrated into his approach. With respect to both anschaulichkeit and anschauung, Heisenberg required more lengthy analysis than Schrödinger and Bohr for a number of reasons: first, his thoughts must be examined in combination with the contributions made by other physicists such as Born and Pauli; second, his approach challenges the more intuitive macrocosmic experience that inflects our concepts and our modes of understanding; and third, he is the most rhetorically minded of the group, and the contradictions in his rhetoric needed to be examined. My choice to include a section on aesthetics in this chapter derives from the fact that it has an implicit connection to both the concept of visualizability and intuition, a connection that is further emphasized by the formal similarities between Schrödinger, Bohr, and Heisenberg’s individual aesthetic sensibilities, their epistemological and philosophical concerns, and their theories of the material world. For Schrödinger, this is manifested in his preference for the simplicity and straightforwardness of experimental evidence, and his rejection of any mathematical elaboration that is not grounded in the essential materiality of the object. Bohr’s sensibility leads him to favor harmonious relationships on three levels: on the quantum level with his Principle of Complementarity, which reconciles contradictory experimental results, the microscopic and the macroscopic, and classical and quantum concepts; on the level of atomic theory, where he attempts to achieve a rapprochement between the theory of relativity and quantum physics; and when he attempts to find common ground between different fields of knowledge. Heisenberg’s sensibility is both less and more abstract than Bohr’s: he speaks directly and at some length about artistic expression—particularly in music and architecture— but his formal, Platonic aesthetic reflects his investment in the abstract

The Physics of Visuality, Intuition, and Aesthetics

123

algebra of his matrix equation, and he argues that the fundamental elements of the arts are in the end expressions of the essential elements of mathematics. In all cases, because the three physicists’ approach to anschaulichkeit and anschauung was directly tied to the atomic model that they defended, examining the debate in light of these concepts inevitably sheds further light onto the specifics of these models in their scientific, epistemological, and philosophical contexts. While I have discussed each of these categories in relation to the quantum interpretation, none of them have an essential relation to quantum physics. All of them provide an opportunity to analyze in ever more subtle ways the relationships between the varied investments of scientists in their theories, ways of knowing, and human experience that reach far beyond the field of physics in particular and science in general.

3

QUANTUM PARADIGMS IN LITERARY CRITICISM

INTRODUCTION

These fictional texts clearly reveal the quantum-influenced strategy that lies … in the very fabric of the narratives themselves. —Samuel Chase Coale

In the 1980s, quantum theory emerged as an analytic tool in literary criticism within the context of a metaphysics dominated by the rigorous and complex methodology of poststructuralism. The principles of quantum theory could be used to add credence to a literary criticism influenced by the philosophical priorities of poststructuralism: the rejection of Western logos, the rejection of the subject/object duality, the decentering of author and authority, and the disruption of the relationship between signifier and signified. As literary criticism moved into the 1990s and then the twenty-first century, applying quantum physics to literary analysis served a critical culture that valued interdisciplinarity and offered a potential solution to the increasing marginalization of the humanities and the growing economic clout of the sciences. To justify dipping into a “hard science” like physics—and indeed into the most mathematically driven, difficult, and obscure field within physics—literary critics had to establish both their competency and their claim to it. To authorize their claim to quantum physics, many critics argued that only those trained in the humanities had the intellectual resources and breadth to appreciate the implications of the new physics beyond the scientific community. In Reading from the Book of Nature: Physics, Metaphysics, and the Novel, Robert Nadeau offers an example of this sort of self-sanctioning:

126

Chapter 3

In my view it might well be fortunate that the vast majority of practicing physicists do not speculate about the potential impact of the new scientific paradigm upon the construction of our symbolic universe in that the rigorous training necessary to become a physicist often precludes development of the tools of analysis and the broad based acquaintance with other aspects of culture that are necessary to engage in such speculation.1

Nadeau acknowledges that humanists need to properly confront and assimilate scientific knowledge in order to apply it in a broader context, and that there are possibilities for scientists with a liberal arts education to similarly venture into the realm of culture; however, a humanities background remains key for him. Nadeau also justifies the use of quantum physics for literary criticism by arguing that artists are hypersensitive to the “intellectual currents that eventually transform us,” which allows them—far earlier than any other group—to appreciate the implications of the new physics and to parse the cultural landscape that is implied in this physics.2 Nadeau’s argument that those possessing a poetic sensibility are more sensitive than scientists to uncertainty and paradox relies on associating scientists with objective, rational, truth-based knowledge. According to Nadeau, “quantum theoretical physicists, who are perhaps the supreme embodiment of the analytic tradition in Western thought, were highly resistant, for example, to their own denaturalization of the either/or distinction that dominated Western belief systems.”3 Here, Nadeau reveals that he is unaware of the extent to which the founders of quantum physics were actively considering the epistemological and philosophical implications of their theory, not to mention the founders’ later philosophical writings that, among other things, delved into the cultural and social context of their work. They may have mourned the loss of a straightforward, referential connection between representation and the material world, but the very nature of quantum physics forced them to leave the tidy boundaries of their discipline and actively confront the problem of language with its origins in our ordinary experience. Far from retreating into certainty and objectivity, they confronted the loss and then set out to explore the broader implications. Throughout the 1980s and 1990s, literary critics overwhelmingly favored modernist literature in their application of quantum concepts.To legitimize the relationship between modernist literature and quantum

Quantum Paradigms in Literary Criticism

127

physics, however, critics needed to establish a line of influence between the two. They had two options: either declare that the new physics had a direct influence on selected writers, or refer in broad terms to shared cultural and intellectual currents. In the first instance, a direct connection between science and artist was established according to the “hypodermic” model, wherein the artist is “injected” with scientific ideas, which then influence the narrative form, characterization, and thematics of their writing. One way of supporting this assertion was to claim that writers were familiar with the science because it was being described in the popular press. Sue Sun Yom, for example, claims that Virginia Woolf ’s tangible relationship to wave/particle duality in her work grew out of the “frequent ruminations on the nature of light, including detailed consideration of the wave-particle duality [that] appeared in such widely read newspapers as The Times and The Saturday Evening Post.”4 The potential advantage of the direct influence approach is in the clear association of a work or author with particular scientific discoveries. In fact, one need not limit oneself to ideas emerging at the same historical moment; one might show, for example, that Don DeLillo was familiar with quantum theory and this familiarity influenced how he constructed the narrative form of Underworld. The problem with the direct influence claim lies in the burden of proof: to justify analyzing a text in terms of a specific scientific discovery, the critic must provide specific evidence of a causal connection, and often the causal connection feels like mere conjecture. In fact, the claim that Woolf read or encountered British physicist and astronomer Arthur Eddington, which seems to have originated with Gillian Beer’s speculation in Virginia Woolf:The Common Ground, has increasingly become a “fact” through an iterative process wherein later critics either cite Beer’s claim or cite someone citing Beer, leaving out the speculative nature of her original claim.5 Holly Henry alludes to, but does not quote, passages from Eddington’s popularized books on astronomy that occur “nearly verbatim” in The Waves and The Years, highlights a reference to British physicist, astronomer, and mathematician James Jeans in Between the Acts, and identifies points where Woolf mentions James Jeans’s more philosophical musings in her diaries (“Do you know what Jeans says? Civilization is the thickness of a postage stamp on the top of Cleopatra’s needle”).6 She also observes that Leonard Woolf reviewed Jeans as editor of Athenaeum.7 She admits that there exists no evidence

128

Chapter 3

that Woolf ever met Jeans, but more problematic is the absence of any real evidence that Jeans’s science directly influenced Woolf ’s narrative form, style, characterization, point of view, or voice.8 Certainly there was nothing that would justify the claim that “Jeans had captivated … Woolf ’s imagination.”9 The alternative is to suggest a more diffuse connection that emerges from a shared epochal gestalt that traverses all forms of expression— literature, science, architecture, medicine, law, and so on. Civilization, like all processes in nature, is a complex interplay of forces, and it is impossible to isolate one idea and identify in specific terms its relationship to another. Nadeau, for example, answers the critique about claiming a direct causal link between science and literature by asserting that “change as it is conceived in quantum physics [i]s analogous to that form of change on the macro level called civilization,” and that we can see the influence of a shared intellectual climate.10 He see linkages in diverse products of a given culture that closely resemble one another in form and meaning in spite of the fact that they might be produced by individuals who had no direct contact with one another’s art or ideas. One can easily make the case, for example, that the novelists would not have been receptive to the ideas from the new physics if other nonscientific developments, like the romantic movement in philosophy and literature, had not been a force in culture, and also that the course of science itself would have taken different directions without such influences.11

Thomas Bohnenkamp similarly argues that each age possesses a “cultural continuity,” so that “the thinkers in any time share certain assumptions about the laws that govern their particular reality.”12 Drawing on the metaphor of quantum field theory (for which Heisenberg laid the foundation), Katherine Hayles rejects any argument implying that one must establish evidence of direct influence between science and other modes of expression in order to link the two conceptually: “to suppose that they require direct lines of influence is to be wedded to the very notions of causality that a [quantum] field model renders obsolete. A more accurate and appropriate model … would be a field notion of culture, a societal matrix … that makes some questions interesting to pursue and renders others uninteresting or irrelevant.”13

Quantum Paradigms in Literary Criticism

129

Finally,Yom claims that “the discourse of light … appeared in widely scattered fields of study, such as the nascent discipline of particle physics. … As part of this pervasive intellectual discourse centered on the metaphor of light, Woolf and members of her literary circle developed deep scientific interests.”14 The advantage of this gestalt approach lies in the fact that one does not need to prove that a writer actually encountered, let alone incorporated, specific scientific concepts. One can argue instead that there exist parallel logics, epistemological concerns, and metaphysical assumptions between science and literature that result from their shared historical moment. This approach is much more convincing; the only disadvantage might be the temptation to associate with the new physics works that simply happen to be written at the same time, making the critic more likely to impose the relationship on the text through their interpretation of it. Yom’s claim, for example, that the circulating “metaphor of light” inspired in Woolf and her contemporaries deep scientific interests appears somewhat arbitrary. Finally (and this applies as well to the direct strategy, although perhaps less so), other critical paradigms or thinkers— for example, Freud or Bergson in the modernist period, and deconstruction or continental feminism in the postmodern period—may be equally or more appropriate and productive than quantum physics as tools of analysis. This last concern proves to be a pervasive problem in “quantum” interpretations of modern and postmodern literature, particularly when these interpretations colloquialize quantum terminology and detach it from the original quantum concepts and the atomic behavior that physicists such as Schrödinger, Bohr, and Heisenberg were trying to describe. WAVE/PARTICLE DUALITY, CHARACTER, AND NARRATIVE FORM

Indeed, Mrs. Ramsay’s is the most fluid consciousness of the novel; hers is the single sustained wave-impulse … against which all that is particular is set. —Miriam Clark

Wave/particle duality is by far the most popular metaphor that is derived from quantum physics and applied to literary criticism, and is most often applied to modernist literature. A thorough survey of articles and books

130

Chapter 3

that use quantum concepts in their analysis turns up an interesting artifact: the works of Virginia Woolf are far and away the most popular objects of analysis with respect to wave/particle duality. This preference may in part result from the way Woolf undermines the boundaries between subjectivity and external reality. It may result from the difficulty of Woolf ’s narratives, in the sense that such complexity is assumed to require an equally complex interpretive paradigm. Most likely, it results from the same sort of critical iteration that followed Beer’s association of Woolf with Eddington: critical momentum has built up around the association of Woolf with waves and particles, and that momentum has been carried forward as ever more works by Woolf are run through the wave/particle interpretive machine. Critical associations of Woolf ’s work with wave/particle duality often display little understanding of what wave/particle duality actually means. What Bohr had in mind with his Principle of Complementarity was a full description of nature that combined the apparently mutually exclusive wave and particle behavior.This notion of inclusiveness—in the sense that the fullness of description is defined as that which encompasses the “both/and” of complementarity—is lost when critics separate out texts or elements of a text and identify them as being either “wave-like” or “particle-like.”Where narrative form is concerned, critics are particularly likely to invoke the “wave-like” nature of Woolf ’s writing. In “Consciousness, Stream and Quanta in To The Lighthouse,” Miriam Clark writes,“The grammar of the passage, with its repetitions, its undertow of subordination, its tidal change of direction, is itself wave-like.The wave-patterns of the novel, like those formed by the diffracted light of physics, represent whole and potent modes of seeing and hearing, thinking and acting.”15 In “Relativity and Quantum Theory in Virginia Woolf ’s The Waves,” Ian Ettinger observes, “[i]n this earlier novel [The Waves] Woolf begins incorporating wavelike structures into the narrative, illuminating particular crests of action and plunging into a ‘trough’ of time.”16 In “Wave Theory and the Rise of Modernism,” and also with reference to quantum physics, Gillian Beer argues that “assonance, overlap between words, iteration and internal rhyme all, in that work, express the wave-like fluidity of a newly imagined universe.”17 Because these interpretations have little connection to the actual quantum wave/particle duality, there exists a tendency to fix on the word “wave,” empty it of anything but its colloquial meaning,

Quantum Paradigms in Literary Criticism

131

apply it in a general way to Woolf ’s style, and elaborate by including our notions of waves—tides, or troughs, or undertows. While there is no relationship between the concept as it appears in quantum physics and the “waviness” of Woolf ’s novels, each of these writers makes specific reference to quantum wave/particle duality as part of their interpretive paradigm. Presumably, they do so in the hope of creating a generative link between two divergent modes of thinking, and perhaps more importantly, of offering a novel way of interpreting texts that have already been thoroughly filtered through a great number of other existing interpretive paradigms. There is another reason, however, that relates once again to the founding language—in this case, Bohr’s reluctant retention of the classical concepts “wave” and “particle.” That they are presented as an either/or duality makes it easier to peel off one of the terms, vacate any relationship to physics—classical or otherwise—and then use it in its most conventional and ordinary form. Most interpretations of Woolf that apply the notion of wave/particle duality can be reduced to the identification of a generic duality in the text that is then mapped onto the “quantum” opposition between waves and particles. In the case of Woolf, the primary lens through which duality is viewed is gender. Critics attentive to gender in Woolf tend to invoke the wave/particle duality in a general manner to refer to the opposing characteristics of female and male: on the one hand, Woolf ’s women are more “wavey,” in the sense that their identities are more fluid and expansive, and they connect more to others. Her men, on the other hand, are more “particle-like” because they are more individualistic and selfcontained, logical rather than intuitive. This gender duality is frequently associated with To the Lighthouse—in particular, the opposition between Mr. and Mrs. Ramsay. Miriam Clark, for example, associates Mrs. Ramsay with a “wave consciousness” or “wave impulse … against which all particularity is set.”18 Clark goes on to discover in Mrs. Ramsay “a rich and menstrual association with the moon and tide … in contrast with the barren particularity of her husband, whose ‘fatal sterility,’ whose ‘arid scimitar,’ are set against her own ‘delicious fecundity.’”19 Paul Tolliver Brown, who similarly associates Mrs. Ramsay with waves, observes, “Mrs. Ramsay is rarely alone or isolated and never questions the existence of her surroundings. However, her intuitive sense of permeable boundaries between subjects and objects

132

Chapter 3

carries with it the risk of a loss of a definable self and the possibility of total submersion into the fluid world.”20 While Mr. Ramsay is overly logical and bounded, Mrs. Ramsay is so entirely “fluid” that her identity is at risk of drowning in the people and settings that surround her. A number of problems surround these quantum interpretations of gender in Woolf. There is no question that Woolf was very much concerned with gender relations, and her work often reflects on the differences between a female and a male way of being in the world. There appears to be no need, however, to refer to wave/particle duality in the quantum sense to talk about fluid, relational, intuitive female nature versus bounded, isolated, logical male nature—these associations are already part of the received cultural lexicon that opposes the feminine to the masculine. Women are already seen to be more intuitive and to have a weaker sense of self; menstrual fluid and fecundity are commonly associated with lunar rhythms, and lunar rhythms with tidal flow.21 Furthermore, opposing wave-like behavior and particle-like behavior, and associating them with opposing gender characteristics loses the sense that the same phenomenon (or man, or woman) can express both wave-like and particle-like qualities, and thus only contributes to reifying essentialist male/female oppositions that obstruct a more nuanced analysis of the characters in To the Lighthouse.22 Consider, for example, this interaction between Mrs. Ramsay and her daughters: She [Mrs. Ramsay] was now formidable to behold, and it was only in silence, looking up from their plates, after she had spoken so severely about Charles Tansley, that her daughters, Prue, Nancy, Rose—could sport with infidel ideas which they had brewed for themselves of a life different from hers; in Paris, perhaps; a wilder life; not always taking care of some man or other; for there was in all their minds a mute questioning of deference and chivalry, of the Bank of England and the Indian Empire, of ringed fingers and lace, though to them something in this of the essence of beauty, which called out the manliness in their girlish hearts, and made them, as they sat at table beneath their mother’s eyes, honour her strange severity.23

As is typical in Woolf ’s work, gender differences are less about essentially masculine or feminine qualities, and more about the limits imposed on women by a male-dominated world—limits that Mrs. Ramsay’s daughters fantasize about transcending. Even Mrs. Ramsey is described as “formidable” and “severe”—hardly “wave-like” qualities, and it is her

Quantum Paradigms in Literary Criticism

133

formidable nature that inspires her daughters to wish for more than she has. Even more radical is Woolf ’s paradoxical assertion that the ringed fingers and lace, which her daughters associate with the essence of beauty, inspire the “manliness in their girlish hearts.” A correct use of wave/ particle duality might have captured the paradoxical fact that the daughters’ hearts are both manly and girlish, and that the opposition between men and women is not ontological, but rather constructed and reinforced by men, such as Mr. Ramsay, who see their wives as irrational. Critics who analyze The Waves do a better job of appreciating the fact that wave/particle duality exists within the object rather than between two opposing objects, although the larger problems concerning the metaphorical use of wave/particle duality persist. Ettinger, for example, argues that the “fluidity of consciousness” possessed by the characters in The Waves “correlates to wave-particle duality, in which seemingly separate entities (in this case embodied personalities) become ‘flowing’ and indeterminately bounded. The characters’ alternating sensations of atomization and fluidity reflect the strange predicaments of a quantum universe in which the individual body is actually a flux of particles and waves.”24 Ettinger goes on to observe that “as the novel’s six characters fluctuate between particulate and wavelike experiences of reality, the jarring discontinuities between unification and separation make it difficult for them to exist definitively in either state, and in the end this tension remains unresolved.”25 For Ettinger, wave/particle duality refers to an instability or “fluctuation” within the quantum particle that stands for a fundamental instability in the characters’ identities and relationships to their environment. Ettinger does capture the sense of an internal duality as well as the alternation between two points of view. His reference to a quantum universe seems unnecessary, however, and his suggestion that particles are inherently unstable abandons the more appropriate notion of an alternating point of view that is connected to what we observe about the world, rather than how it really is. Even in this revised form, the quantum concept contributes little to his basic claim, which in this case is that that the characters experience an unresolved oscillation between unification and separation. Not surprisingly, wave/particle duality proves to be a seductive notion for those analyzing Orlando, and here the original sense of the concept is

134

Chapter 3

retained. Of all of Woolf ’s novels, Orlando inspires the greatest number of interpretations that retain the notion that a single element can express two different states. According to Renee Benham, Victoria Smith’s assertion—that Orlando is neither a man nor a woman but both— “mirrors the duality present within physical matter evidenced by the double-slit experiment.”26 Benham goes on to observe that the “duality of light to behave as both a particle and a wave and the paradox of altered and unpredictable behavior after observation is similar to Woolf ’s representation of sex and gender duality within Orlando.”27 Benham recognizes the centrality of observation in wave/particle duality, although the reference to “altered and unpredictable behavior after observation” is a misrepresentation, since the behavior of a particle is precisely determined after observation. The larger problem, however, concerns the introduction of confusion rather than clarification when she imports the observer effect into literary analysis. Several questions arise: how does “after observation” apply to the novel? Observation by whom? The other characters? The reader? Is Orlando only both sexes when he/she is not observed? Again, observed by whom? Benham illustrates one of the problems with using a quantum concept to interpret literature: the critic allows the “scientific” feel generated by the metaphor’s source to stand in for clarity at the level of analysis. As well as character analysis, wave/particle duality is applied to the narrative form of modernist literature. Modernist literature in general is frequently associated with a wave logic, but there exists a number of other ways in which wave/particle duality is invoked. It may be seen to represent a duality within the text, a duality between two different types of modernist literature, or between two different critical approaches to the text. Identifying the duality between different kinds of texts is designed to provide a critical guide for approaching specific works: some modernist texts/authors are more wave-like, some are more particulate, and the goal is to identify where the writers fall in relation to this opposition so that one may apply the correct critical approach. Daniel Albright, for example, associates a duality between two types of texts with a duality between two critical approaches, and asserts that it is possible—if “tricky”—“to map certain authors according to the wave and particle models.”28 “Particulate” modernist poems contain “hard palpable balls of meaning, and promote a scientific view,” while “wave” modernist poems

Quantum Paradigms in Literary Criticism

135

“refuse to disengage themselves cleanly from the surrounding tissue of words,” so that “the meaning is distributed rather evenly over the whole text”—indeed, over the poet’s entire canon.29 Thus, “Pound, with his images, vortices, hard bits of rhythm, was a researcher into elementary particles … whereas Lawrence was a student of feeling-waves, the radiations that interconnect man, woman, snake, cow, moon, sun [emphasis in original].”30 After having established the texts or authors as either wave-like or particulate, Albright applies wave/particle duality to elaborate on the appropriate approach to a given text. In “particulate aesthetics,” the reader attends to the “details of syntax, alliteration, vowel-sequence, ictus, and other minute textual gestures.” Focusing on discriminations and boundaries, the reader creates “a whole hierarchical literary physics, from smaller to larger units of meaning.”31 In wave-like aesthetics, which tends to abolish boundaries, “the words per se are of little account: what matters is the telepathic stream, uniting writer and reader in a state of electrical immediacy.”32 Because he uses aesthetics to describe both the narrative form of the text and a given critical approach to the text, it is sometimes unclear which one Albright means. In the end, though, Albright seems to be saying that certain texts require certain approaches: “If a poem is built around a fundamental particle, it is possible for the critic to skewer that particle and declare, this is what the poem is about. But in the poetry of the wave, all the words have roughly equal density and value, and the poem resists any attempt to abstract a single control-element.”33 The role of the critic, then, is to identify whether a work is particle-like or wavelike, and then apply either a particle-friendly or wave-friendly interpretive paradigm.The problem with this approach, which is not exclusive to using a quantum interpretive framework, lies in its tautological nature: what one chooses to highlight in any given text (the particulate; the wave-like) becomes the basis for how one categorizes the text itself, or vice versa. A better approach would be to shy away from defining texts as particulate or wave-like, and simply demonstrate what can be made of the text by applying each lens to it.What are the “particulate” elements in the text, in other words, and what are the “wave-like” elements? How do they work with or against one another, or both? Ultimately, much of the confusion associated with using a wave/ particle approach returns us to the necessity of using classical terms to

136

Chapter 3

describe quantum phenomena. This terminology remains caught up in the fact that we tend to understand “duality” as a stable opposition between two different entities or ideas (for example, good versus evil; masculine versus feminine). Less familiar is the notion that the expression of opposing qualities is not really attached to an “entity,” but rather is a function of how we interact with it.There is much interpretive potential in the wave/particle model, if one were to use it to question the notion of stable identity, to discuss the nature of metaphor, or to explore how texts or gender differences are constructed by the manner in which we “see” them. It becomes less useful, however, if one uses it to support a dualist approach that is itself constructing its object of analysis, without acknowledging that our approach both enables and constricts what we observe in the text. THE PRINCIPLE OF COMPLEMENTARITY: POETIC FORM AND HYBRID IDENTITIES

Where Bohr’s epistemological tool is the inclusive power of complementarity, Frost’s is metaphor. —Guy Rotella

When the idea of complementarity rather than duality is emphasized, it tends to be applied to poetry rather than prose. This may be because poetry is more economical and imagistic, tending toward more local juxtapositions, wherein opposites appear alongside one another. In poetry, two competing points of view can be expressed almost simultaneously, whereas in the novel the alternations are more asynchronous. In other words, the imagistic nature of the poem can present us with “mental pictures” that express multiple perspectives. Thus the notion of alternating between complementary points of view rather than choosing between opposing points of view comes to the fore. A fine analysis of how poetry expresses this sort of complementarity can be found in W. John Coletta’s and David H.Tamres’s argument that Bohr’s Principle of Complementarity finds its aesthetic complement in Robert Frost’s poem “For Once, Then, Something,” where “readers are shown two mutually exclusive modes of perception: a mirror (‘shining surface picture’) and a window (‘too clear water’).”34 Integrating Heinz Pagels’s explanation of the

Quantum Paradigms in Literary Criticism

137

Principle of Complementarity with their interpretation of the poem, Coletta and Tamres write, “quite literally, in the image Frost has created, there exist complementary properties [mirror and window] of the same object of knowledge [water], one of which if known will exclude knowledge of the other.”35 Frost’s knowledge of “mirror reality” excludes knowledge of “window reality” and vice versa.36 This interpretation nicely acknowledges Bohr’s original meaning: that we can observe (have knowledge of) particulate or wave-like expressions, but never at the same time.Throughout, Coletta and Tamres demonstrate a command of the several quantum concepts to which they refer, frequently either citing the originators of the concept, or informed explanations by fellow physicists. Their application of Bohr’s principle here is both accurate and productive. Their introductory explication of the mutually exclusive nature of wave and particle in physics deepens our appreciation of the tension between two mutually exclusive images in Frost’s poetry, and they allow themselves to linger on this central aspect of the poem long enough for it to enrich our understanding of complementarity in quantum physics. Burt Kimmelman similarly refers directly to Bohr’s Principle of Complementarity in his analysis of a poem from George Oppen’s 1934 Discrete Series. Like Coletta and Tamres, Kimmelman offers a subtle description of complementarity, and one that acknowledges its fundamental origins in our knowledge of things, rather than in things themselves. Kimmelman observes, for example, “that [in] the theory of complementarity, a theory that understands the truth of things as lying somewhere between two adjacent or, more to the point, two complementary stipulations of ‘fact’ … the real, the truth of things, can never be articulated [emphasis in original].”37 Kimmelman’s observation that the only truth of things lies in two complementary “stipulations” recognizes the fact that external reality is only present in the promise/premise contained in our description of it. In articulating Oppen’s philosophy on the relationship between words and world, Kimmelman associates complementarity’s fundamental premise that truth is provisional and that the world “both includes and excludes language” within Oppen’s poetic process, wherein “Oppen must attempt to invoke the world—to bring the world into his poem—yet his language-as-world cannot simply be words.”38 Kimmelman suggests here that Oppen’s poetry expresses the tension between “reality” and language,

138

Chapter 3

where language affects our perception of reality, which nevertheless cannot be reduced simply to a linguistic artifact. Kimmelman also reframes Rachel Blau DuPlessis’s “non-quantum” interpretation of Oppen in a manner that adds a layered nuance to DuPlessis’s analysis. He primes his quote from DuPlessis with a directed explanation of Bohr’s Principle of Complementarity that highlights the aspect of Oppen’s poetry he wishes to demonstrate: the fact that “one truth ‘complements’ another; actual truth exists somewhere indeterminately within their resultant complementary relationship.”39 He continues: “As DuPlessis has observed, Oppen’s ‘poetry constantly expresses a dilemma—in the original sense of that word: dilemma of two propositions. In the prototypical meditative drama of the poetry, Oppen will hold one position, weigh it, test it out, and then become compelled to consider an opposing proposition, equally true, which counters the first one.’”40 By positioning DuPlessis’s analysis within the Principle of Complementarity, Kimmelman gives it new life, creating a productive resonance between the quantum dynamic and poetical form, and in the process deepening our understanding of both.41 Rebekka Edlund’s use of complementarity in her analysis of hybridity and “creolization” in Wilson Harris’s The Carnival Trilogy proves to be the exception to the association of complementarity with poetry. Edlund’s use of complementarity is weaker than those critics previously cited, partly because she relies on a single secondary source—physicist Nick Herbert—whose reading of quantum concepts is already very idiosyncratic, and then adds her own spin to the especially peculiar aspects of Herbert’s interpretation. She begins by citing Herbert’s claim that the quantum world consists of tendencies wherein “everything remains strictly in the realm of possibility.”42 So far, this seems a legitimate reference to the fact that the description of quantum phenomena is often limited to the identification of probabilities. Edlund veers off track, however, when she introduces Herbert’s claim that the whole world is composed of “quantumstuff,” which he describes as “a physical union of particle and wave.”43 There is no “physical union” of particle and wave— only an alternation of perceived behavior entirely dependent on whether we chose to observe—or not observe—it on its path. It is the association

Quantum Paradigms in Literary Criticism

139

of wave/particle complementarity with “quantumstuff,” however, that ends up leading Edlund down the rabbit hole. She writes: I spent quite some time trying to imagine this simultaneity of particle and wave, and of course it is practically impossible. For the purpose of reading The Carnival Trilogy, however, I find it is useful to visualize the concept of “quantumstuff ” as a big fluid, and malleable puzzle, or an infinite number of tiny elements that can be randomly disconnected and reconnected. The narrative consequence of this malleability, which applies to space as well as time, is that linear storytelling becomes obsolete.44

Edlund would have done better to leave out the notion of “quantumstuff ” altogether, for in her attempt to visualize it, she abandons any connection to complementarity or wave/particle duality, and ends up offering an impressionistic set of images (“fluid,” “puzzle,” “tiny elements that connect and disconnect”). How these images cohere into a metaphor for the obsolescence of linear storytelling is difficult to discern. When Edlund does leave the notion of quantumstuff behind, her interpretation becomes more legible, although significantly, she abandons the key notion of alternation in favor of true simultaneity when she introduces the notion of a “physical union of wave and particle.” Her observation that the wave/particle model transcends the logic of either/ or provides a foundation for her claim that complementarity models a conception of multiculturalism that transcends “the need for hierarchizing or choosing.”45 Edlund then extends the metaphor to drive home her main point: that Harris’s narrative moves identity beyond the “binary opposition between Western and African heritage” into concepts such as “creolization” and “interculturation” that allow for the possibility of a unified rather than fragmented postcolonial identity.46 According to Edlund, dispensing with oppositions leads to a situation in which “the individual no longer has to decide whether he or she is African, European, Asian, or Amerindian. He or she can be all at once, none of the above, and/or some of these things some of the time.”47 Edlund then concludes by specifying the linkage between complementarity and postcolonial identity: In quantum theory, however, with its “Principle of Complementarity,” neither the particle nor the wave state of the electron is complete in itself—both are necessary to give us a complete picture of reality. This provides us with a

140

Chapter 3

whole different basic understanding of the universe. Suddenly, what Homi Bhabha has termed “hybridity,” seems to be the natural state of things in general. If the electron is wave and particle at the same time, it is not only possible, but even natural for a person to feel, say, African and European and Asian at the same time [emphasis added].48

In naturalizing complementarity, and then applying it at a macrocosmic scale, Edlund tends to lose sight of the fact that it is already a metaphor, which leads her to claim that quantum behavior of the electron provides not merely a conceptual model, but rather ontological proof that “hybridity” is an innate quality of the universe. The second naturalization, this time of hybridity, glosses over the extent to which all identity is culturally constructed and threatens to install a new fundamentalism around identity. At the same time, Edlund ignores the fact that wave or particle behavior depends on whether a measuring apparatus is used—in other words, whether one interferes with it through the act of observation. If she were to acknowledge this fact, she would be forced to conclude that a person’s identity is created by how they are perceived. But this is precisely the problem with the colonial dynamic that she hopes to challenge: subjected to the colonizer’s gaze, the colonized are (re)created in the image of the colonizer’s reductive, essentializing fantasy. Similarly, if she were to use the term “complementary” properly, she would be suggesting the necessity of alternating between identities—again, precisely the imposition that notions such as hybridization and creolization are designed to challenge. Part of the effectiveness of using the Principle of Complementarity to analyze poetry lies in the careful way in which the critics of poetry that I encountered took the time to familiarize themselves with the concept; however, part also appears to lie in the appropriateness of the comparison between poetry and complementarity—particularly when it is limited to formal analysis. Bohr said that both wave-like and particle-like behavior must be included for one to get a full picture, but the mathematics and experimental evidence proved the fact that whichever behavior was manifested was entirely dependent on the presence or absence of an observing device. The best interpretations, then, appreciate the fact that wave/ particle behavior is not about the fact that that some things (characters, or works) are waves, and some are particles, but rather that the same element can be both, depending on how you look at it.

Quantum Paradigms in Literary Criticism

141

THE UNCERTAINTY PRINCIPLE: LANGUAGE AND INDETERMINACY

Heisenbergian indeterminacy applies to both physicist and poet: the physicist is a prisoner of matter and the poet is a prisoner of language: neither can get outside of his medium. —Steven Carter

While not as popular as wave/particle duality, the Uncertainty Principle is nevertheless one of the more active metaphors in literary criticism, perhaps because the term is even more generalizable than wave/particle duality. One tends to see this term associated with postmodern rather than modern literature; works of modernist literature, for all their complexity and disruption of linearity, tend to be ordered, closed systems. In their juxtaposition of opposites, they are amenable to the notion of duality, but they do not ultimately express the kind of radical undecidability that characterizes postmodern literature. Few critics using the Uncertainty Principle trace it back to its origins—in this case, the fact that it is impossible to measure conjugate states simultaneously. In applying the Uncertainty Principle, critics tend to introduce semantic slippage between “Uncertainty” (in the quantum sense), “uncertainty” (in the generic sense), and the postmodern narrative’s indeterminateness, undecidability, fragmentation, and resistance to closure. There do exist, however, some effective applications of the Uncertainty Principle. Along with applying the Principle of Complementarity to Oppen, Burt Kimmelman makes liberal use of the Uncertainty Principle. Unlike solidly modernist writers such as Virginia Woolf, George Oppen stands at the border between the modern and the postmodern. When Kimmelman analyzes Oppen in terms of complementarity, he is viewing Oppen through the lens of modernism, and when he associates Oppen with the Uncertainty Principle, he brings to the fore Oppen’s postmodern qualities. This second emphasis is expressed in the fact that Kimmelman introduces the Uncertainty Principle when he compares Oppen to Robert Creeley, who was a central member of the Black Mountain poets, and who was influenced by Charles Olson—the first American poet to label himself postmodern. Of the similarity between Creeley and Oppen, Kimmelman writes:

142

Chapter 3

The two poets’ intuition of language is strikingly sympathetic to the problems confronted by quantum physicists and mathematicians who in the early twentieth century anguished over the fact that the reality they were detecting was not expressible in words. Heisenberg went so far as to construct a mathematics unconnected to any picture of what existed in the physical realm in order to work out his Uncertainty Principle.49

Kimmelman argues that, for Oppen, words in a poetical context “are a reality that cannot be comprehended, as if marked by an indeterminacy such as Heisenberg assigned to electrons whose position and velocity might not be ascertained at one and the same time.”50 As with his use of the Principle of Complementarity, Kimmelman demonstrates a sound understanding of the Uncertainty Principle, including a rare appreciation of the extent to which quantum physicists agonized over the difficulty of expressing their discoveries in language. This opens the door for Kimmelman to focus on issues of representation, and “the inherent metaphoricity of words, their power to turn something into what it is not.”51 While the “metaphoricity” of language troubled Heisenberg, precisely because it turned something into what it was not, in the context of poetry, the generative potential of the metaphor can serve to enrich the work. Once again, Kimmelman uses a quantum concept to good effect, making a persuasive link between poetic language and the epistemological challenges faced by quantum physicists. Kimmelman’s analysis of Oppen in relation to complementarity illustrates that successful use of a quantum phenomenon to interpret literature requires a thorough understanding of its conceptualization in its original context. His use of the Uncertainty Principle, however, illustrates that a truly original and provocative use of the concept can emerge from an appreciation of those of its aspects that are typically passed over—in this case, the material impact of language. In his analysis of Paul Muldoon’s poetry, Wayne Kobylinski similarly uses the Uncertainty Principle effectively to note a unique tension in Muldoon between a highly structured poetic form and radical uncertainty. Kobylinski observes: “Just as, for all its notions of built-in uncertainty, the world described by quantum mechanics operates in a mathematically demonstrable manner, the world described by Muldoonian poetics represents through formal structures an inherent order with uncertainty built into it.”52 Here, Kobylinski references another typically

Quantum Paradigms in Literary Criticism

143

ignored tension that appears in the Uncertainty Principle itself: the fact that uncertainty is expressed through a very precise mathematical structure. More than this, as Kobylinski implies, quantum physics does not describe a chaotic world, but rather an ordered world that has uncertainty built into it. He then uses this tension between structure and uncertainty as a metaphor for illustrating how Muldoon imposes highly constraining formal techniques such as double sestinas, terza rima haikus, and restrictive rhyming schemes “to express features of the observed world that contradict the fixity inherent in those techniques.”53 Like Kimmelman, Kobylinski pays particular attention to the restrictive nature of conventional language; he then connects this to the impact of the observer in order to illustrate Muldoon’s consciousness of the possibilities produced by formal precision, wherein “solving for one variable only accentuates the inability to pin down another.”54 Many critics of the postmodern who purport to be using the Uncertainty Principle fix on the word “uncertainty” and then apply it in a generic way to the aesthetics of a literary work. This tendency deprives their analysis of a productive cross-fertilization between quantum physics and literature, and renders their reference to the physical principle arbitrary and unnecessary. Susan Strehle’s analysis of Donald Barthelme offers a prime example among many of how “uncertainty” or “indeterminacy” gets detached from the Uncertainty Principle. In selecting Barthelme to illustrate how the Uncertainty Principle plays out in a work of literature, Strehle appears to be right on the mark—in a rare critical windfall, Barthelme actually invokes the Uncertainty Principle with reference to his writing. During an interview with J. D. Harris later published in the Paris Review, Barthelme stated, “In this century there’s been much stress placed not upon what we know but on knowing that our methods are themselves questionable—our Song of Songs is the Uncertainty Principle.”55 Unfortunately, Barthelme’s association of the Uncertainty Principle with “knowing that our methods are themselves questionable” already misses the mark, but it is Strehle who empties it of meaning when she observes that Barthelme’s Paradise fictionalized the Uncertainty Principle in its “looseness, openness, and uncertainty.”56 Of Barthelme’s characters, Strehle observes that they “remain uncertain how to construct or express intangible values. They speak ironically, obliquely, metaphorically—like

144

Chapter 3

Barthelme’s fictions themselves—because saying the not-true is, for subscribers to the Uncertainty Principle, the closest approximation to cloudy ‘truth.’”57 According to Strehle, the ironies that abound in Barthelme’s works leave the characters “suspended in uncertainty,” and the novel, in its “open-endedness,” leaves both the characters and the readers “in the midst of uncertainty.”58 Uncertainty in these passages is associated, variously, with narrative open-endedness, lack of conventional structure (looseness), confusion on the part of the characters, and a lack of surety (not-knowing). It is a product of obliqueness, irony, and indeed of metaphor itself. Strehle in fact offers a very good reading of how, in Barthelme’s Paradise, irony undermines the presumptions of the realist narrative with its causal logic and representation of stable selfhood. Her understanding of how irony functions is both subtle and thorough. She also begins with an excellent account of the Uncertainty Principle, and she has grounds to use it with respect to Barthelme, given his own statement. Unfortunately, when she attempts to integrate the principle into her textual exegesis, all of its specificity is lost, and the word “uncertainty” is reduced to nothing more than its conventional dictionary definition. This general use of the term is what inspired Sean Kinch’s observation that Strehle’s “quantum adjectives” are too vague, and that they could be applied to countless literary texts that appeared long before the quantum revolution.59 This tendency to tack the word “uncertainty” onto analyses that focus on textual indeterminacy and the undermining of a mastery-based knowledge economy, and to then invoke the Uncertainty Principle, is by no means unique to Strehle. Guy Rotella, for example, falls back on the same strategy in his discussion of Frost’s indeterminacy in “Comparing Conceptions.”This strategy does not necessarily undermine the overall validity of the analysis, although some of the more misinformed attempts by critics to apply the principle lead to incoherencies, as in Jorge Carrion’s discussion of Heisenberg and James Joyce in “The Bicephalous Writer: The Commingling of the Creative Writer and the Critic in a Single Body.” Here, its use lends a merely surface complexity to the analysis, while at the same time introducing unnecessary and ultimately unproductive detours. In his critique of how the methodologies of poststructuralism—in particular deconstruction—have assimilated post-Newtonian scientific

Quantum Paradigms in Literary Criticism

145

principles, James Solomon cautions against the tendency of poststructuralist criticism to associate the uncertainties of scientific investigation with the aporias of aesthetic inquiry. In particular, he points out that “critical uncertainty and quantum uncertainty are not equivalent: the former refers to aesthetic indeterminacy, while the latter refers to a specific interpretation of atomic measurement and experimentation.”60 While the Uncertainty Principle can be used effectively in analytic literary aesthetics, Solomon nicely sums up the problematic way in which “uncertainty” is often used in an indiscriminate way as a word-bridge between quantum physics and literary form, a strategy that either leads to incoherencies or to a situation wherein the term is so diluted as to stand for nothing other than a general lack of definiteness. Solomon’s primary point is to insist on the fact that our understanding of the world is referential, and “is grounded in the real behavior of nature itself and not simply in the figures of metaphysics.”61 He thus warns of using quantum theory, and the Uncertainty Principle in particular, as a justification for poststructuralist epistemology and critical methodologies “intent of demonstrating the absolute uncertainty and indeterminacy of textual interpretation.” Solomon explains: Often, in fact, such scientific developments are used to defend deconstructive critical practices. This has been the case especially with Werner Heisenberg’s famous statement of scientific “uncertainty” in the fact of the quantum atom. This statement has been taken as a general metaphor for a global epistemological uncertainty that prevents any objectively determinate knowledge of the world.62

Part of the problem, according to Solomon, is again in the misinterpretation of the Uncertainty Principle, which is not about “the indeterminacy and subjectivism of our knowledge of reality,” but rather about “the means by which its potential behavior may be calculated and known.”63 Christine Froula’s “Quantum Physics/Postmodern Metaphysics: The Nature of Jacques Derrida” appears to provide an ideal opportunity to evaluate Solomon’s more epistemology-oriented critiques.64 In the depth and complexity of her analysis of Derrida, as well as in her detailed, informed descriptions of quantum phenomena and relativity, Froula’s essay on Derrida is in many ways a tour de force. Froula so tightly weaves

146

Chapter 3

together the various concepts of the new physics with her analysis of Derrida’s critique of Western logos that it is difficult to separate them out and still do justice to her interpretation.65 Nevertheless, I will attempt to isolate her use of the Uncertainty Principle to discuss Derrida’s notion of “trace.” As Froula points out, Derrida’s “trace” challenges a logocentrism based on truth, stable meaning, and “being as presence.” This assumption of being as presence is in part a function of the belief in language as a system that refers to external truths and stable meanings. For Derrida, what replaces presence is a free play of meanings wherein the sign never leads to the extra-linguistic “thing,” but rather to another sign, so that each element of writing is constituted on the basis of the trace within it of the other elements of the chain or system. Froula, like Kimmelman and Kobylinski, is one of a select number of critics who recognize and capitalize on the fact that the quantum interpretation confronts the limits of language and representation. Froula points out that Heisenberg … suggested that the problem lay not in the theory but in a fundamental discontinuity between the mathematical language in which the theory was stated and the language used to express observational findings in classical physics: “For the time being, we have no idea in what language we must speak about processes inside the atom. … I assume that the mathematical scheme works, but no link with the traditional language has been established so far.66

In highlighting Heisenberg’s Uncertainty Principle as a problem of language, Froula departs from the typical association of the Uncertainty Principle with some generalized form of aesthetic or epistemological indeterminacy. This focus on the problem of language enhances her assertion that quantum physics is the “scientific prototype” of “Derrida’s critique of the Western writing founded upon a belief in the logic and reason that attends linear writing/language.”67 Just as quantum physics undermines the referential assumption and investment in rationality associated with classical concepts, so does Derrida’s model of writing as an immanent “play of differences,” wherein there can be no “dualistic conjoining of a sensible signifier with an intelligible signified,” challenge the assumption of rationality and referentiality of Western logos.68 To reframe Froula’s observations in the language of my first chapter, I return to my discussion of conceptual metaphor theory and the problem

Quantum Paradigms in Literary Criticism

147

of the referent in quantum theory. Derrida would, of course, see Conceptual Metaphor Theory’s argument that language is based on embodied experience as a part of a misguided belief in presence. I would respond that, while writing may end up being an immanent play of differences, figurative aspects of language emerge from, but are not defined by, our ordinary sense-based experience. Furthermore, when using language in its conventional, everyday communicative role, we necessarily contract to assume at least a provisional referentiality. Our dependence on this contract is precisely why language fails in quantum physics, and supplies the context for Arthur Eddington’s comment that we have yet to invent a mode of narration that could describe quantum physics.This, then, is the link between quantum physics and Derrida that Froula’s article rightly points out: quantum physics dispatches the classical, everyday assumption of a stable referent and undermines the notion of presence. The consequence is both restrictive and (not necessarily desirably) productive, as far as communicating its concepts goes. The rejection of classical terms virtually silenced Heisenberg, who retreated to the mathematics of his matrix mechanics, and Bohr and Schrödinger’s retention of conventional terms produced myriad latter distorted interpretations of quantum phenomena. Froula explains that quantum physics “implies not only the incompleteness and relativity inherent in any view of the objective world but the absolute, theoretical impossibility of knowing the world as ‘objective,’ and hence a denial that ‘reality,’ conceived as independent of our observation of it, exists at all [emphasis in original].”69 She explains, “it is not simply that subatomic ‘things’ have a reality that we are unable to observe because of technological limitation; rather, an irreducible uncertainty exists in the nature of things at the most fundamental level [emphasis in original].”70 That Froula associates this notion of the subatomic particle as a “non-entity” with Ernest Fenollosa’s comments on the decomposition of the noun corresponds with my observations in chapter 1 about how quantum physics challenged entity-based language and nominalism. The difference lies in the fact that my concern was with the linguistic “brick wall” that Schrödinger, Bohr, and Heisenberg ran into with this decomposition of the noun. Froula, in contrast, is suggesting how quantum physics can be made operational in narrative—here, by pointing out that the “decomposition of the noun” corresponds to Derrida’s notion of

148

Chapter 3

trace.71 At the same time, her analysis highlights the fact that there exists no straightforward mode of narration—in the sense of telling a story— that could communicate quantum concepts. One can only do so indirectly by showing, as Froula, Kimmelman, and Kobylinski do, a congruence between a poetical or linguistic system and the logic of quantum physics. Concerning the Principle of Uncertainty, the most successful critics appear to be those who anchor their analysis in questions of language and poetic form in a focused and restrained way. These critics not only demonstrate a thorough and subtle understanding of the principle, but they apply select aspects of it with precision. Kobylinksi, for example, effectively uses the tension in the Uncertainty Principle between indeterminacy and the precision of mathematics to illustrate an analogous tension in Muldoon’s poetry. A less successful application of the Uncertainty Principle occurs when critics empty it of its nuanced specificity and then use the word “uncertainty,” decontextualized and reduced in a vague way to stand for narrative indeterminacy or subjective disorientation. It is not the case, as Solomon argues, that the Uncertainty Principle cannot be imported productively to discuss aesthetics or epistemology, but rather that the integrity of the physical principle must be maintained, and the critic must clearly establish how and why defined features of the principle are appropriate to their object of analysis. SUPERPOSITION, WAVE/PARTICLE COLLAPSE, AND THE OBSERVER EFFECT: THE SUSPENSION AND COLLAPSE OF MEANING

Only an Ultimate Observer outside the loop would be able to see the result of the experiment as one unified reality or a split universe. This Ultimate Observer is ostensibly a reader who represents a Final Observation. —Jason C. Smith

A number of critics appeal to the observer effect to discuss the breakdown of the boundary between subject and object, and the transformation of the object by the perceiving subject.At times, the observer-induced wavefunction collapse is used in its traditional sense: the effect that observation has in reducing a system to a single state. At other times, critics draw on the related notion of superposition in order to discuss the prolif-

Quantum Paradigms in Literary Criticism

149

eration of meaning in language. In the first case, critics focus on the determinate outcome of observation; in the second, they focus on the indeterminate state of a system prior to observation and then demonstrate how a state of superposition endures beyond the act of observation. This second approach relies on decoherence or the “hidden variables” theory, as well as Everett’s Many-Worlds Interpretation. The outcome-oriented approach relies on a familiar misinterpretation of the quantum observer effect: that the act of observation—which quantum physics locates at the point of the detecting mechanism—really happens only once the outcome reaches the observer’s consciousness. In his discussion of the act of poetic composition, Kobylinski writes: This idea of superposition links with [Roland] Barthes’ formulation of the zero degree, which claims that the modern poetic word is “accompanied by all its possible associations. …” The Muldoonian view ascribes this state of expansive association not to poetic language but to some general realm of language that presents itself to the poet, who in turn controls or limits the possible readings. The poet thus functions as the observer who in the act of getting a fix on the appropriate meaning of the word collapses the superposition of associations.72

The system of poetic language, Kobylinski suggests, may be understood as initially being in a state of superposition, where all possible meanings exist. The poet then “collapses” these multiple associations by arranging words in a particular form and thus delimiting their meaning. Kobylinski’s use of superposition holds up; superposition does refer to all possible states of a subatomic physical system. One might question his assertion that the poet fixes meaning in the same way that measurement fixes subatomic behavior, however. This argument ignores the persistent associative nature of language, which continues to possess an excess of meaning even when specific words are arranged on the page. Kobylinski also ignores the role that the reader (as “observer”) has in defining meaning. In his reading of James Joyce’s Ulysses, Keith Booker offers a slightly different take on how the author delimits available meanings: Joyce has thrown off once and for all the notion of the distant, disinterested author, … rather, the author becomes involved in the narrative in an intimate way. Thus Joyce’s author becomes similar to Heisenberg’s observer, who

150

Chapter 3

cannot separate himself from the results of his observations. According to Heisenberg the very act of observation influences the observed results, just as for Joyce the very act of communication influences the very concepts and descriptions being communicated.73

Here, Booker incorrectly associates the Uncertainty Principle—which is concerned with how accurately we can measure the position and momentum of a particle at the same time—with the observer effect, which is more properly associated with wavefunction collapse. It is a common confusion, however, and one that even Heisenberg at times seemed to encourage. If one brackets the reference to Heisenberg and instead focuses on Joyce’s role as author, one can see that Booker, like Kobylinski, associates the author’s “interference” with choices that he makes. Booker focuses more on disturbances introduced by the act of communication, however (rather than the author’s choice of language), and in doing so he implies a more dynamic relationship between author and language. More than Kobylinski, then, Booker highlights the interactive over the restrictive aspect of the observer effect in order to emphasize how the author is an immanent part of the textual system. In his analysis of John Fowles’s The French Lieutenant’s Woman, Jason C. Smith also references the manner in which the observer “controls” the outcome of the experiment, this time shifting from the author to the reader. His understanding of the relationship between narrative and “reader as observer” are quite different than Booker’s and Kobylinski’s, however. For Smith, the observer/reader’s power to affect the meaning upsets traditional notions of authorial control, thus granting the reader an empowering and subversive role in constituting the narrative. He argues: “What Fowles achieves, then, is to put the reader (though no doubt an Ideal Reader) into the position of Ultimate Observer with access to the Final Observation. In this way, Fowles’s text offers at least the possibility of freedom from the linear deterministic narrative of Enlightenment science.”74 According to Smith, the fact that Fowles offers more than one ending forces the reader into a creative, participatory role in determining which storyline will prevail, and assigns a provisionality to the narrative that undermines the determinism of locating the meaning entirely in the text, there to be discovered by the reader. There remains an unacknowledged determinism in Smith’s model, however—but in this case the power to

Quantum Paradigms in Literary Criticism

151

determine the outcome/meaning of the narrative is relocated to the reader. In “A Place to Step Further: Jack Spicer’s Quantum Poetics,” Steven Carter makes several straightforward analogies between the poet’s investigation of language and the quantum physicist’s investigation of matter, particularly with respect to wave/particle duality and the Uncertainty Principle.75 In his article “Robert Duncan and Erwin Schrödinger: Esthetics of the Wavefunction,” Carter differs from Kobylinski and Booker in distinguishing the material world from the poetic world. Rather than presenting them as coextensive, he contrasts how the observing mechanism in physics initiates wave/particle collapse with how the reader/observer of poetry can sustain multiple meanings because of the nature of language. In his discussion of Robert Duncan’s poetry, Carter asks, “[w]here then do poet and physicist part company? Schrödinger’s cat, once the consciousness of the observer comes into play, can only be experienced as alive or dead; the reader of a poem, on the other hand, can entertain several manifestations of language in his mind, consciously and unconsciously at the same time.”76 While the observing mechanism in physics collapses the particle into a single state, the observer/reader of poetry can maintain a state of superposition, partly because of the nature of language, and partly because of our relationship to it. When he considers a particular poem, in this case Duncan’s “At the Loom,” Carter appears to argue the opposite point. He begins with the observation that “At the Loom” “is a poem that, like Schrödinger's cat, carries on a dual existence in the mind of the reader/observer.”77 Carter pursues this notion of two opposing states when he observes: The reader collapses the wavefunction of one thematic category by focusing on the other and by approaching the poem either ethically, exploring “how it means,” or ontologically, examining what it “is.”The choices the reader makes, however, can never subsume or define the poem either thematically or mechanically because Duncan insisted on the “luminosity” of the system as a whole.78

Carter suggests that the excess of meaning exists in the poem, not in the reader’s mind, and that the reader sustains a superposition of opposite meanings only until that moment when he or she formulates an

152

Chapter 3

interpretation. Carter’s final point is that Duncan’s poetical “system” escapes reduction to one state regardless of the reader’s choices—that there is an excess to the object. Here, Carter is more confusing. Because he keeps returning to Schrödinger’s thought experiment, which was in fact designed to illustrate the incoherence of superposition on the macrocosmic scale, he sees superposition as the simultaneous existence of only two states, whereas superposition actually refers to the maintenance of all theoretically possible states. This leads him to partner incorrectly quantum superposition with a dualistic choice between an epistemological approach and an ontological one. Carter is thus arguing that Duncan’s poetry escapes the deterministic relationship between observer and object in the quantum wavefunction collapse (i.e., in Duncan, the object refuses the observer’s attempt to collapse it to one state/meaning). Carter’s oscillation between the reader’s interpretive “superposition” and the poem’s superposition of meaning (which the reader collapses) compromises the effectiveness of his argument. Superposition and the observer effect are associated with the ManyWorlds Interpretation, which denies the wavefunction collapse entirely. Instead, each possible state is said to continue to exist within a potentially infinite number of different worlds or universes.When he focuses on the reader, Jason C. Smith emphasizes determinate outcomes; when he focuses on the “bifurcation” of the narrative in The French Lieutenant’s Woman into two endings, he turns to the Many-Worlds Interpretation. With reference to Schrödinger, Smith argues that the opening of the box is not the moment when a single potentiality becomes reality; the opening of the box merely marks the moment when the universe to which the observer belongs is recognized.79 Similarly, the point at which the narrative of The French Lieutenant’s Woman bifurcates is not the point at which the two endings are initiated; rather, they are introduced earlier, when Charles and Sarah encounter each other in Exeter. After that moment (in chapters 61 and 62), Sarah is absent from the narrative (in a state of superposition), and during this absence both Sarahs exist, each with their own actions and choices.80 Out of chapters 60 and 61 emerge not two versions of Sarah’s life, but rather two different Sarahs altogether, possessing different psychologies and occupying different universes. There would seem to be no reason to use quantum physics to make this point, although the narrator in the book does describe himself as a

Quantum Paradigms in Literary Criticism

153

“gamma ray particle that effects a change in the text by effecting a change in Sarah at the molecular level.”81 This allows Smith to observe that “like Heisenberg’s gamma-particle, the narrator becomes something which both observes and effects.”82 There is the Sarah that the narrator changes, and the Sarah that he does not change. One does not get pregnant, and the other does. The problem with Smith’s interpretation is that he seems to want to have it both ways: when discussing the reader or narrator, he wants to maintain the traditional observer-triggered wavefunction collapse that the Many-Worlds Interpretation rejects, while when discussing the narrative, he wants to adopt the multiverse model of the ManyWorlds Interpretation that depends on the absence (or, more accurately, the irrelevance) of the observer effect. One might imagine an interpretation that does in fact associate the narrator with one phenomenon, and the narrative form with another, but Smith does not make this potentially productive distinction. As a result, there is an unintended and unacknowledged “bifurcation” in Smith’s interpretation that makes his argument appear contradictory. Critics turn to the observer effect in search of a model that preserves a measure of interpretive or creative control (either at the level of author, reader, or narrator), while at the same time allowing for narrative experimentation and an excess of meaning. Most critics fit into one of two camps: they either favor a determinate, outcome-oriented application of the observer effect, or they favor the indeterminacy of the superposition that exists prior to the wavefunction collapse. The first delimits meaning and is typically associated with the author/poet/reader; the second emphasizes the proliferation or excess of meaning, and is typically associated with the nature of language in general and narrative form in particular.Whether they focus on superposition, or the subsequent wavefunction collapse, or the system as an interactive totality determines whether a critic associates the observer effect with sharply defined outcomes or with provisionality and indeterminacy.The relative emphasis on superposition or wavefunction collapse thus depends on what the critic wants to highlight in the work. Kobylinski’s systemic approach stresses the inextricability of the observer (author) and the observed, while Smith’s determinism stresses the empowerment suggested by the observer’s (reader’s) influence over the observed. Carter alternates between the observer’s state of superposition and the superposition of his object of analysis,

154

Chapter 3

opposing the responsiveness of subatomic systems to observation to the poem’s resistance to observation, although his ultimate conclusion seems to be that the reader/observer will never be able to collapse the poem into a single state because it possesses an ineffable quality that exceeds any single interpretation. ENTANGLEMENT AND NONLOCALITY: PROXIMATE CHARACTERS

Quantum entanglement provides Eleanor and Max a private language for describing their love. —Sean Kinch

Quantum entanglement—where an action on the state of one system (or particle) produces a response in the state of a system at potentially great distances from the first—tends to serve as a model for how characters in a novel interact with one another, particularly the manner in which their interactions or co-presence is mutually affecting. Entanglement or nonlocality appears to be vulnerable to two sorts of misrepresentation when used in this way: first, it is represented as based on proximity or direct interaction, wherein entanglement is represented as a kind of closeness or intermingling that is characterized by the loss of boundaries. This misses the whole point of entanglement: that it is interaction at a distance. The other, less fundamental error is to see entanglement as a sort of to-and-fro dynamic based on mutual impact. This interpretation discards the determinism of quantum entanglement, which is based on a unidirectional causal effect wherein a change made to one particle affects the other in an equal but opposite way. In his analysis of The Waves, Ian Ettinger argues that “in their essential fluidity, the six personalities impact and intermix with one another. By coming into proximity, they fundamentally change one another’s perceptions and senses of selfhood, correlating with Werner Heisenberg’s theory of “entanglement.”83 Ettinger’s most obvious error here is to associate entanglement with Heisenberg, when the notion of “entanglement” emerged, somewhat inadvertently, from the thought experiment that Einstein, Podolsky, and Rosen developed to demonstrate that the Uncertainty Principle was incomplete.84 The other two main figures involved in entanglement were Schrödinger, who worked on nonlocality and came

Quantum Paradigms in Literary Criticism

155

up with the term “entanglement,” and J. S. Bell, whose theorem demonstrated that there were no hidden variables involved in connecting the two particles. Ettinger also misrepresents the two basic tenets of entanglement: action at a distance, and one-way causality. Instead, the characters affect one another in undefined ways because of their “proximity” and direct “intermixing.” At another point, Ettinger observes that, when in Neville’s presence, Bernard “begins to see reality through another’s eyes.”85 Ettinger introduces a similar example when he refers to Neville’s observation, “as he approaches I become not myself but Neville mixed with somebody.”86 According to Ettinger: This suggests that one’s sense of self is not only affected by others but actually generated in relation or reaction to them. This is exemplified in moments in which the characters are experiencing “supraconsciousness” and the self suddenly coheres in the presence of a familiar face, which Bernard describes as being “contracted by another person into a single being.” Since personalities exist in a continuum, in which they both mirror and alter one another, “entanglement” often occurs as “Meeting and parting, we assemble different forms, make different patterns.”87

That a sense of self could be affected or generated by another is consistent with entanglement. The “creation” of someone through direct interaction, however, is outside of the purview of entanglement, and directly contradicts it. In entanglement, nothing is created, only altered; and the point is that the bodies affect one another, not through proximity, but rather at a distance. Again, the confusion derives from the limitations imposed by Schrödinger’s necessary use of conventional language; in its conventional usage, entanglement does mean direct interaction, and without further context, the paradoxical nature of quantum entanglement at a distance is lost, and the term reverts to its everyday meaning. Just as in his use of wave/particle duality, the manner in which Ettinger uses entanglement demonstrates slight understanding of the quantum concept. His interpretation of the way that the characters interact is certainly valid—they are indeed “entangled” with one another—but it has nothing to do with the quantum theory of entanglement. What is a decent interpretation of the intersubjective dynamics between the characters in The Waves is

156

Chapter 3

undermined in Ettinger’s effort to make the reading appear more sophisticated by introducing a distorted reading of a quantum concept. In his discussion of DeLillo’s novels Underworld and The Body Artist, Samuel Coale offers a similarly confusing application of quantum entanglement that collapses entanglement with the Uncertainty Principle in a manner that fails to do justice to either. Like Ettinger, Coale both misrepresents quantum entanglement, and detaches it from the quantum phenomenon so that it means little more than inseparability. Coale’s justification for using quantum entanglement in the first place is somewhat weak. He writes, “Patrick O’Donnell, in his article on Underworld, refers to ‘a post modernism that, in Fredric Jameson’s hands, is founded on a modernity where the entangled relationships among “self,” “object,” and “world” have undergone a fundamental change,’ the idea of entanglement being a very quantum quality.”88 Coale’s comparison rests on a simple metonymic slippage: O’Donnell mentions the words “entangled” in relation to Underworld, and Coale, keen to invoke quantum concepts in his own interpretation of Underworld, uses this passing reference to entangled relationships to justify his introduction of quantum entanglement. Coale then uses the “linkage” between entangled quantum entities to argue that “maybe nothing is separate to begin with, and perhaps an ultimate entanglement lies at the heart of the quantum flux at all times.”89 In a further remove from actual quantum entanglement, Coale then suggests that “if entanglement leads on to more entanglement, then perhaps the perpetual flux is all that can possibly exist in the microcosmic world.”90 Coale’s frequent mention of quantum flux may refer to quantum fluctuation; however, quantum fluctuation has nothing to do with quantum entanglement.91 In linking the two terms Coale can further associate entanglement with a sort of undulating, unpredictable, nonlinear, and mystical flow that he claims characterizes the three qualities expressed in DeLillo’s novels: “the randomness of existence, the mysterious nature of human perception, and the nature of time.”92 Coale exacerbates the confusion when he advances an interpretation of Louisa Gilder’s reading of entanglement: “Louisa Gilder writes, ‘In entanglement, the quantum state the particles find themselves in is indefinite—neither here nor there, neither this nor that, neither yea or nay—but if one is measured and found to be ‘yea,’ the other is ‘nay.’”93 Instead of recognizing Gilder’s main point and the main point of

Quantum Paradigms in Literary Criticism

157

quantum entanglement—that influencing one particle immediately produces an effect on the other—Coale appears to pick up on the reference to indefinite states, and then to claim that Gilder’s theory of entanglement provides more support for the quantum nature of DeLillo’s fascination with randomness, human perception, and the created narrative of time. This set of associations, followed by several more that include the collapsing entanglement and uncertainty, reaches its peak in the following set of observations: What, then, are some of the ramifications of Don DeLillo’s art in these two novels, which so depend on his own wrestling with quantum theory in terms of randomness, human perception, and the concept of time and space? Entanglement again seems the key, the quantum idea that demolishes the polarities—particle/wave—that measurement, our observation of our measuring devices, and our construction of language and logic rely upon to convey as best we can what the quantum world produces.We cannot exist without polarities, it seems, but DeLillo constantly sets them up to deconstruct them, creates them to annihilate or at least undermine them as the quantum flux theoretically does as well. From this perspective so much is entangled and ultimately ambiguous that nothing can really be separate or independent from anything else.94

First, quantum entanglement does not demolish polarities; it is founded on them. Second, quantum entanglement is not connected to wave/ particle duality. Furthermore, one wants to ask: in what way does the measurement of quantum particles (and here we are into the Uncertainty Principle) “rely on” polarities? How does quantum “flux”—where particles randomly pop in and out of existence—deconstruct or undermine polarities? And how does all of this lead to the association of entanglement with ambiguity, followed by the conclusion that “nothing can really be separate or independent from anything else”? Unlike Ettinger, who misuses entanglement but does offers a coherent, albeit nonquantum interpretation of the characters in The Waves, Coale’s collection of incorrectly described and arbitrarily applied quantum concepts exacerbates what is already a largely impressionistic argument. In Coale’s analysis, quantum concepts are not just irrelevant to the argument, they actively impede it. The result is further misrepresentation of these concepts and a set of observations that obscure rather than elucidate DeLillo’s novels.

158

Chapter 3

The example of Ettinger’s and Coale’s untidy use of quantum concepts reinforces the fact that, if literary critics are to maintain any credibility in their use of scientific paradigms, they must first ensure that they understand them; to misrepresent scientific concepts is to propagate misinformation rather than knowledge, and to render suspect their application to literature. It undermines efforts in the humanities to broaden their interdisciplinary reach, and opens the discipline up to the kind of parody that physics professor Alan Sokal snuck by the editors of Social Text in 1996— an act that quickly went viral within the academic community and beyond.95 The editors of Social Text later justified their decision to publish Sokal by professing their eagerness for work that combined science and the humanities/cultural studies. Because they did not, at that time, send out submissions for vetting by readers and who might have had the expertise to reject Sokal’s article as absurd, the reputation of this preeminent journal in cultural studies was deeply compromised. Setting aside the ethical question of whether it is right to misrepresent one’s work when submitting it for publication, Sokal seemed to have no trouble finding misinformed quotations that used quantum physics and science in general to support all manner of claims about postmodernism. The kind of casual, ill-informed approach to the sciences that led to the publication of his article in a leading academic journal proved to be a major public embarrassment for the humanities, and lent legitimacy to Sokal and Jean Bricmont’s 1998 follow-up screed, Fashionable Nonsense: Postmodern Intellectuals’ Abuse of Science.96 Contra Ettinger and Coale, Sean Kinch’s precise use of quantum entanglement to interpret characterization and narrative structure in Hopeful Monsters is both informed and justified.To begin with, unlike The Waves, Hopeful Monsters is obviously appropriate for two reasons: the characters openly use the concept of quantum entanglement to articulate and understand their relationship, and the author builds nonlocality into the narrative structure by virtue of the fact that the characters spend much of their time apart. Kinch takes pains to lay out the specific ways in which quantum entanglement plays out in the novel: In Hopeful Monsters, Mosley uses quantum entanglement to connect the two main characters on two levels: it is the metaphor Max and Eleanor themselves employ to describe their relationship, and it is the structuring principle Mosley imposes on their narratives. … After the emotionally charged meta-

Quantum Paradigms in Literary Criticism

159

phorical uses of the phase entanglement, Mosley leaves to Max the challenge of explaining the theory more thoroughly.The care and precision that Mosley takes in elucidating nonlocality alerts us to its broader poetic meaning. … [T] his passage gives readers a clear conceptual framework for understanding Max and Eleanor’s relationship and, more broadly, for imagining a new mechanism for human interconnectedness.97

Here, Kinch clearly sums up how quantum entanglement is woven into the fabric of Hopeful Monsters in three ways: as a central metaphor that the characters employ, as a model for how the reader can interpret the narrative structure, and in its extended interpretation by Max. Kinch outlines precisely how the characters understand quantum entanglement as a metaphor for how their love is structured: first, they express the belief that science provides powerful metaphors for depicting emotional life, and in fact has emotional content itself, “so it is natural for [Max] to see himself and Eleanor as entangled particles. … For Eleanor and Max, nonlocality supplies a means for them to sustain the plausible reality of their love over space and time.”98 The characters need to believe in quantum entanglement, because to do so allows them to maintain faith in their love even when they are far apart. More than this, observes Kinch, “entanglement steers their lives through parallel events and leads them to formulate similar ideas about science, politics, art, friendship and, as we have seen, love,” as well as similar thought patterns.99 Quantum entanglement thus does not merely provide a metaphor for the characters, it also describes the deeper organizing structure of their relationship. When Kinch talks about narrative structure, he focuses on how the concept of quantum entanglement in Hopeful Monsters enables the reader to appreciate the reciprocal coincidence of events in the characters’ lives, without the author having to articulate in each instance the content of the nonlocal effect that one has upon the other. Describing the first six chapters, where Max and Eleanor take turns recounting their early lives through their college years, Kinch argues that quantum entanglement gives readers a model for understanding how Mosley structures their alternating narratives. The basic premise of quantum nonseparability is that, for mathematically paired particles, measuring one particle also reveals attributes of the second, no matter how far apart the particles have traveled. In Hopeful Monsters, when we learn what Max or Eleanor is experiencing at any moment, we know that the other one is in an analogous

160

Chapter 3

situation. … Mosley constructs this section of the novel so that every major event in Eleanor’s narrative, from her perilous crossing of the Straits of Gibraltar to the chance meeting that leads her to Max, coincides with Max’s.100

Kinch offers a persuasive and detailed account of the parallels in the characters’ experiences, although, as throughout, his focus on parallels drops the key notion of causality in favor of a more general similarity based on analogous correspondence. Kinch also misses the fact that, in entanglement, a change made to the state of one particle produces the opposite effect in the other. Needless to say, this would preclude using entanglement to describe the characters’ parallel experiences in Hopeful Monsters. That Kinch is following the characters’ sense of how quantum entanglement works in their lives helps to redeem the reading, however, and in spite of this error, Kinch offers a cogent and multilayered analysis based on the metaphor. One might argue that a danger in using a very specific concept like quantum entanglement to describe the structuring of the narrative would lie in the tendency to see parallels in the characters’ actions and experiences where none exist, or to lose sight of the nuanced singularity of their separate experiences in the interest of drawing similarities. This loss of nuance, however, is the effect of using any interpretive tool—and Kinch himself acknowledges, “[w]hen we use quantum mechanics in criticism, we should carefully limit our conclusions to specific features of novels, and we should be wary of sweeping claims about texts being inherently ‘quantum.’ Quantum mechanics is one tool among many, and complex texts often require multiple approaches for full comprehension.”101 In fact, on the one hand, Kinch appears too cautious at times about the use of quantum concepts, suggesting that there must be specific quantum references in the actual text to justify a critic’s use of such concepts. On the other hand, given some of the abuses to which these concepts are put, ones that Kinch specifically outlines in the first half of his article, his caution is understandable. Still, it limits literary critics to the pursuit of extremely low-hanging fruit, where they can do little more than reiterate how the references are used in the novel. Moreover, as Kinch’s interpretation bears out, allowing the author’s representation of quantum concepts to guide one’s critique can lead to reiterating distortions of the concepts that are already present in the work.

Quantum Paradigms in Literary Criticism

161

HOLISM AND THE IMPLICATE ORDER: TEXTUAL TROPES AND ENFOLDMENT

The complex hierarchical order that can be analyzed within a novel is never truly fixed. In a great work of art, it is dynamical and always used in a creative fashion. —David Bohm

David Bohm believed that all aspects of the universe, including both matter and consciousness, are enfolded into one another and are expressed, in different ways, within an “implicate” and an “explicate” order.The explicate order refers to our normal perceptions of space and time, as well as our perception that particles exist separately from one another. The implicate order, in contrast, refers to a deeper and more fundamental reality. The total order of the material world is contained, or enfolded, within each region of space and time, just as aspects of the mind are enfolded in each region of the brain—in the same way that each region of a holographic photographic plate contains within it the entire threedimensional image, the totality of which can be viewed from an infinite range of perspectives. In Science, Order, and Creativity, Bohm and F. David Peat examine how the implicate order manifests itself iteratively in orders of varying detail and complexity in art. Bohm’s theory of art, and the novel in particular, is based on the principle of an initial, generative (implicate) idea that is then “enfolded” into (explicate) definite forms. However particular and detailed those forms may be, they will always have enfolded within them the overall generative idea.102 Bohm explains: Hierarchies are no longer fixed and rigid structures involving domination of lower levels by the higher. Rather, they develop out of an immanent generative principle, from the more general to the less general. The novel … is an example of such a hierarchy, for it grows out of the basic generative order within the author’s mind through the generative suborders of plot, character, atmosphere, means of expression, and so on. In addition, this generative order must be expressed within various conventional forms of syntax as they apply within the sentence, paragraph, and chapter, and to the tacit conventions of the novel. Therefore, while within a particular sentence the orders of syntax and semantics may appear to dominate, they are in fact serving the much larger generative order of the novel as a whole.103

162

Chapter 3

The important point here is the iterative nature of the relationship between the explicate orders and the implicate totality, a relationship that follows the same logic as the holograph: each specific expression, whether at the level of plot, character, or sentence, is an explicate expression of the same generative idea or principle. Bohm’s metaphor of the holograph is enlisted by critics to identify a type of narrative structure, to discuss the relationship between characters and their external environment, and to describe the reader’s direct intuition of the whole. The reader’s holographic perception may be understood as a function of the text’s formal qualities, or it may be associated with the nature of reading itself. Mary Ellen Pitts, for example, makes a general case for using Bohm’s hologram as a model for the reader’s interaction with the text. She observes, “The holographic model is especially useful in approaching literature and science today because it suggests the viewer’s active participation as he shifts positions to discover the different views and colors that may be obtained from different angles.”104 This is a reasonable, if somewhat vague, argument for reader participation, but she might have been better off using the metaphor of the kaleidoscope instead of the holograph, given the fact that changing one’s position relative to the holograph does not produce new impressions— rather, each position reflects the same whole. Pitts goes on to claim that “[a] major implication of the holographic paradigm for the individual reader is a greater sense of involvement with, as well as responsibility to, the whole. The reader participates in the whole creative process; reading itself, as reader-oriented theorists have insisted, is a creative process that undermines the notion of total autonomy and separateness of the text.”105 While Pitts pursues a valid point, Bohm’s holograph does not fit into her model of empowering reader/text interaction. It is true that, for Bohm, nothing is separate because the totality is always reflected in each part, and that perception—within the explicate order—is integral to Bohm’s model. The two “inseparabilities” (of Bohm and Pitts) are quite different, however; Bohm is talking about a state of unified being, while Pitt is talking about the transformative action of one element (the reader) upon another different element (the text). Pitts would have fared better had she relied on the reader response theory to which she alludes elsewhere. In doing so, she would have faced a greater challenge in

Quantum Paradigms in Literary Criticism

163

distinguishing herself from other critics who assign power to the reader, but she would at least have avoided distorting Bohm’s paradigm by trying to make it square with a mode of reading that it does not support. In his analysis of To the Lighthouse, Mark Hussey links Woolf ’s characters to enfoldment when he observes that “at several moments in the novel, boundaries between seemingly solid ‘selves’ and the world around them dissolve, giving rise to a sense that ‘everything is enfolded into everything else.’”106 As an example, he includes an extended quote from the book that describes Mrs. Ramsay’s reflection on how she is not distinct from objects upon which she dwells: Not as oneself did one find rest ever, in her experience … but as a wedge of darkness. Losing personality, one lost the fret, the hurry, the stir; and there rose to her lips always some exclamation of triumph over life when things came together in this peace, this rest, this eternity; and pausing there she looked out to meet the stroke of the Lighthouse, the long steady stroke, the last of the three, which was her stroke, for watching them in this mood always at this hour, one could not help attaching oneself to one thing especially of the things one saw; and this thing, the long steady stroke, was her stroke. Often she found herself sitting and looking, sitting and looking, with her work in her hands until she became the thing she looked at—that light, for example.107

Following the quotation, Hussey remarks: It is usually difficult not to quote Woolf at length—her modes of composition might well be termed holographic, so that clusters of imagery are “enfolded” throughout a work (The Waves perhaps best exemplifies this). In the passage above, the division between observer and observed is dissolved in the act of perception; the “self ” of Mrs. Ramsay might be called, in Bohm’s term, a “relatively autonomous sub-totality” of a larger flowing movement.108

In his interpretation of the passage, Hussey connects To the Lighthouse with enfoldment on two levels: its narrative structure, and the nature of Mrs. Ramsay. In terms of narrative structure—his reference to the clusters of imagery in the novel—Hussey appears to use the term “enfolded” to mean something like “integrated”—so that distinct, self-coherent, and bounded metaphorical conceits appear throughout the text. Hussey does not quite capture the nature of enfoldment, however: whereas Bohm posits each region as an explicate, iterative unfolding of the implicate totality, Hussey is just suggesting the repetitive use of imagery in the novel.

164

Chapter 3

In the second half of the preceding quotation, Hussey associates the observer/participant dynamic with the fact that Mrs. Ramsay becomes the thing she is looking at, and then extrapolates from this to claim that her dissolution into the material object reflects how Mrs. Ramsay is part of a larger narrative flow. In the end, however, Bohm is not necessary either to support Hussey’s claim about the structure of Woolf ’s imagery, or to support his claim about Mrs. Ramsay. In fact, using Bohm is misleading: enfoldment is not the dissolution of an element into the totality—it is the reflection of the totality within that element. Summed up, Hussey is merely saying that Woolf integrates extended metaphorical conceits into her book, and that Mrs. Ramsay has poor boundaries. The latter observation reinforces a stereotypical “female” attribute that critics frequently associate with Mrs. Ramsay’s poor ego boundaries—albeit one that is often viewed as a positive openness, empathy, or fluidity. When Hussey turns his gaze toward the reader’s experience, he concludes, “if memory functions according to a ‘holographic’ paradigm, and Woolf ’s fictions are analogous to holographs, the reader can bring to consciousness the virtual ‘wholeness’ of the work from any point in the text, reading being analogous to shining a laser on a holographic plate.”109 Hussey is right to cite Bohm’s argument that consciousness and memory function in a holographic manner, but he misses Bohm’s point that enfoldment in the mind and enfoldment in matter represent analogous poles, not common parts of a single totality. More significantly, since he has not at any point in his argument actually demonstrated that Woolf ’s narrative is holographic, Hussey’s claims about the reader’s “mystical” and “direct intuition of the implicate order” of the total text falls flat.110 There is some potential in arguing that a single text, or even all texts, are holographic in the sense that each part reflects the whole. This claim would have to be supported by first establishing what the text in its entirely expresses—what its generative principle is—and then through increasingly close readings to demonstrate how that principle is expressed on the level of aesthetics, theme, plot, character, image, and syntax. The goal is thus not to establish how the reader’s shifting perspective introduces variety or generates the text (Pitts), or to show how the parts dissolve into the whole (Hussey). Rather, the goal is to establish the continuity or sameness that underlies the distinct parts of the text, which, according to Bohm, are relatively autonomous subtotalities that

Quantum Paradigms in Literary Criticism

165

nevertheless reflect the whole.While Bohm is attempting to solve a quantum problem, and while the quantum is frequently associated with fragmentation, indeterminacy or excess of meaning, and lack of closure, Bohm’s model can only support an interpretation based on coherence, predictability, highly determinate order, and tight closure. Bohm’s model, in its holistic approach, is distinctly romantic (and thus doesn’t support Hussey’s association of Bohm’s quantum discourse with postmodernism), and it is not amenable to the paradoxes of modernist literature.111 Bohm’s model is not appropriate when the goal is to expose postmodern discontinuities, fragmentations, and open-endedness; instead, it serves the search for continuities, wholeness, and the essential “genius” of the work. CONCLUSION

There exist some basic prerequisites for a persuasive use of quantum phenomena to discuss elements of literature such as theme, narrative form, characterization, and the role of the narrator/author/reader. Kinch, in the first half of his article on Hopeful Monsters, provides a very cogent summary of these conditions. The first and most basic requirement is a thorough understanding of the quantum concept that you are using. As Kinch observes,“[f]or critics productively to use science as reading models, their understanding of the scientific concepts must come not from the literary text, but from reading scientific sources.”112 In other words, the scientific model, properly understood, must lead the analysis, rather than be derived from the literary text. Ironically, Kinch is one of the offenders here: he relies on Muldoon’s somewhat distorted version, expressed via his characters, of entanglement’s central dynamics. However, Kinch does much better than those who merely identify an uncertainty or duality in the work and then invoke, by virtue of its name, a quantum concept. Often there is no reason to reference the quantum concept at all, and the critical use of them appears to reflect a desire to dress up references to indeterminacy or oppositions present in the text.The critic may simultaneously promulgate misleading and inaccurate representations of quantum science, and offer confused interpretations of the text. This undermines attempts to bring the sciences and the humanities together and to advance the field of science studies—and fails to take up the responsibility of sharing information that is rigorous, thoughtful, and genuinely productive.

166

Chapter 3

A second prerequisite is an understanding that designations such as the Uncertainty Principle and the Principle of Complementarity are already (imperfect) metaphors, and to be sensitive to the problems of language that inhere in their use to describe quantum phenomena. Again, Kinch nicely sums up the challenge: “A further problem for critics who want to use quantum mechanics to read literary texts is that the science itself is difficult, if not impossible, to describe in common language. … We must acknowledge that any verbal account of quantum mechanics … is essentially metaphoric, making our use of it in criticism doubly metaphoric.”113 Critics such as Froula and Kimmelman recognize the fundamental limitations of language with respect to quantum physics, and use it to good effect in their analysis.When critics use a quantum concept as a model for a textual feature, they must recognize that they are working on the level of metaphor, and must avoid the temptation—into which Edlund falls, for example—to apply the language of quantum physics literally in order to support ontological claims about the nature of reality and human experience. The most successful users of quantum concepts are those who are scrupulous in describing the concept, and who then apply it to the text in a precise and delimited manner.When critics elide important details, such as the location and determining effect of the measuring apparatus in the Uncertainty Principle; or associate two disparate phenomena, such as entanglement and the observer effect; or apply the wrong dynamic to the wrong phenomenon, such as indeterminacy to wave/particle duality, they build in inaccuracies that threaten to make incoherent their interpretations of the text. When they use generalized references to “the quantum world” in order to make overarching claims about, for example, postmodern literature or the role of the reader in relation to all literary texts, they both dilute the usefulness of quantum physics as an interpretive tool, and abandon any nuanced sense of the distinctiveness of individual texts and the varied manner in which readers interact with them. Delimiting the use of quantum concepts requires an accurate assessment of which texts, and which aspects of a given text, are appropriately viewed through this lens. It also requires the recognition that a “quantum” interpretation does not exhaust the potential meaning of the novel. No analytic paradigm is exhaustive, and for the most part, critics do limit themselves to the analysis of a particular aspect of a literary work. It is

Quantum Paradigms in Literary Criticism

167

possible to take this restriction too far, however, as Kinch does in limiting the application of quantum physics to “physics texts” in which the authors or characters specifically refer to physics in general, and quantum physics in particular.114 In fact, Kinch argues that not even every novel that refers to quantum physics merits quantum physics as an interpretive tool— those that follow conventional narrative conventions, for example, should be excluded.115 This seems overly selective; clearly, there are novels and poems that, while not including any reference to quantum physics or to science at all, benefit from the careful use of quantum concepts in a way that expands our understanding of the text. The use of quantum physics as a means to interpret texts thus brings with it many potential pitfalls.To avoid these pitfalls, critics must demonstrate due diligence in researching and understanding the science, and exercise care in applying quantum metaphors to literary works. They must assess, with an open mind, whether quantum concepts are really necessary to their interpretation, which also means understanding the literary text with the same thoroughness that they understand the scientific concepts. They must focus on specific dynamics within the text, select the appropriate quantum concept in all its specificity, and avoid sweeping claims about “quantum” critical methodologies or entire categories of literature. If these basic conditions are met, and if the critic possesses the interpretive skill to integrate literary analysis with quantum physics in a dynamic and productive manner, the use of quantum concepts can both open up new perspectives on a given literary text and expand our understanding and appreciation of quantum theory and its epistemological and linguistic implications.

4

NEW AND POST-NEW AGE APPROPRIATIONS

INTRODUCTION

Concepts that at first express axiomatic principles in their original field can, over time, accrue and dispense meanings that expand to become touchstones for wider-ranging sensibilities. Quantum physics, repollinated by the popular imagination fifty years after its inception, has been enlisted to lend credence to a host of unconventional practices and philosophies—from quantum touch healing to the notion of quantum jumping into alternative versions of oneself.That the quantum has proved so amenable to this kind of co-optation seems both unlikely, given its inaccessibility, and inevitable, given the originary distortions produced by its translation into language. In fact, it is the very combination of inaccessibility and distortion that has made the quantum world so pliable to such a distant range of interests and agendas and that has carried it so far from its origins. I begin this chapter by introducing the notion of “quantum consciousness,” and by tracing how the initial language used to describe quantum phenomena has been used to position the self in mutually affecting relationships to other subjects and to the cosmos. I do so by analyzing the rhetorical strategies used in New Age texts that advance the concepts of quantum consciousness, quantum mysticism, and quantum healing. I begin with a summation of the concept of quantum consciousness and its distance from the quantum interpretation’s original conception of the relationship between observer and observed. I examine two foundational New Age books written in the 1970s: Fritjof Capra’s 1975 The Tao of Physics and Gary Zukav’s 1979 The Dancing Wu Li Masters, showing how

170

Chapter 4

these two books midwifed the complementary relationship between physics and Eastern philosophy. I then turn to the field of politics, considering the burgeoning field of “quantum politics,” which typically defines itself against the liberal individualism of Newton’s era, and invokes quantum concepts in an effort to generate new language that better models the global complexities of our time. Advocates of quantum politics are not scientists; rather, they are political philosophers or political scientists who look to quantum concepts to provide metaphors for new ways of resolving social fragmentation, alienation, and cultural conflict. I consider the following: how successful are these models at offering a new form of political engagement? What conceptual consistencies link the various attempts to import quantum metaphors into the arena of politics? To what extent are they able to envision concrete applications of quantum politics to the specific socio-political challenges that we face? I conclude this chapter by exploring how quantum phenomena have been massaged by post-New Age practitioners to underwrite the promise of spiritual and financial gain. I address the following questions: if traditional self-help and personal growth literature is typically packaged and sold to consumers in the language of common sense, why would those interested in offering accessible and engaging self-help models choose this most conceptually inaccessible of sciences in order to draw in their clientele? How do post-New Age practitioners of quantum healing and vendors of quantum get-rich sites reconfigure quantum particles as vivified agents whose unique movements and interactions promise to secure commodified forms of health, happiness, and wealth? What individual and societal priorities inspire popular constructions of the quantum, and what does the post-New Age use of quantum language tell us about dominant forms of subjectivity in the late twentieth and early twentyfirst century? NEW AGE QUANTUM MYSTICISM

If Bohm’s physics, or one similar to it, should become the main thrust of physics in the future, the dances of East and West could blend in exquisite harmony. Physics curricula of the twenty-first century could include classes in meditation. —Gary Zukav

New and Post-New Age Appropriations

171

Partly because of the metaphysical and ontological questions that quantum physics introduces, several of the original quantum physicists, including Schrödinger and Wolfgang Pauli, expressed an interest in aspects of mysticism. David Bohm’s unconventional theory of the implicate and explicate order, developed in the 1980s, is the best known and most cited work by those seeking to forge links between consciousness and the material world. Bohm argued that the implicate order enfolded both consciousness and matter, although he saw them as two different poles that expressed the same relationship between the implicate and explicate order. Bohm’s proposal that there exists an implicate but invisible cosmic oneness has offered fertile ground for those looking to quantum theory for confirmation of a deeper reality beyond our immediate grasp. The parallels between matter and consciousness that Bohm posited have proven particularly appealing to New Age followers looking for evidence of an inclusive and animate universe. New Age quantum rhetoric was inaugurated with the publication of Fritjof Capra’s 1975 The Tao of Physics and Gary Zukav’s 1979 The Dancing Wu Li Masters. Capra is typically cited as the first significant New Age popularizer who wedded quantum physics and Eastern mysticism, an association that has endured into the twenty-first century. In his widely read book, Capra argues that the basic oneness of the world advocated by Eastern mysticism is similarly revealed in the modern physics of the early twentieth century. Capra promises readers that as we come to understand the concepts of the new physics, we shall see that these concepts express again and again the same insight: the constituents of matter are all interconnected, interrelated, and interdependent, and cannot be understood as isolated entities.1 In his interpretation of the observer effect, Capra claims that subatomic phenomena and the measuring device are part of a single system that culminates in the consciousness of the observer/scientist.2 Capra’s emphasis on the interconnectedness of object, measurement, and the consciousness of the observer in the quantum world leads him to conclude: “Quantum theory forces us to see the universe, not as a collection of physical objects, but rather as a complicated web of the various parts of a unified whole.”3 Having established the foundation of quantum physics as interconnectedness, Capra compares what he calls this “web philosophy” of quantum physics to Eastern mysticism, claiming that Eastern

172

Chapter 4

mystics “express their experience in words that are almost identical with those used by atomic physicists.”4 As proof of this similarity, he offers the following quotation from the Hindu philosopher Sri Aurobindo: “The material object becomes … something different from what we now see, not a separate object on the background or in the environment of the rest of nature but an indivisible part and even in a subtle way an expression of the unity of all that we see.”5 Capra executes here one of his recurring strategies: he creates a “quantum context,” and then uses his own language, and not that of physicists, to establish the link between quantum physics and Eastern mysticism. Capra’s claim also extends well beyond the original articulation of the observer effect, which was intended to refer only to the impact that the measuring/detecting instrument has on a system, quite independent from our witnessing of the result. In fact, Heisenberg took pains to reinforce the fact that the impact of observation in quantum phenomena should not be misunderstood to imply that subjective features ought to be brought into the description of nature: Therefore, the transition from the “possible” to the “actual” takes place during the act of observation. If we want to describe what happens in an atomic event, we have to realize that the word “happens” can apply only to the observation, not to the state of affairs between two observations. It applies to the physical, not the psychical act of observation, and we may say that the transition from the “possible” to the “actual” takes place as soon as the interaction of the object with a measuring device, and thereby with the rest of the world, has come into play; it is not connected with the active registration of the result by the mind of the observer.6

Victor Stenger refers this confusion between instrument and consciousness back to the original ambiguity, and he is worth quoting at length: Ironically, the seemingly profound connection between quantum and mind is an artifact, the consequence of unfortunate language used by Bohr, Heisenberg and the others who originally formulated quantum mechanics. In describing the necessary interaction between the observer and what is being observed, and how the state of a system is determined by the act of its measurement, they inadvertently left the impression that human consciousness entered the picture to cause that state to come into being.This led many who did not understand the physics, but liked the sound of the words used to describe it, to infer a fundamental human role. … If Bohr and Heisenberg had

New and Post-New Age Appropriations

173

spoken of measurements made by inanimate instruments rather than “observers,” perhaps this strained relationship between quantum and mind would not have been drawn. For, nothing in quantum mechanics requires human involvement.7

While Bohr and Heisenberg did often use the term “measuring” or “detecting” device, they also often simply used the word “observer,” and it then became canonized as the “observer effect.” Stenger’s clarification would not deter Capra, however, who counters that quantum physics has simply not gone as far as Eastern mysticism in abolishing the distinction between observer and observed, subject and object, and that he anticipates future developments in quantum physics will make it necessary to include human consciousness in its description of the world.8 This anticipatory strategy is also typical of Capra: he refers to the quantum concept, revises it to support his argument, and then justifies the revision by anticipating an inevitable future advance in quantum physics that will bring it more in line with the tenets of Eastern mysticism. Capra performs similar moves in relation to wave/particle duality, quantum field theory, matrix theory, and the Uncertainty Principle. He combines, for example, the Uncertainty Principle, which states that we can never simultaneously know two conjugate states of a system, with the related concept of probability waves—the probability of where we might detect those aspects of the system that we cannot measure precisely. Capra begins with an accurate representation of probability: “the waves, associated with particles … are not ‘real’ three dimensional waves, … but are ‘probability waves’; abstract mathematical quantities which are related to the probabilities of finding the particle in various places and with various properties.”9 Capra then jumps to the unsupported conclusion that the probability function introduces a fundamental opposition between existence and nonexistence: “We can never say that an atomic particle exists at a certain place, nor can we say that it does not exist.”10 Here, he turns a problem of detection into an existential fact by excising two critical words from his initial description of probability: the word “finding” (which maintains probability at the levels of knowledge), and the word “properties” (which reminds us that it is not just location that we are talking about, but rather the total state of the particle/system). He does retain the phrase “what we can say,” which acknowledges that we are still concerned with our knowledge of the system, but effaces the conjugate

174

Chapter 4

relationship between momentum and position by referring only to position, or “place.” The second half of the sentence, “nor can we say that it does not exist,” dispenses with the word “place” altogether and focuses on existence itself. Capra ends with the observation that quantum physics “transcends the concept of existence and non-existence,” and that the particle “manifests a strange kind of physical reality between existence and non-existence.”11 This last phrase alludes to Heisenberg’s assertion that the probability wave represents “a tendency for something. … It introduced something standing between the idea of an event and the actual event, a strange kind of physical reality just in the middle between possibility and reality [emphasis added].”12 The phrase “idea of an event” locates Heisenberg’s observation squarely in the realm of concepts and processes, but not existence. Even the more existential phrase “between possibility and reality” refers not to existence versus nonexistence, but rather to the potential represented in our lack of complete knowledge. As I said in chapter 1 regarding this same quotation from Heisenberg, he is referring not to being, but to the tendency toward being—a tendency that is not located in the ontology of the particle, but rather in the limits of measurement. Once Capra has translated the Uncertainty Principle into terms more befitting his goal, he juxtaposes a quotation from Oppenheimer with one from an Eastern mystic. Capra first quotes Oppenheimer on the Uncertainty Principle: “If we ask, for instance, whether the position of the electron remains the same, we must say ‘no’; if we ask whether the electron position changes with time, we must say ‘no’; if we ask whether the electron is at rest, we must say ‘no’; if we ask whether it is in motion, we must say ‘no.’”13 Curiously, Capra leaves out Oppenheimer’s next sentence, which specifically associates physics with Buddhism: “The Buddha has given such answers when interrogated as to the conditions of a man’s self after his death.”14 The full quotation indicates that Oppenheimer is writing about the limits of knowledge, but by leaving the last sentence out, Capra can interpret Oppenheimer’s words as expressing what Capra views as quantum physics’ most fundamental insight: the transcendence of the duality between existence and nonexistence. Capra follows Oppenheimer with a quote from the Ashvaghosha:“Suchness is neither that which is existence, nor that which is non-existence, nor that which is at once existence and

New and Post-New Age Appropriations

175

non-existence, nor that which is not at once existence and nonexistence.”15 From the initial slippage wherein “probability” becomes “non-existence,” and the “idea of an event” becomes “existence,” and juxtaposition of the two quotations, Capra is able to attribute existential import to matters that, in the quantum interpretation, concern idealizations of description and observation. Sal Restivo aptly describes Capra’s basic strategy as the “parallelist method of juxtaposed quotations.”16 This method, notes Restivo, rests on the basic assumption that “if the rhetorical, imagery, and metaphoric content of statements on physics and mysticism is similar, then the conceptual context must be similar.”17 While Restivo is correct, Capra does more than “discover” metaphoric content—he actively transforms the original quantum language through a combination of metonymic slippage and misrepresentation in order to force an alignment between the discoveries of quantum physics and the insights of Eastern mysticism. Like Capra’s, Zukav’s descriptions of quantum concepts are extensive and well informed; like Capra, he has a tendency to abandon his rigor whenever he wants to establish a connection between quantum physics, human experience, and Eastern mysticism. In constructing these connections, Zukav engages in a number of rhetorical strategies that include distorting or reinventing the language of quantum physics, applying faulty logic characterized by unwarranted conceptual leaps, and collapsing the distinction between microscopic and macroscopic behavior. Most of Zukav’s book is about physics and human experience, with a few sentences here and there commenting on the similarities between quantum and Eastern language. Only toward the end does he include more extended comparisons to Eastern philosophy. Nevertheless, his recontextualization of quantum concepts sets the stage for a worldview that aligns with Eastern mysticism. Zukav’s title derives from the combination of the Chinese word “Wu,” which he translates as “matter/energy,” and “Li,” which he translates as “universal order/organic pattern.” Thus, Wu Li can be interpreted as “organic patterns of matter.” Zukav describes a process wherein having discovered that the term “Wu Li” is also the Chinese [Mandarin] word for “physics,” he found that “[h]ere, at last, was the vehicle through which we could present the seminal elements of advanced physics.”18 Having introduced his own definition of the physical world as patterns of organic

176

Chapter 4

energy, Zukav proceeds as if this description were canonical within quantum physics, with the observation that “we talk of physics as patterns of organic energy.”19 While there is no evidence that anyone other than Zukav uses this phrase in relation to quantum physics, his introduction of the term “organic” (which he then attaches to the term “alive”) in relation to quantum physics prepares the way for a series of associations that support a conclusion Zukav returns to again and again throughout the book—that there is a living correspondence between humans and subatomic particles. Zukav introduces the idea of this correspondence with the following passage: “We would like to think that we are different from stones because we are living and they are not, but there is no way we can prove our position. … We cannot establish clearly that we are different from inorganic substances. That means that, logically, we must admit that we may not be alive. Since this is absurd, the only alternative is to admit that ‘inanimate’ objects may be living.20 It seems that there would be plenty of ways to establish our difference from inorganic substances, but once having concluded that we cannot, Zukav elaborates on it in his discussion of the double-slit experiment. Zukav proclaims, “the astounding discovery awaiting newcomers to physics is that the evidence gathered in the development of quantum mechanics indicates that subatomic ‘particles’ constantly appear to be making decisions!”21 Certainly, Zukav’s allusion here and elsewhere to “knowing” particles is another example of the unfortunate original use of an everyday term to describe the behavior of subatomic particles. Zukav takes the term literally, and claims that the double-slit experiment proves that subatomic particles “process information and act accordingly.” He concludes, “strange as it may sound, [photons] seem to be organic. Since we are also organic, there is a possibility that by studying photons (and other organic quanta) we may learn something about us.”22 The chain of association goes something like this: photons process information and act on it in the double-slit experiment; something that processes information and acts on it is organic; photons are organic; we are organic; in studying photons we are studying ourselves. In this roundabout way, Zukav establishes not only that subatomic particles are “conscious,” but also that we are very much like they are. Zukav uses the same strategy of transitioning seamlessly from the microscopic to the macroscopic—particularly where humans are

New and Post-New Age Appropriations

177

concerned—in his interpretation of the observer effect. He begins by claiming that the nature of light (as particle or wave) is constituted by our interaction with it, and then arrives at the following conclusion: “It appears that light has no properties independent of us! To say that something has no properties is the same as saying that it does not exist. The next step in this logic is inescapable. Without us, light does not exist. … This remarkable conclusion is only half the story.The other half is that, in a similar manner, without light, or, by implication, anything else to interact with, we do not exist! [emphasis in original].”23 Zukav’s first assumption is that we are creating the nature of light, rather than merely exposing one aspect of its behavior. He then suggests that our interaction with the photon occurs not just at the level of consciousness, but at the level of our very “self.” Needless to say, this line of thought ignores the fact that it is really the detecting device that influences the behavior of the photon—behavior that we can only observe after the fact. Zukav then proceeds to make his point through a series of logical fallacies. Imagine that in the following string of associations, “H” stands for “human” and “P” stands for “photon.” Zukav’s logic goes something like this: if H’s action upon P produces a particular outcome, then P’s action upon H must produce the same outcome. Specifically, if H must interact with P to bring it into existence, then P must interact with H to bring it into existence. If the absence of H means that P does not exist, then the absence of P means that H does not exist. To put this line of reasoning in the language of mathematics, Zukav’s logic assumes that H and P are commutative operands (that H * P = P * H) and that they correspond in every significant way.This ignores the fact that quantum effects do not pertain on the macrocosmic level, and that the causal effect of our observation is unidirectional. Zukav’s reasoning here supports his fundamental claim (which echoes and extends that of Capra’s): the universe is one organic web of interchangeable and interaffecting parts wherein distinctions between the macroscopic and microscopic do not pertain. Zukav’s analogies between physics and Eastern practices and philosophies tend to contain the same leaps of logic that characterize his collapse of the distinction between animate and inanimate, microcosmic and macrocosmic. In reference to the use of electron colliders to discover new particles, Zukav writes:

178

Chapter 4

We can say as well that an electron moving through space emitted a photon and went out of existence at that point! A new electron was created in this process and it departed the scene with a new momentum. There is no way of knowing if this interpretation is correct or not because all electrons are identical. However, it is simpler and more consistent to assume that the original particle was annihilated and a new particle was created. The indistinguishability of subatomic particles makes this possible [emphasis in original].24

By Zukav’s own admission, his assertion that particles are actually being created and destroyed (rather than revealed, however briefly) is on shaky ground, and is dependent again upon the assumption that we cannot prove otherwise. Despite the speculative nature of Zukav’s claim, he proceeds as if it were a proven fact, and arrives at the mystical-sounding conclusion that “subatomic particles are this unceasing dance of annihilation and creation [emphasis in original].” Zukav then claims that: [t]his twentieth-century discovery, with all its psychedelic implications … is very similar to the way much of the population, including the Hindus and the Buddhists, view their reality. Hindu mythology is virtually a large-scale projection into the psychological realm of microscopic scientific discoveries. Hindu deities such as Shiva and Vishnu continually dance the creation and destruction of universes while the Buddhist image of the wheel of life symbolizes the unending process of birth, death, and rebirth.25

Once again, Zukav begins with a questionable premise about subatomic behavior, and then applies it to the macroscopic realm. His observation that “Hindu mythology is virtually a large-scale projection into the psychological realm of microscopic scientific discoveries” is difficult to parse. What does “the psychological realm of microscopic scientific discoveries” mean? What does it mean that Hindu mythology “projects” into it? The best answer seems to be that Zukav views Hindu mythology as based on getting in touch with subatomic phenomena, and living as they “live.” Later, Zukav forges a conceptual parallel between the terminology of physics and Eastern practices. Zukav fixes on the term “massless particle,” which is used to describe an element in a mathematical structure and which, Zukav points out, appears paradoxical when translated into language.26 Zukav then defines this as a type of Zen Buddhist koan, which is an apparently nonsensical question upon which the Zen practitioner meditates with the goal of transcending the limits of thought (“what is

New and Post-New Age Appropriations

179

the sound of one hand clapping?” or “what did you look like before your parents were born?” are two familiar examples). While Zukav defines koans as paradoxes, this is not quite accurate—they are more along the line of questions that cannot be “thought through.” Nevertheless, Zukav is correct in claiming that “paradoxes are common in Buddhist literature,” and he is not entirely out of line in claiming that the directive to “picture a massless particle” is consistent with a Zen approach. Zukav’s final conclusion is a stretch, however. He asks the reader, “Is it a coincidence that Buddhists exploring ‘internal’ reality a millennium ago and physicists exploring ‘external’ reality a millennium later both discovered that ‘understanding’ involves passing the barrier of paradox?”27 It does in fact appear to be coincidental, and physics is certainly not the only Western practice in which paradox dominates; one need only to examine modernist and postmodernist literature and thought to see the same sorts of paradoxical structures. To compare the use of the term “massless particle” to the practice of meditating on nonsensical questions is also to impose an entire belief system regarding the nature of “reality” upon what is merely a matter of the limitations of language in describing mathematical constructs. Capra and Zukav, whose books were written in the 1970s, clearly align with the Western New Age movement in subscribing to a philosophy wherein the universe is characterized by holism, universal consciousness, and an altered relation to reality. Both Capra, who has a PhD in theoretical physics, and Zukav, who is best known for his teaching on personal spiritual growth and what he calls “authentic power,” offer detailed, informed, and sophisticated explications of developments in physics during the twentieth century. For someone who is not a physicist, Zukav demonstrates a very impressive command of the concepts. In fact, part of what makes Zukav’s conclusions appear so contrived is his initial thorough description of the concepts. Had Zukav begun with sweeping generalizations about quantum phenomena, and had he not been so committed to a “scientific method” in reaching his conclusions, his logical fallacies would not have been so apparent. In the case of both authors, their knowledge of physics exceeds their familiarity with Eastern mysticism. In Capra’s case, associations to Eastern practices in his book often appear as throwaway lines or afterthoughts buried within an exposition of physics or tacked onto the end of chapters. Capra acknowledges that he

180

Chapter 4

often sums up Eastern practices only in passing, but then justifies it by asserting that mysticism simply cannot be taught in a book. Zukav begins by using the phrase “Wu Li” as a metaphor to organize his thoughts about physics, and would have fared better had he maintained its status as a metaphor, and resisted the temptation to make ontological claims about the identity between quantum phenomena, human existence, and Eastern mysticism. Both writers end up distorting the quantum concepts that they so deftly explain at the outset, compromising to some degree what are otherwise cogent, accessible explanations of the development of quantum physics and the concepts that define it. While they don’t always apply quantum concepts to Eastern mysticism in a manner that does justice to both, Zukav and Capra remain important for how central their works were for renewing the interest in theoretical physics and in quantum physics in particular that emerged in the 1970s, when Zukav’s and Capra’s works were published. Both Zukav and Capra were also involved with the New Age collective the Fundamental Fysiks Group. The Fundamental Fysiks Group was founded by Elizabeth Rauscher and George Weissmann in May of 1975 as an informal discussion forum, and attracted a number of academics and promising young physicists. It was heavily influenced by the New Age counterculture, and its members, alongside tackling the fundamentals of quantum physics for the first time in decades, dabbled in psychedelic drugs, transcendental meditation, consciousness expansion, psychic mind reading, telekinesis, and séances. While the group’s cultural milieu influenced their more radical pursuits, they were also influenced by prominent academics in the fields of physics and mathematics, most notably Princeton Professor and Nobel Laureate Eugene Wigner and Princeton Professor John Wheeler. Wigner is best known for his mathematical formulation of quantum mechanics and his research into the structure of the nucleus, and his most significant publication is the now-classic 1960 article on the philosophy of mathematics and physics, titled “The Unreasonable Effectiveness of Mathematics in the Natural Sciences.”28 Wigner’s interests extended well beyond mathematical formalism, however, and it is his thinking on the relationship between quantum effects and human consciousness that inspired members of the Fundamental Fysiks Group. Wigner proposed that consciousness plays a central role in quantum mechanics, and in 1961

New and Post-New Age Appropriations

181

published a widely reprinted article entitled “Remarks on the MindBody Problem,” in which he stresses the basic role played by consciousness in quantum theory.29 In this article, Wigner offers a simple thought experiment, often referred to as “Wigner’s friend,” which he claims proves that in quantum mechanics the conscious subject has a separate but tangible role from the inanimate device—specifically, that the wavefunction collapses due to its interaction with consciousness.Wigner uses the example of a photon, considering the question of what initiates the wavefunction to collapse and to cause a flash on a screen. In the thought experiment, the “friend” is inside the lab, and either sees a photon flash or does not. The other person is outside the lab and does not know the outcome. Wigner argues that without human consciousness, this leaves the world in a superposition of states, an unsustainable proposition. He resolves the paradox by saying that it is the friend’s consciousness that collapsed the wavefunction.30 In his 1970 collection of essays, Symmetries and Reflections: Scientific Essays, Wigner states that “it was not possible to formulate the laws of quantum mechanics in a fully consistent way without reference to consciousness.”31 John Wheeler was central in reviving interest in general relativity in the United States after World War II, in explaining the basic principles behind nuclear fission, and in linking the term “black hole” to objects having a gravitational collapse—but his influence on the Fundamental Fysiks Group concerned the role of consciousness in quantum effects. Wheeler argued that observers participate in creating the reality they measure, and proposed that a physicist’s decision to measure a particle’s position rather than its momentum changes the objective properties of the real world.32 At a 1974 conference at Oxford, Wheeler stated that we can put in a device to measure position or we can insert a device to measure momentum. But the installation of the one prevents the insertion of the other.We ourselves have to decide which it is that we will do.Whichever it is, it has an unpredictable effect on the future of that electron.To that degree the future of the universe is changed.We changed it.We have to cross out that old word “observer” and replace it by the new word “participator.” In some strange sense the quantum principle tells us that we are dealing with a participatory universe.33

182

Chapter 4

By “backing up” the moment of action from the instrument to the observer’s conscious choice concerning how to insert the instrument, Wigner launches a direct challenge to Heisenberg’s insistence that observation concerned only the measuring instrument, and that subjectivity plays no role in the process. Later in his life,Wheeler speculated that reality itself is created by observers in the universe. In a 2006 radio interview titled “The Anthropic Universe,” Wheeler said, “we are participators in bringing into being not only the near and here but the far away and long ago.We are in this sense, participators in bringing about something of the universe in the distant past and if we have one explanation for what’s happening in the distant past why should we need more?”34 Like Zukav and Capra, Wigner’s “Participatory Anthropic Principle,” as Wigner coined it, expanded out from a quantum-experimental context to a context of cosmic proportions. A leading voice within the Fundamental Fysiks Group was American theoretical physicist Jack Sarfatti, who combined the ideas of Wigner and Wheeler. Sarfatti proposed that if every quantum object is interconnected with every other via quantum entanglement, and if consciousness plays a central role in quantum mechanics, then modern physics might provide a natural explanation for parapsychological or psychic faculties and phenomena.35 In 1975 Sarfatti was involved with Werner Erhard in setting up the nonprofit think tank, the Physics–Consciousness Research Group. In January of 1976 Sarfatti and the Physics–Consciousness Research Group gathered for a month-long conference on physics and consciousness, and it is said that Zukav’s The Dancing Wu Li Masters emerged out of his attendance at this conference.36 Working largely outside academia, Sarfatti focused on and continues to specialize in the study of quantum physics and consciousness, arguing that mind is crucial to the structure of matter. Another member of the Fundamental Fysiks Group was American mathematical physicist Henry Stapp who, like Wigner and Wheeler, favored the idea that quantum wavefunctions collapse only when they interact with consciousness—only when conscious minds select one among many alternative quantum possibilities. Stapp was known in particular for his proofs of strong nonlocality properties, and he published many papers pertaining to the nonlocal aspects of quantum mechanics. Stapp’s interest in nonlocality was shared by many members of the group,

New and Post-New Age Appropriations

183

and found its source in Bell’s theorem, formulated by J. S. Bell in 1964.37 Bell’s theorem states that the predictions of quantum theory cannot be accounted for by any local theory, and his theorem may be seen as an extension of the Einstein-Podolsky-Rosen paper from which Schrödinger derived the notion of entanglement. It was the concepts of nonlocality and entanglement that so captured the imagination of many members of the Fundamental Fysiks Group. In his 2011 How the Hippies Saved Physics, David Kaiser undertakes a book-length investigation of the Fundamental Fysiks Group. As Kaiser tells it, the group was responsible for a striking and essential renaissance of the ideas that had occupied Bohr, Heisenberg, and Schrödinger, as well as later refiners of quantum concepts such as David Bell and David Bohm. According to Kaiser, “[t]heir concern … was to broaden the physicists’ range of approaches or methods beyond the hyperpragmatism that had marked the earlier Cold War years.They strove to expand the physics profession’s collective mental space, to push beyond what they considered a narrowness of vision that hardened after a quarter century of instrumentalist thinking.”38 Kaiser describes a physics community that was foundering until the Fundamental Fysiks Group emerged as the “full-color public face of the ‘new physics’ avant-garde.” He makes large claims for the group’s influence, asserting that while the group existed on the margins of mainstream physics, its members nevertheless “managed to parlay their interest [in quantum physics] into a widespread cultural phenomenon.”39 Kaiser argues that the Fundamental Fysiks Group saved physics in three ways, the first of which concerned the manner in which they thought and interacted.The group was more speculative, less constrained, and more engaged than postwar physicists had been up to that point, and they were concerned with the kind of big-picture ideas that had occupied the minds of Bohr, Heisenberg, and Schrödinger. Second, argues Kaiser, the group rescued entanglement and Bell’s nonlocality theorem from obscurity.40 By Kaiser’s account, in the years since the Fundamental Fysiks Group was active, topics such as Bell’s theorem and quantum entanglement have moved to the center of legitimate physics.41 Finally, they used Bell’s theorem to develop the “no-cloning” theorem that is behind the science of quantum information.42 The significance of the Fundamental Fysiks Group, then, rests on their radical methods, their

184

Chapter 4

renewed focus on quantum theory, and the applied outcome of their theorizing. On the applied side, Kaiser argues, the brainstorming sessions of the Fundamental Fysiks Group provided the “intellectual bedrock” for the field of quantum information science, including quantum computers and the quantum encryption technology used in bank transfers and electronic voting.43 Kaiser also associates the Fundamental Fysiks Group with Steve Jobs and Steve Wozniak’s development of the Apple computer, arguing that “alongside the Fundamental Physics Group, cutting-edge technology found an easy place within the burgeoning counterculture.”44 In the first instance, Kaiser makes a strong claim for the influence of the Fundamental Fysiks Group, suggesting that they had a direct impact on later technological innovations that made use of concepts that they studied and wrote about. In the second instance, Kaiser makes the weaker claim that the group shared a cultural gestalt with those who developed pioneering technology. Kaiser’s argument about the influence and legitimacy of the Fundamental Fysiks Group rests on a number of other claims: that from their inception they received attention from mainstream media, that their ideas received support from mainstream and sometimes renowned physicists, and that their work was published in respected journals. Kaiser cites, for example, an early review article on Bell’s theorem by group member John Clauser that became a classic and has been cited nearly eight hundred times in scientific literature.45 Kaiser also includes reference to current widespread efforts to reconcile quantum physics with relativity, claiming that “every single one of the now-standard responses for how to accommodate Bell-styled nonlocality with Einstein’s relativity came in either from participants in the Fundamental Fysiks Group or from other physicists’ concerted efforts to comprehend or critique their ideas [emphasis in original].”46 Kaiser similarly cites the fact that Nobel Laureates now debate the interpretation of quantum physics, that textbooks now commonly include topics such as Bell’s theorem and quantum entanglement, that curricula now include efforts to expose students to “quantum weirdness,” and that “where once influential leaders of the field had castigated philosophy as a waste of time—even when it came to plumbing the deep mysteries of quantum mechanics—the latest journals, conferences, and

New and Post-New Age Appropriations

185

books on quantum information science feature contributions from cardcarrying philosophers alongside those from professional physicists.”47 In Kaiser’s narrative, the Fundamental Fysiks Group had a direct impact on the renewed interest in quantum physics that emerged in the 1970s and that carries through to the present, as well as a direct impact on technologies related to quantum effects. At the same time that Kaiser applauds the mainstreaming of the ideas of the Fundamental Fysiks Group—a process that he describes as beginning almost from its inception—he also celebrates its outsider status, for it is this status that challenges long-held assumptions about the hierarchical nature of scientific study and production. Along these lines, Kaiser argues that “the multiple entanglements between the fundamental physics group and leading physicists of the day strain philosopher Karl Popper’s great good hope that clear criteria might demarcate authentic science from pseudoscience.”48 Kaiser writes that while marginalization itself will not fundamentally challenge Popper’s rules of demarcation, examples like Fundamental Fysiks Group can teach us other important lessons about how science works. Hippie physicists’ exploits illuminate the relationship between science and counterculture, the intertwining of ideas and institutions, and the ultimate roots for the revival of interest among today’s leading quantum physicists in questions that had once been derided as “mere philosophy.”49

Popper’s demarcation is strained precisely because the ideas of the Fundamental Fysiks Group are assumed to have gone from marginal to mainstream. It is precisely in this crossing over from marginal to mainstream, then, that the ideas of the Fundamental Fysiks Group challenged cherished beliefs about what constitutes “real” science. From this perspective, their marginal status and association with counterculture was necessary for their science to be to be revolutionary, and it was especially their more offbeat New Age activities that accorded them their unique status. It seems likely that the Fundamental Fysiks Group participated in a movement away from applied physics toward a more theoretically oriented physics, as well as a renewed interest in quantum concepts, particularly entanglement and nonlocality. It is certainly true that, in the last two decades, Bell’s theorem has been central to research in quantum

186

Chapter 4

information science, where the nonlocality of quantum theory underpins many of the advantages afforded by a quantum processing of information. It is more difficult to prove that the Fundamental Fysiks Group had a direct, causal influence on the resurgence of interest in quantum physics. It is also unclear to what extent the ideas that made the Fundamental Fysiks group what it was, and what truly put it on the margins— notions such as the role of consciousness in quantum effects or quantum parapsychological faculties, for example—carried over into mainstream physics. As Kaiser himself acknowledges, the two works connected to the Fundamental Fysiks Group that received the most recognition were The Tao of Physics and The Dancing Wu Li Masters. Both of these books contain excellent, accessible explanations of quantum physics, both have been central to the popularization of physics in general and quantum physics in particular, and both are admirable for the scope of their undertaking. Their primary reception, however, has been popular rather than academic, and in this respect they don’t challenge Popper’s demarcation. The blending of mainstream and marginal physics seems best represented by figures such as Wigner and Wheeler, whose major contributions were to orthodox quantum theory (Wigner), and to relativity, nuclear fission, and black holes (Wheeler), but who also explored in some contexts the interaction between quantum phenomena and consciousness. Their example also doesn’t pose any challenge to the demarcation problem, however, since it doesn’t represent a cultural shift so much as an individual one. Still, it remains true that the Fundamental Fysiks Group represented a new and refreshing way of doing science that broke free from the established, conservative science community. If the Fundamental Fysiks Group was not as singularly influential as Kaiser claims, it was nevertheless notable as a cultural phenomenon in its own right—one that combined the spirit of the New Age with a passion for quantum physics. QUANTUM POLITICS

Quantum politics must be based on the unpredictable, contradictory nature of human beings. —Theodore Becker

New and Post-New Age Appropriations

187

In an attempt to generate symbolic language for both new methodologies in political science and new forms of political organization and sociality that better reflect our contemporary historical reality, a handful of people have turned to quantum physics. The field of quantum politics is relatively new, having begun in the mid-1980s and continuing on through today. Several of those who write in the field draw heavily from the language of New Age quantum mysticism, while at the same time they attempt to update the language and concepts to address contemporary social, political, and economic realities. Almost universally, they oppose their models to the Newtonian atomism that funded the eighteenthcentury philosophy of self-interested liberal individualism. Within the field of quantum politics, the Newtonian social structure is critiqued for the way it supports the abstract, independent, and isolated individual.The persistence of liberal individualism is viewed as both anachronistic and counterproductive—it is not only inappropriate for our contemporary global socio-political and economic relations, but also complicit in perpetuating a dynamic based on conflict and hierarchy. Efforts to generate a new politics based on quantum logic can be split roughly into two camps, although there is considerable overlap: those who tend to embrace a New Age rhetoric that invokes Eastern and especially Buddhist philosophies and are concerned with generating a new worldview, and political scholars concerned with developing a new methodology for better understanding personal, national, and global relations. Along the lines of the first camp, Danah Zohar, in her article “Forces of Reaction,” emphasizes relationality and context as key quantum dynamics that can inform a new sociality. Like proponents of quantum mysticism, Zohar invokes the interaction between the observer and the observed in quantum physics, similarly locating the act of observation at the level of human consciousness. Zohar replaces the Newtonian subject/ object split by what she calls “observer participancy,” wherein “the quantum observer stands inside what he observes, his own goals and consciousness and intentions help to ‘make’ the reality he observes. This replaces the old objectivity with a new kind of ‘truth within a situation,’ or engaged truth.”50 Rather than substituting the old objectivity with a new subjectivism, however, Zohar’s observer/participant is a social being who pursues greater dialogue between disparate cultures.

188

Chapter 4

In a curious paradox, Zohar’s observer/participant model stands for both the individual’s power to create their own reality, and a participatory politics in which reality and truth, viewed as contingent, are created though a collective, dialogic process. The individual’s potential to create his or her own reality appears directly to contradict the notion that “reality” is explicitly social and constructed through a participatory process. The paradox arises in part from the problem of the original use of “observer,” which, as Stenger notes, allows room for the experimenter’s consciousness to enter into the equation. It also arises from the way in which the measuring instrument’s impact is construed. In the original sense of the observer effect, the measuring instrument “creates” the result through its impact on the behavior of the system. Add consciousness and give it an existential spin, and you arrive at Zohar’s intention-driven individual, who creates their “reality.” On the other hand, view the measuring instrument (now translated into consciousness) as an integral part of the overall experiment, which includes matter, measurement, and result, and you arrive at the social notion of a system with interacting parts. From this, one can derive the notion of a dialogic, participatory sociality—the model that Zohar ultimately privileges. One of the primary quantum metaphors that Zohar invokes is wave/ particle duality, or more specifically, the metaphor of the wavicle.51 The term “wavicle” provides the metaphor for the kind of pluralism Zohar advocates, which she distinguishes from both a “plural dualization” that offers an unambiguous, either/or choice between absolutes, and a postmodern eclecticism that leads to “the privatization of meaning, [and] the fragmentation of shared, public meaning.”52 According to Zohar, a pure “wave” state leads one to become “brainwashed, overwhelmed by the other’s view,” and a pure “particle” state leads to a situation wherein one “ merely tries to win the argument.”53 Zohar’s “third way” is based on the “both/and” of the wavicle, which involves a constant, enriching negotiation between individuals or discrete groups, and a larger collectivity. Zohar advocates a creative, Socratic dialogue that involves an intermediate state between pure waves and pure particles and where “we treat both our own and the other’s viewpoint as equally worthy of consideration.”54 Throughout, Zohar’s program for social change is largely abstract and figurative, which may explain the unsatisfying nature of her conclusion. With reference to this Socratic dialogue, she writes, “What conditions

New and Post-New Age Appropriations

189

would foster it? This question is vital if our pluralistic society is to regain meaningful social cohesion, a consensus that recognizes and thrives on difference.”55 To be sure, Zohar is engaging more in social philosophy than in imagining new political structures; nevertheless, one might hope for a brief indication of some of the conditions that might provide a foundation for her new social dynamic. In Quantum Society: Mind, Physics, and a New Social Vision, Zohar teams up with Ian Marshall to recapitulate her emphasis on achieving creative consensus and unity while at the same time celebrating diversity. Again, they invoke the wave/particle model, the embracing of which will enable us to see that diversity and consensus are not mutually exclusive, because the conjoined yet separate wavicle suggests a constructive mediation between localized identities and the collective. I want to focus here, however, not on Zohar and Marshall’s wavicle model, which simply reiterates Zohar’s claims in “Forces of Reaction.” Rather, I will take up their promotion of “contextualism,” a term they claim to have derived from quantum philosophy. Rather than associating the responsive reality of contextualism with the observer effect, Zohar and Marshall claim that “the full force of quantum contextualism shows itself most dramatically in the famous ‘two-slit experiment.’”56 In quantum contextualism, “reality shifts its nature according to its surroundings,” in the same manner that the electron or photon is in “constant creative dialogue with its environment.”57 Here, Zohar and Marshall make the same move that Zohar makes in “Forces,” combining the assertion that we create our own reality with an ideology of participatory social interaction, and the same contradiction pertains. Along with anthropomorphizing “creative” subatomic particles, Zohar and Marshall get the double-slit experiment wrong. They begin by describing how a photon’s behavior is dependent on whether we erect a “particle detector” or a “wave detector” in their path (which they identify as the screen through which the observer hopes to observe the particles as they strike).58 This depiction of the experiment is incorrect.There is no such thing as a “wave detector,” and the screen behind the plate that contains the double slits is simply used to register, not participate in, the results. The experiment is much more straightforward: if photons are allowed to travel through the slits undetected, they will always reveal a wave pattern on the screen. If a detecting device is used, they will always

190

Chapter 4

behave like particles. Thus any interference by the observer will always cause the wavefunction to collapse into a single (particulate) state, and no interference will always result in a wave pattern. This misinterpretation compromises the main point that Zohar and Marshall derive from the experiment: that we see what we are looking for. Zohar and Marshall carry this misinterpretation over to the social realm when they argue: The relationship between the observer’s way of looking at a quantum experiment and the outcome of what he sees is very like the link between our social expectations and what we perceive. If we look at a group of people as a collection of individuals, we will perceive them as individuals. But if we look at the same group as a collective unit, take an “average over the individuals,” we will see a collective phenomenon. More strongly still, the way we look at a group of people can actually affect the group’s behavior.59

In fact, whenever we “look” during the double-slit experiment, we will always “see” the same thing: particulate behavior. Only when we are not looking will a wave pattern appear. To pursue their own analogy, this means that the “collective” only exists when we are not interfering or interacting with it in any way.This, along with the vague claim that “we” affect “their” behavior, does not support the dialogic unity in diversity that represents their ideal society. In “Physics, Buddhism, and Postmodernism,” Dawne McCance critiques the “atomistic pathology” of abstract individualism for promoting a logic of fragmentation that “pits person against person, creates divided nations and political and religious factionalism, and ends in violence and war.”60 While Newton comes under fire, McCance’s main critique is directed toward Descartes and certain outdated Cartesian concepts such as “entity,” “totality,” and “transcendence” that render nature an object independent of humans.This cluster of concepts, writes McCance, creates a view of the transcendent subject founded on a “social physics” of abstract individuality wherein a disconnected, encapsulated, self-justifying subject operates independently of all others.61 According to McCance, this worldview threatens not only our physical survival but also the survival of human community.62 Like Zohar, McCance finds inspiration for a new worldview from the quantum claim that we can no longer sustain the notion of a division

New and Post-New Age Appropriations

191

between subject and object. From this, McCance extrapolates the primary principle of a new, postmodern and post-Cartesian epistemology: a movement away from the ideology of the detached observer and toward an “interactive point of view which recognizes the participation of the knower in the known,” and is defined by notions of “process,” “context,” and “relationship.”63 She also finds inspiration in Bohm’s rejection of the theoretical assumption that we can provide direct descriptions of reality— an assumption that leads us to believe that the world consists of separate fragments, and then to act in such a way that we produce the very fragmentation implied in our attitude. Invoking Fritjof Capra’s book on holistic systems theory, The Turning Point: Science, Society, and the Rising Culture, McCance offers a string of metaphors including “holism,” “system,” and “ecology” that are intended to provide an alternative to social fragmentation.64 The weakness of McCance’s argument derives in part from the fact that she attempts to link too many conceptual frameworks: Buddhism, quantum physics, and postmodernism. In fact, she repeatedly refers to “contemporary postmodern quantum physics” without any consideration of the anachronism of associating quantum physics with postmodernism. McCance’s strength and primary focus seems to be Buddhism, and in her attempt to fit both quantum physics and postmodernism into a Buddhist paradigm, she at times collapses quantum physics and postmodernism, as if they derived from the same historical moment and expressed the same logics and priorities. McCance also alternates between references to epistemology/theory, and political ethics; at times, she seems to be reaching for a new philosophy, at other times, a new form of political engagement. While she includes specific references to politics in her critique of Cartesian and Newtonian worldviews, when she presents an alternative, she fails to advance beyond detached, figurative terms (process, context, relationship, holism, system, ecology) that lack real political content or context. Both Zohar and McCance are suggesting a new manner of social and political engagement. It is not clear, however, how thinking of other individuals and oneself in a different way will actually work on the ground. How will an emphasis on “unity in diversity,” “process,” or “context” manifest itself structurally? How will it transform educational or religious institutions, the interaction between levels of government, the shape of

192

Chapter 4

non-governmental organizations, or voting practices? To be fair, both McCance and Zohar are exploring a general shift in our relation to the world and to ourselves, and their interest is more in supplying the metaphors that would articulate this worldview than in developing any concrete political program. As an editorial review of Quantum Society in Kirkus Reviews observes, however, “without concrete examples to illustrate such theories, this … work remains abstract and speculative.”65 This failure might be forgiven, if their language did not clearly call for new forms of collective organization, forms that they stop short of delivering. Those who identify more as political scientists or political philosophers typically turn to quantum physics not just for new metaphors of sociality, but also for new methodologies in political science and new concrete forms of political engagement. Law professor and political scientist Asghar Kazemi, for example, focuses on the methodological challenges that face contemporary political theorists. He observes that “in recent years many authors have questioned the wisdom of continuing to rely on the Newtonian philosophy to deal with the emerging problems in world affairs and domestic issues.”66 According to Kazemi, a political methodology that relies on Newtonian cause and effect, sovereignty, and political determinism no longer reflects the complex global political, economic, and social realities. Kazemi states that the anxiety political scientists experience when facing the inherent uncertainty in politics impedes their endeavors—to the extent that those endeavors include the pursuit of objective, provable, and enduring theories about society. As a solution, Kazemi suggests embracing the Uncertainty Principle, which he links with the postmodernist assertion that all “truths” are subjective. If political scientists accept that there are no immutable laws to be discovered, that everything is probabilistic and perhaps even random, and that they cannot predict what will happen in the future, then, says Kazemi, they will no longer suffer from the malaise that grows out of the fact that they can only ever achieve partial and contingent knowledge. Kazemi briefly extends his conclusions to the larger political arena when he argues that embracing the idea of uncertainty means accepting that we will disagree and being willing to admit when we are wrong—both of which will lead to a more open society. Ultimately, however, Kazemi is more interested in establishing new methodological principles of analysis than new forms of political engagement.

New and Post-New Age Appropriations

193

Kazemi connects his quantum politics to both postmodernism and chaos theory. His justification for linking quantum theory to postmodernism rests solely, however, on Christopher Norris’s questionable claim in Quantum Theory and the Flight from Realism: Philosophical Responses to Quantum Mechanics that “quantum theory has been a major influence on postmodernism.” In a move characteristic of those who link quantum physics and postmodernism, Kazemi refers to the fact that both fields question the legitimacy of truth claims. This is not precisely true; while quantum physics demonstrates the limits of our knowledge, its theory rests on complex and meticulous mathematics that have been refined, tested, and experimentally proven over time to be correct. Furthermore, for Kazemi, questioning truth claims in politics appears merely to mean recognizing that when politicians claim to be telling the truth, they are usually hiding something. Kazemi’s use of chaos theory and the butterfly effect to describe the sensitivity of global systems to unknown factors and variables is somewhat more coherent, but he does not suggest how one might do anything more than acknowledge this sensitivity. Despite his focus on methodology, Kazemi does not offer any direction about what a new approach might look like. What impact might the embracing of uncertainty have, for example, on the (often contradictory) use of statistical data to make truth claims? How might it affect assessments of the relative value of quantitative versus qualitative analysis? How would it affect the researcher’s relation to his or her object of study? More problematic, however, is the fact that Kazemi has no reason to invoke the Uncertainty Principle in the first place.Were Kazemi to take the principle seriously, he might have used the impact of the measuring device on the object to reconceive the relationship between the political scientist and their object of study. Even here, the notion of the observer/participant has been amply theorized, particularly in the field of anthropology and without any reference to quantum theory. Instead, however, Kazemi is merely enjoining political scientists to embrace the fact that they can never be certain of what they think they know, and thus should open themselves up to alternative points of view. Kazemi takes a standard dictionary definition of the word “uncertainty” and then cloaks it in references to quantum physics, postmodernism, and chaos theory in order to lend a cutting-edge feel to his methodology.

194

Chapter 4

In the anthology Quantum Politics, editor Ted Becker argues that substituting quantum theory for classical physics as the basic metaphor for political science will move us toward a more participatory democratic theory that offers a more humane and peaceful alterative to the dysfunctional regimes of the present day.67 Like several of the authors in the anthology, Becker places a great deal of emphasis on the notion of democracy as participatory—drawing, like McCance, from the observer/ participant formulation of Heisenberg’s Uncertainty Principle.68 Becker extends the notion of uncertainty to argue that quantum phenomena function according to chance, and then he creates an analogy to a form of political engagement that “is based on the unpredictable, contradictory nature of human beings,” and the fact that “chance plays a major role in any political organization.”69 The association of the Uncertainty Principle with “chance” is a fallacious one, however; the impact of measurement and the reliance on probabilities does not mean that quantum behavior is “random”—it merely reflects an inherent limitation on what we can know about it. In Becker, as in the anthology in general, it often is not clear whether the primary interest is in developing a new methodology for political science, or in conceiving a new form of politics; nevertheless, some key through lines emerge. Many of the authors use the observer effect to argue that quantum politics provides a conceptual ground for participatory democracy. Through a series of metonymic associations, they then argue that quantum politics emphasizes relational, interdependent, and collectivity-oriented engagement rather than atomistic, isolated individualism; that it promotes direct democracy; and that the dynamics of social systems and understanding of social “realities” is actually dependent on the priorities and knowledge of the observer.This last claim appears to be another version of McCance’s argument that we can create our own reality, and while the recognition of this fact might help us appreciate the contingent nature of “reality,” its determinism is at odds with the principles of collective participation. In “The Quantum Mechanics of Politics,” Flora Lewis—who is admittedly neither a political scientist nor a political philosopher—ventures to offer more concrete political applications for a quantum worldview. In what has become a trend in quantum politics, most of Lewis’s argument is again based on the elevation of common uncertainty to the

New and Post-New Age Appropriations

195

Uncertainty Principle. According to Lewis, “the demand for certainty represents a danger, because the illusion of being absolutely right acts as moral blinders.”70 Lewis asserts: “There is a profound urge to grasp at whatever will allow one to say, ‘I know I am right and you are wrong. It’s as clear as the nose on your face.’ It is much harder to live with the value of uncertainty—a value that requires respect for and a degree of deference to other peoples’ observations and ideas.”71 In reference to the “close calls” during the nuclear arms race, Lewis rejects “elaborate equations about the balance of terror and the balance of forces” that offer the illusion of truth, and instead argues for national and international relations that acknowledge the fundamental uncertainty that defines human relations.72 For Lewis, this uncertainty is not inherently destabilizing; instead, it “always leaves room for hope of doing things better” and she encourages us to keep trying to get it right.73 Nevertheless, asserts Lewis, uncertainty underlines the need for checks and balances in the political arena, including the separation of governmental powers, the independence of institutions, and the regular consultation of the electorate and freedom of citizens to publish information and opinions.74 It is not clear how Lewis gets from uncertainty to the kinds of checks and balances that she advocates, and her other political applications of quantum phenomena seem even more of a stretch. Complementarity leads to the assertion that any attempt to impose order and stability is impossible; the “wave of probability” inspires the recognition that new troubles and challenges will always appear; and quantum mechanics “requires the perception that events are made up of particles, individuals,” although they “move in waves which can be discerned even if no individual move can be predicted.”75 It is true that Born’s reinterpretation of Schrödinger’s wavefunction is connected to the limitation wherein only the probability of finding a given particle at a given place and time can be defined; however, Lewis simply seems to be using “probability” in place of “probably.” Lewis offers no political application for that fact that the state of particles is only available statistically in the movement of (probability) waves, and the assertion that complementarity provides a model for the inherent instability of social relations is somewhat puzzling. In the case of many of these authors, the use of quantum concepts as a model for political structures and analysis tends to be vague and

196

Chapter 4

metaphorical. Since the field of “quantum politics” is fairly new, however, it is not surprising that these metaphors have not cohered into a mature methodology for political analysis, nor into a program for political organization, interaction, or activism.The references to holism (cast as collectivism), and the emphasis on participation and intersubjectivity is also not surprising, given their longstanding priority status in other fields of political study, and given the fact that this path has already been cleared by expansive New Age notions of quantum consciousness. Nor is it surprising that these quantum concepts have been, at times aggressively, massaged or misrepresented to fit a particular political worldview. More problematic is the almost ubiquitous use of “quantum” terms or concepts in place of perfectly serviceable everyday words or more trodden political science terminology—merely in an effort to capitalize on the latest fashion in cutting-edge analysis. There is, of course, always a danger in adding another layer of metaphor to terms that are already metaphorical. This practice is particularly fraught where quantum physics is concerned, since the language, necessarily drawn from classical concepts, is already approximate and misleading. To then import this terminology into a social or political context by way of analogy allows for even more conceptual drift, not to mention reductionist comparisons (just as subatomic systems can be both particlelike and wave-like, so can we be both individuals and participants in a community). In any introduction of the comparison of quantum physics and contemporary socio-political dynamics, one might hope for a recognition of the cross-pollination of intellectual currents at any given historical moment, and an explanation for why a scientific paradigm established seventy or eighty years ago should only recently prove such an appropriate instrument for political analysis. Absent this explanation, its use appears to be merely a function of an effort to generate novel theoretical tools for analysis. POST-NEW AGE SELF-HELP: QUANTUM GUIDES TO HEALING YOURSELF AND GETTING RICH

The EPFX/SCIO is a Quantum Biofeedback Device … [that] can be connected directly to a client by means of a “harness” around the ankles, wrists and head, but can also effectively work at a distance—through the subspace.

New and Post-New Age Appropriations

197

By adding the intent of Prosperity, Money, Abundance into various therapies, the instrument will work to harmonize those particular issues. —Quantum Prosperity Program76

As theoretical and philosophical works, both Zukav’s and Capra’s texts are more interested in universal dynamics than in the individual subject. Following a long tradition characterizing the study of the physical world, Capra and Zukav view it worthy of consideration as a wonder in itself. This orientation distinguishes them from later efforts to conscript quantum physics as a foundation for a radically subjective relation to the world. The concern with self-fulfillment is reflected in the emergence of postNew Age quantum “self-help” programs toward the end of the twentieth century, at a time when “subjectivity” enters academic and popular discourse as a dominant concern. The historical origins of this emphasis on the subjective might be attributed to a number of factors. On the one hand, it may be traced back to the fragmentation and collapse of collective movements such as the New Left and second wave feminism, and the ensuing emphasis on an eventually depoliticized and individualized identity politics.77 The rise of the Internet has similarly encouraged an increasing personalization, both in terms of individual consumption and the self-publication enabled and encouraged by social sites such as Facebook, Twitter, LinkedIn, Instagram, and so on, and the explosion of personal blogging sites. At the same time, there exists a perceived loss of control that can be traced to the acceleration of a surveillance society wherein governments, corporations, and hackers can map or steal individual identities, and to a globalized, nodal power structure that is diffuse and intangible, so that it is no longer clear who is doing what to whom. Uncertain job prospects and mounting personal debt, combined with an economic machinery that becomes less and less intelligible at the same time that it becomes more and more unstable, exacerbate this sense of powerlessness and disorientation. It is not surprising, then, that popular culture is shot through with products that promise a renewed control over the direction of one’s relationships, finances, career, and health.The offer to “take back the reins” is evidenced by the number of books that offer strategies that will assist individuals in their search to regain a lost sense of agency and control, as well as the proliferation of “life-hackers” who offer endless advice on

198

Chapter 4

how to maximize your productivity, improve your mental acuity, economize your fitness regimes, or perfect your diet. Complementing this offer of self-empowerment is an emphasis on one’s right to personal happiness and well-being that dominates popular culture, and that has been taken up by both academia and the State in the form of proliferation of metrics assessing health and happiness.78 Post-New Age “quantum practitioners” conscript such concepts as nonlocality, uncertainty, and wave/particle duality to advance their own brand of observer/observed dynamics. In doing so, they filter the fundamentals of quantum physics’ disruption of causality through an anthropocentric perspective that takes the undecidability of events at the quantum level for a form of animated potential that has, as its main characteristic, the plasticity of the material world to the human will. The original explanatory analogies and metaphors of quantum physics are literalized and then reconstituted to confer a strong cause and effect relation between human consciousness and the material world, wherein the mind is free to seize or create opportunity through the manipulation of its physical surroundings. The immensely popular 2004 film What the Bleep Do We Know? sums up the extent to which post-New Age quantum rhetoric is associated with both subjective possibility and personal empowerment.79 Toward the beginning of the film, theoretical physicist Amit Goswami states that “quantum physics, very succinctly, is the physics of possibility.”80 Later in the film, Stanford Professor of Materials Science and Engineering William Tiller observes, “if reality is my possibility, then the question is: how can I make it better, how can I make it happier?” Chiropractor and “Neuroscientist of Change” Joe Dispenza advises that we ought not to buy into the idea that we have no control, nor should we continue believing that the external world is more real than the internal. “This new science [of quantum physics],” says Dispenza, “is just the opposite. It says that what’s happening within us will create what’s happening outside of us.”The film reveals a general post-New Age emphasis on our power to create our own reality, with the goal of optimizing our personal satisfaction and prosperity. For post-New Age Quantum practitioners, the relationship between the observer and the observed proves once again foundational. Jeremy Campbell states that “the notion that a quantum happening is relative

New and Post-New Age Appropriations

199

to an observer … has been elaborated into a more daring hypothesis: namely, that observation is the whole point of the universe, and that all physical law is relative to the observers.”81 In post-New Age quantum consciousness, subjectivity expands to include the entire universe, whose existence becomes an artifact of human consciousness, at the same time that it contracts, to reflect the will of the individual. Here, the dynamics of matter are harvested to construct a one-way causal relation between the consciousness of an observer and his or her physical surroundings. At the same time, the laws of quantum statistical probabilities are transformed into the notion of human potentiality and empowerment. The conceptual framework of quantum physics, combined with unrelated concepts in other fields, is used to support a host of loosely related methodologies that, once operationalized, will demonstrate how to take advantage of the infinite choice and possibility available to the individual. POST-NEW AGE MYSTICISM: HEALING THE MIND AND BODY

Kidneys never make decisions alone; they work in constant consultation with the quantum mechanical body. —Deepak Chopra

After Capra, the best-known (and most successful) popular proponent of the connection between consciousness and quantum physics is physician Deepak Chopra. According to Chopra, we participate in a cosmic connection to the quantum world, and the “inconceivable” region from which photons emerge is the same as that from which we fetch thought and experience. Chopra acknowledges that physicists could object that he is just making metaphors, and that an unlocatable particle is fundamentally different from the hidden worldmind; however, he insists on the notion of quantum nonlocality on a cosmic scale. He argues that a particle separated by immense distances in spacetime from another particle really does know what the other one is doing, and that this knowing demonstrates the fact that the entire universe is knitted together by a kind of “memory network,” or universal consciousness.82 Chopra offers an interpretation similar to Zukav’s, wherein the original metaphorical use of the word “know” to describe the interaction of entangled particles is translated literally to mean that particles at a distance are actually conscious of

200

Chapter 4

one another. Chopra then extends this interaction on a cosmic scale in order to support his claim that the universe itself is conscious. The existence of a universal consciousness is essential to Chopra’s main business: assisted quantum healing. Since both body and mind ultimately are made up of the cosmic stuff of consciousness, writes Chopra, we can cure all our ills simply through the application of our mental energy to our bodies. “If asked for a definition of quantum healing,” he writes, “I would say this: quantum healing is the ability of one mode of consciousness (the mind) to spontaneously correct the mistakes in another mode of consciousness (the body).”83 According to Chopra, our organs work in “constant consultation” with the individual quantum body—itself an interconnected field of intelligence and experience. The specific connection between mind and body begins at the atomic level, but it also extends into more macrocosmic elements such as molecules and DNA itself, which he claims also possess quantum properties.84 As with all quantum events, says Chopra, something inexplicable happens beneath the surface to form the all-knowing intelligence of DNA. DNA lives at the point of transformation, constantly delivering messages from the quantum world to ours, tying new bits of intelligence to new bits of matter.85 The point of “diving into” the realm of the quantum body, then, is to “change the blueprint itself,” thus transforming our individual physiology in its entirety.86 As one of many examples, Chopra presents the case of a female patient who healed herself of cancer, with Chopra’s guidance, through the mere application of her mind.87 Chopra calls this patient’s case a “quantum event” because the fundamental transformation she enacted on her body went deeper than her organs and traveled directly to the source of the body’s quantum existence in universal time and space. Whereas proponents of quantum politics use quantum language as metaphorical models for their ideal of a transformed sociality, Chopra at once literalizes and expands the already metaphorical concepts of quantum phenomena. And whereas New Age philosophers of quantum mysticism are interested in uncovering a reality deeper than that available to our senses, Chopra’s utilitarian approach emphasizes human intention and our ability to capitalize on quantum dynamics for our own benefit. Chopra tends to use quantum terminology (combined in an undifferentiated way with the mass/energy theory of relativity) in a manner that is

New and Post-New Age Appropriations

201

both allusive and elusive—invoking quantum concepts without talking about their origins or specificity. This decontextualization and vagueness differentiates him from Capra and Zukav, who are consistently precise and thorough in their explanation of quantum physics. Chopra’s vagueness works to his advantage; while Zukav is forced to make a series of awkward deductions to achieve his point, Chopra can simply invoke the language and apply it as he wishes. He frequently uses the term “quantum leap,” for example, to describe the alchemical transformations that emerge out of the application of mind to body, and his strategy mystifies these transformations at the same time that it grants them a scientific air. Despite his fuzzy science and radical claims, Chopra has managed to convert his assisted quantum healing into a multimillion-dollar financial juggernaut that is founded on a variety of workbooks, workshops, and consultations available to the consumer wishing to benefit from Chopra’s guided quantum self-help program. “Independent Research Professional” James A. Putnam offers a more novel take on the notion of quantum intelligence. According to Putnam, our experience comes to us through the intermediaries of photons: our bodies are constantly bombarded by photons that originate from an immeasurable number of sources, each photon striving to pass on some small bit of cosmic information.88 The photons “notify” us, Putnam says, to search inside our being and discover a form of knowledge that is always already there for the taking.The information that we can glean from photons, properly selected and interpreted, then wakens our genetically inherited intelligence, potentially leading to self-actualization.The information that photons provide must be decoded, however—rerouted and analyzed internally with the help of an enlightened guide. Summed up, Putnam’s theory represents an anthropic interpretation of the subatomic realm that casts humans as the central actors in and beneficiaries of animated particles that are constantly striving to help us achieve our potential. Here, particles are not merely “knowing” or “conscious”; they now possess a human-like intentionality whose main goal is to help us achieve our potential. Particles are miniature homunculi who exist to serve humankind. It is never made clear, however, precisely how quantum physics enters into the picture. The notion that we are constantly bombarded by atomic particles is hardly unconventional, although the proposition that they speak to us certainly is. Putnam provides no explanation

202

Chapter 4

for how he arrives at his theory, omitting even the typical slippage from observer to participant to universal consciousness. Putnam is a researcher, not a vendor, which distinguishes him from the proliferating “quantum touch” sites that disarticulate the term quantum from its origins even more radically, and that offer energetic healing using light touch, controlled breathing, and a variety of other techniques to bring about physical and spiritual well-being. Along with selling a host of products, these sites promise to put visitors in touch with quantum touch therapists, whose techniques appear to focus on facilitating healing by amplifying our life-force energy or chi. In “Resonance, Life-Force, and the Principles of Quantum-Touch,” Richard Gordon invokes two scientific concepts in promoting its method—neither of which are especially associated with quantum physics.The first is “resonance”: the tendency in a system for even small forces to produce large-amplitude oscillations. The second is “entrainment,” wherein two interacting oscillating systems assume a common period. Using the quantum touch technique, the description says, “we can create a high frequency of life-force energy. If we place this field of high energy around an area of pain, stress, inflammation, or disease, the body can entrain to the higher frequency, thus amplifying the body’s ability to heal itself.”89 The quantum touch is a personalized touch; through direct contact, the consumer’s own innate capacity to heal himself or herself is nurtured and expanded, according to the site. As a way of establishing the relation between scientific terms and modes of healing, the use of science rhetoric here dispenses altogether with metaphors grounded in quantum theory. Instead, it relies on a metonymic strategy of annexing “amplitude” to New Age “energy fields” (which suggest, but have nothing to do with quantum fields) in order to create a rhetorical bridge that establishes the connection between science and the Eastern concept of chi. This metonymy is even more evident in the site’s title, where the relation between the quantum and healing is established merely by juxtaposing the terms “quantum” and “touch.” While purportedly grounded on scientific methodologies, there is no real attempt to connect this mode of healing with actual quantum concepts. The only quantum presence is in the name of the practice, and the term “quantum” is reduced to a commodified cipher. This process of commodification becomes overt on Jack Lyons’s Quantum Touch site, where

New and Post-New Age Appropriations

203

the registered trademark symbol at once suggests the site’s preeminence (its process must be protected from imitators) and underscores the ultimate conversion of “quantum” into a commodity.90 Most quantum-based healing programs exist on the Internet. While quantum healing programs do continue to appear in the form of books, the Internet’s reach, flexibility, and interactivity offer several advantages for the packaging and sale of quantum healing. Sites can offer layered access, wherein the homepage functions as a teaser that contains vague reference to some sort of novel quantum approach. To learn more about the program and to realize the promised results, consumers must navigate away from the homepage, where they are typically presented with the opportunity to register for workshops and buy a host of products. Internet sites can offer simultaneously a number of unique products and workshops, at once allowing visitors to individualize their purchases according to their individual means and needs, and enabling the vendors to adjust their use of the term “quantum” according to shifting cultural priorities. PROFITING FROM THE QUANTUM: GET-RICH SCHEMES

When you use these techniques, what you’re doing is connecting with the Quantum Universe on a very scientific level. Like iron filings to a magnet, you’re literally “sucking in” all the resources you need in your life—effortlessly. —Robert Anthony

The commodification of quantum logic present in the marketing of quantum healing programs is accelerated in personal wealth-oriented “quantum enrichment” programs. In these mostly online programs, financial gain is no longer associated with the necessary cost of sustaining a business whose main goal is to help others. Rather, it is offered directly to consumers in the form of online get-rich quick schemes that propose applying the principles of quantum physics to achieve near-instant financial success. Many of these sites leaven their focus on the secrets of making money with healthy doses of rhetoric about general payoffs that include finding peace of mind, developing a healthier body, and resolving inner conflict; however, the prospect of getting rich is never far from sight. On his site http://www.robertanthonyonline.com, Robert Anthony dispenses altogether with the packaging of balance, health, and happiness,

204

Chapter 4

and gets right to the point: making money.91 “Knowing how to be rich,” observes Anthony, “is far more important than the actual making of money.”92 After much seeking, along the lines of a mystic pilgrimage, Anthony realized that what he needed was not some new “system” or set of “teachings” that promised wealth. Instead, he says, “What I desperately needed was PROOF: Scientific, Indisputable, Immediately Verifiable Proof … And that proof came in the form of QUANTUM PHYSICS … It was irrefutable. It was scientific, and it was indisputable.”93 Anthony’s program is offered via a six-week “crash course” consisting of a series of audio presentations made available to the consumer one week at a time. The program appears to be founded upon a loose combination of the observer/participant principle of quantum physics, the Newtonian law of attraction, and, once again, New Age energy fields.94 By changing our energy fields, Anthony promises, we can produce physical changes in our surroundings, literally attracting financial success.The combination of the law of attraction with quantum physics is not unique to Anthony; a host of “prosperity” sites offer selected and detailed explanations of quantum physics, including the double-slit experiment and quantum entanglement, and then spins out a relationship between each phenomenon and the law of attraction. As with Anthony, the term “law of attraction” is used to introduce both intentionality and familiarity into the methodology, while the less familiar quantum concepts are used both to mystify the process and to lend it scientific caché in an already crowded field of vendors who link the law of attraction to making money.95 Anthony is so confident that his clients will make revolutionary financial gains that he is willing to “swallow all the risk and lay bare [his] most prized secrets for the taking.”96 Anthony’s money-back guarantee rests on the immutability of the physical law: “I guarantee that this program will work for anyone who actually uses it. I know that it is impossible to fail with this program. It is impossible not to get results because that would be the first time in history that the laws of physics failed [emphasis in original].” Anthony combines the scientific legitimacy of his program with the typical testimonials, as well a classic American trope: the story of himself as self-made man who rose from generational poverty to great riches.The combination of this familiar, foundational American myth with the trendy science of quantum physics allows Anthony both to tap into

New and Post-New Age Appropriations

205

enduring aspects of the American psyche, and to capitalize on the typically American demand for both proof and novelty. On his site quantumjumping.com, Bert Goldman launches arguably the most improbable success-oriented use of the oft-invoked concept of quantum jumping. Goldman tells his reader: [My] Visualization Technique Will Transform You Into A Universe-Hopping Utopian Being. Once I’ve shown you how, you’ll be able to use the untapped power of your mind to ‘jump’ into alternate universes, and visit alternate versions of yourself who already have all the skills, knowledge and experience you desire … The smarter you. The richer you. The healthier you. The sexier you.They’re all out there, and all you need to do is talk to them.97

Goldman’s slick, multilayered site has several modest free offers; clicking on any one of them will take you to the $200 nine-CD collection of the “Ultimate Edition of Quantum Jumping,” and clicking on the CD collection link initiates a video offering access to the many quantum jumping courses associated with Bert Goldman.To ground his claims, Goldman combines the notion of a quantum jump with the Many-Worlds Interpretation, which denies the wavefunction collapse in favor of the position that all potential (superimposed) states persist, and that every possible outcome of every quantum event survives within its own history or world. While the scientific concept of quantum jumping has nothing to do with either wavefunction collapse or superposition, their combination provides a kind of scientific logic to the otherwise dismissible idea of hopping between universes. Like Anthony, Goldman provides a long list of experimental verifications in the form of enthusiastic testimonials.“Sarah,” for example, affirms how she quantum jumped to a successful version of herself and then visualized herself interacting with this other self. After decoding a cryptic phrase from her other self, concerning “cleaning up,” Sarah experienced an epiphany and subsequently started up a profitable business readying houses for the market. “Stan” had always wanted to write, so he quantum jumped into a parallel world and connected with his already widely published doppelgänger to learn the craft of writing and to pick up some strategies for successful publishing. In another jump, he encountered his globally recognized public speaker self. As a result of this encounter, Stan says, he now knows that if he should be asked to speak he “can do so with

206

Chapter 4

all the flair and skill that [he] learned during [his] Quantum Jump.” Finally, “Vincent,” a would-be musician, visited his own multiple gold recordwinning double, who provided him with ideas, titles, and in some cases music for new songs. Instead of expanding subatomic particles on a cosmic scale like Chopra, or humanizing them like Putnam, Goldman shrinks individuals to the level of subatomic particles who move at will between different versions of themselves. The social implications of this relationship to atomic phenomena are far from insignificant; in fact, Goldman’s model nicely illustrates the socio-political dynamics informing post-New Age quantum self-help discourse. Like the liberal humanism of Newton’s time, the consumer-oriented atom/individual is thoroughly disconnected from any larger political or social reality, and the focus is solely on individual gain. Unlike Newton’s liberal humanism, the instantaneous transformation that Goldman offers releases people from the arduous task of learning from life experiences and applying these experiences in a gradual process of overcoming social and individual obstacles to gain insight into their personal nature and societal context. Most get-rich sites combine the economic term “prosperity” with the more generalized term “abundance”—the latter suggesting a kind of undefined plenitude that will banish hardship and need. Using terms associated with the economic register but that also possess extraeconomic connotations increases the chance that anything of benefit that happens to the client during or after taking the program can be interpreted as a direct outcome of their participation. Finally, by casting money as part of a cosmic bounty available to everyone and dispensed by a generous and caring universe, quantum enrichment practitioners anoint people with the belief that they deserve to be wealthy. That quantum concepts are so mysterious and counterintuitive only helps practitioners to emphasize the fact that they possess some sort of elemental secret—“secret” being a term that appears at least once and typically many times on each get-rich site. At the same time, the inaccessibility of quantum concepts to laypeople justifies the need for a quantum master to guide consumers eager to better realize their personal potential, achieve health, or simply make money. Many of these gurus speak of traveling around the world, studying or apprenticing under primarily Eastern mystics who revealed secrets reserved only for the initiated.These insights

New and Post-New Age Appropriations

207

prove futile, however, until practitioners learn the secrets offered by quantum physics. Eastern methods of achieving enlightenment typically require disciplined commitment, and take years or even decades to yield results. By taking this burden upon themselves, quantum gurus release clients from a difficult and protracted labor and instead offer them instant gratification. For people dealing with the modern malaise of loneliness, alienation, and stress, people who lead hectic lives with little time or patience for undertaking years of arduous meditative practice, the instant personal alchemy promised by quantum-fueled transformation offers an attractive alternative. Because almost all quantum experts cloak their techniques in secrecy, it is often difficult to discern, without signing up and paying for the workshops/CDs/books/consultations, precisely how the practitioners incorporate the premises of quantum physics into their actual methodologies. Indeed, many of the references to Eastern mysticism seem precisely to serve the function of situating the vendor’s “quantum knowledge” at an inaccessible remove from the consumer. The week-by-week teaser lists of topics, however, suggest that most quantum programs, broken down to their constituent parts, offer mostly traditional instruction in sound business practices, the power of positive visualization, selfconfidence, and commitment. Quantum testimonials function in a similarly conventional way, with the exception that they add quantum references that transform these testimonials into experimental evidence that legitimizes a foolproof (because scientific) product. CONCLUSION

The language through which scientific advancements are relayed is neither neutral nor transparent. Rather, this language reflects specific social, political, and cultural needs and expectations, as well as specific constellations of hopes and anxieties. Constructions and applications of atomic discourse provide a material touchstone that is no less tangible than any other aspect of scientific enquiry. From holistic quantum consciousness, to the participatory democracy of quantum politics, and finally to the contemporary commodified and personalized atom of quantum healing and quantum enrichment programs, one consistently finds in atomic models the expression of societal tendencies. In this sense, our

208

Chapter 4

understanding of the atom and its component parts has served and reflected worldviews just as much as it has reflected scientific advances, with existing societal priorities driving the relative degree to which a given model is accepted and applied. Reflected in the work of Fritjof Capra and Gary Zukav, New Age quantum consciousness retains a close allegiance to the theory and experimental results in the physics of the twentieth century. The authors construct themselves as teachers and guides whose main agenda is simply to enlighten the public on the semi-mystical wonders of the cosmos and our relationship to it. In opposition to this, post-New Age quantum vendors transform particles into animate agents whose unique movements and interactions with individuals will secure the health and happiness of selfinterested individual subjects. In their promotion and consumption, quantum healing and quantum enrichment programs reflect the structure and priorities of advanced capitalism.The process of quantum commodification extends increasingly to encompass areas of subjectivity—for example, spirituality—that historically have been considered immune to overt commercialization. This extension of the commodification process is evidenced in the way that quantum methodologies are commercialized and then sold to people as a means of advancing, not just their financial interests, but their spiritual well-being as well. The new economy of the atom also emerges from the late twentieth and early twenty-first century retreat from the public sphere and the attendant atrophy of the public sphere as a site of interpersonal engagement. As such, the specifically public and political nature of the earlier configurations of atomism as evidenced by those who followed Newton is supplanted by the subjective language of personal betterment and individual gain. Following the manner in which atomic behavior is cast and then deployed rhetorically is an exploration not just of matter, but also of what matters. The transition from an objective to a subjective orientation toward the material world popularizes quantum language in a way that accommodates a mode of being-in-the-world that is self-directed, depoliticized, and, arguably, narcissistic. At the same time, however, the collectivity-oriented ideology of quantum politics is emerging. In fact, the utopian political use of quantum concepts appears expressly designed to remedy the dominance of commodity-orientation subjects who are

New and Post-New Age Appropriations

209

invited to exploit quantum dynamics for their consumption of self-fulfillment. The invocation and application of quantum rhetoric in the past twenty years touches on a deep contemporary sense of being unmoored and the increasing desire for structured guidance as a means toward a renewed sense of control over one’s life.The nomadic potential of quantum metaphors and concepts ensures that, no matter what the individual or social complaint/desire, there exists a quantum strategy to ameliorate or realize it.This remarkable adaptability marks late twentieth and twenty-first century use of quantum discourse as unique, not only within the discipline of physics, but relative to all fields of scientific inquiry.

5

QUANTUM VERSUS NUCLEAR DISCOURSE

INTRODUCTION

If only I had known, I should have become a watchmaker. —Albert Einstein

I began this book with a summary of the history of quantum physics, and an explanation of quantum phenomena and concepts: the Uncertainty Principle, wave/particle duality, the observer effect, superposition, the Principle of Complementarity. I moved quickly, however, from the details of the science of quantum physics to its representational and philosophical significance. I began by looking at the challenges inherent in representing quantum phenomena in language, and the ways in which these challenges intersected—according to the founders of quantum physics— with the nature of language itself. I analyzed the relationship between quantum phenomena and visuality, in particular the debate around whether quantum phenomena could be visualized, and examined how intuition and aesthetics played out in relation to the field of quantum physics. In chapter 3 I shifted my attention to the cultural realm and literary criticism, and in chapter 4 I examined social contexts and constructs, concentrating on quantum discourse and how quantum concepts fed the rhetorical strategies that various individuals and groups used to promote their interests and agendas. In this chapter, I examine the outcome of another atomic phenomenon: the nuclear bomb. I do not delve into the science behind the bomb, because the terms of my analysis of nuclear physics do not concern the structure of the atom per se, but rather the effect that its dramatic behavior had on society at the time it was manipulated and deployed. The

212

Chapter 5

primary linkages that I make between quantum and nuclear phenomena include, but are not limited to, those aspects of the quantum that I have mentioned already: language, visuality, and rhetorical strategies. Insofar as I compare the two, the terms of my comparison between the quantum and the nuclear are cultural, social, and above all discursive, for the splitting of the atom was a societal phenomenon, and the mapping of its significance highly rhetorical. In this sense, my analysis of “nuclearism” extends the direction in which I have been heading throughout this book, one that leads toward the social and the rhetorical. The key developments in nuclear physics occurred around the same time that quantum physics was establishing itself. In 1919, on the back of Niels Bohr’s 1913 discovery of the nucleus, a New Zealand-born British physicist named Ernest Rutherford became the first person to deliberately transmute one element into another by subjecting nitrogen to alpha radiation and transforming it into oxygen.The first nuclear reaction had been achieved, and elemental transmutation went from an unreachable dream to a reality. In 1932, Rutherford’s students John Cockcroft and Ernest Walton achieved the first fully artificial nuclear reaction when they used a particle accelerator to bombard lithium with protons and separate it into alpha particles. Humans had succeeded in splitting the atom by artificial means, and this success led to an important realization: the atom was possessed of almost unimaginable power that could now be harnessed by humankind. While Rutherford was aware of the atom’s tremendous potential, the existing method of releasing atomic energy was extremely inefficient, which led Rutherford to conclude that no practical source of energy could be produced by splitting the atom. In his September 11, 1933 address to the British Association for the Advancement of Science, Rutherford observed, “anyone who looked for a source of power in the transformation of the atoms was talking moonshine.”1 In 1939, war broke out with Germany. During that same year, in the basement of Schermerhorn Hall at Columbia University, American physicist Leó Szilárd and Italian physicist Enrico Fermi conducted an experiment that produced the multiplication of neurons in uranium, proving that a chain reaction was possible. The new “philosopher’s stone”—an atomic bomb—was not only attainable, but also inevitable. Szilárd later described the event: “We turned the switch and saw the flashes. We

Quantum versus Nuclear Discourse

213

watched them for a little while and then we switched everything off and went home. … That night, there was very little doubt in my mind that the world was headed for grief.”2 That grief, it turns out, caused another chain reaction: the massive rhetorical energy spent in an effort to express, control, and manipulate the implications of the atomic bomb. This chapter is about that rhetorical reaction, which took form in the “nuclear discourse” that emerged with the first atomic bomb trial in July of 1945, and reached maturity in the 1950s. My analysis of this discourse and its rhetorical strategies is grounded in its similarities to and differences from quantum discourse—in the relationship it expresses between the material world, sense-based experience, and language; in its relationship to visuality; in its construction of the relationship between subject and object, observer and observer; and in its dominant metaphorical conceits. My comparison is premised on the belief that examining their similarities and differences results in a more robust understanding of both the particularity of these two atomic discourses, and a further understanding of the relationship between language and human experience. I argue that the fundamental connection between the advent and evolution of nuclearism and quantum concepts lies in the fact that they both address aspects of the physical world that defy description—and for the same reason: in both cases there exists no experiential point of reference. Like quantum phenomena, I propose, the implications of splitting the atom could not be apprehended via the ordinary sense perceptions upon which language rests. In quantum physics, on the one hand, this limitation emerges from the fact that the fundamental aspects of the world that we experience through the senses—causality, temporal and spatial continuity, boundedness, and self-consistency—simply do not apply. Splitting the atom, on the other hand, could not be understood via everyday experience and ordinary sense perceptions because its spectacular apotheosis in the nuclear blast overwhelmed the senses and exceeded the existing store of experiences. In this chapter I trace how, like quantum phenomena, the nuclear blast could not be narrated, and like quantum physics, those who initially witnessed and sought to represent the blast spoke openly about the limits of language in communicating what they had apprehended. In this chapter I demonstrate that, unlike quantum discourse—which initially was characterized by a restrained and, at least

214

Chapter 5

on the part of Heisenberg, even a suspicious view of language—nuclear discourse was characterized by a proliferation of it. The unfolding of the quantum world took place within a highly constrained arena, among the very small number of people who could even begin to understand, let alone generate, the complex mathematics that described it. The later use of the quantum in popular culture and literary criticism tended to be abstract, apolitical, and either directed at the individual or applied to support lone efforts at textual exegesis. Where it was politicized—for example, in quantum politics or in critiques of Western logos and dominant forms of subjectivity—the outcomes remained academic and speculative. The result of the splitting of the atom, however, was public, highly politicized, and global in its implications. Much of the rhetoric around the bomb, then, was similarly politicized and public oriented. In its most political form, this rhetoric was deployed in a controlled and directed way to manipulate and redirect public opinion, to repress American accountability for the bomb, and to affirm American technological dominance and exceptionalism. The fundamental distinction between quantum and nuclear discourse lies in the rhetoric of mastery that characterizes nuclear discourse. In this chapter, I show how the respective relationship to mastery in quantum and nuclear discourse can be traced back to the specific atomic phenomena that each undertakes to describe. In its essence, quantum physics is about giving up control over the material world. This abdication of control is forced upon us in a number of ways: in the fact that it is not possible to know conjugate states of the same system simultaneously; in superposition and the observer effect, which prevent us from knowing the state of a subatomic system prior to measurement; and in the “spooky action at a distance” of quantum entanglement. Later, the observer effect came to be associated with a participatory relationship between humans and the material world that suggested a more cooperative and mutually constituting relationship. Even when a quantum guru claims to be manipulating nature, they describe their process as merely activating our already existing quantum nature. In nuclear physics, in contrast, the atom is viewed as subject to the human will—its power measured, captured, and controlled. Rather than indirect observation of the subatomic world, nuclear physics engages in direct and aggressive action upon it, violently rending the core of the atom through unnatural means. Splitting the

Quantum versus Nuclear Discourse

215

atom, I argue, is not about participating in cosmic forces; it is about taming them. It is not about giving up control to nature; it is about wresting control from nature and bringing it under the command of humanity. If quantum phenomena reflect nature’s power to defeat our understanding, the splitting of the atom is about humanity’s willful power over nature. NUCLEAR DISCOURSE AND THE INEXPRESSIBLE

There was almost no talk I can remember on our trip back to the base. It was just too much to express in words, I guess.We were all in a kind of state of shock. —Joe Stiborik, Enola Gay radar operator

In The Counter-Memorial Impulse in Twentieth-Century English Fiction, Sarah Henstra observes that at the heart of nuclear rhetoric is the awareness that we are speechless, that there is no way of narrating the bomb, and that nuclearism thwarts chronological narrative.3 Referring specifically to the need for novel language after the detonation of the bomb, Charles Gannon similarly observes that “size makes mute any discursive attempt to establish connection between individual experience and the overwhelming total reality of a nuclear explosion.”4 Gannon concludes that, where description of the atomic bomb is concerned, “language’s capacity for signification is stunted.”5 Finally, Paul Chilton argues that “[t]he postHiroshima world has had to create new images and vocabulary to encapsulate the inconceivable—literally inconceivable—phenomenon of nuclear fission/fusion and its moral implications.”6 All of these observations echo two key and interrelated problems that were expressed by Bohr and Heisenberg in particular: there exists a disconnect between atomic phenomena and ordinary experience, and there exists no language for describing these phenomena directly. One way of understanding the difference in similarity of quantum physics’ and nuclearism’s inexpressibility is through the lens of I. A. Richards’s concept of vehicle and tenor. While “vehicle” corresponds roughly to Conceptual Metaphor Theory’s source domain, and “tenor” to its target domain, it is useful to apply Richards here for the way he considers the manner in which vehicle and tenor interact with one another to

216

Chapter 5

create metaphor. According to Richards, a key aspect of metaphor is the active interplay between vehicle and the tenor, which constitutes the “two halves” of a metaphor.7 For Richards, it is not merely the case that the vehicle ascribes meaning to an “inert” tenor; rather, a metaphor emerges in the mutually constituting activation of the similarities and dissimilarities between vehicle and tenor. He writes,“the co-presence of the vehicle and tenor results in a meaning (to be clearly distinguished from the tenor) which is not attainable without their interaction.”8 To explain the interaction between similarity and difference in metaphor, Richards introduces the simple example of a table leg. He distinguishes this metaphor from a literal description such as the leg of a horse, in the sense that the leg of a table (the tenor) possesses only some characteristics of the leg of a horse or person and lacks others. He elaborates on the relation between similarity and difference in metaphor by proposing that whether a word is being used literally or metaphorically depends on whether “it presents both a tenor and a vehicle which co-operate in an inclusive meaning. If we cannot distinguish tenor from vehicle then we may provisionally take the word to be literal; if we can distinguish at least two cooperating uses, then we have metaphor.”9 Richards identifies these common characteristics, or cooperating uses, as the essential “ground” of the metaphor.10 According to Richards, the tenor is the underlying principle subject, while the vehicle provides the means.11 The status of the tenor as underlying principle, as well as the necessity for cooperation between tenor and vehicle, provide the keys to comparing the unique predicaments expressed in the quantum interpretation and in nuclear discourse. As Elizabeth Leane observes, quantum theory implies a radical failure of metaphor in the sense that “the metaphors popularizers employ, like the equations employed by particle physicists, are ‘all vehicle and no tenor.’”12 The cause of this imbalance lies in the fact that in quantum physics there is no such thing as an identifiable subatomic “entity” that exists prior to or outside of observation—there are only discontinuities, tendencies, and probabilities. Any entity that does emerge from our act of observation has been created by that act, just as the tenor (quantum phenomena) is created by the source that defines it. Because in the quantum world, subatomic phenomena cannot be said precisely to “exist,” they cannot fulfill their metaphoric role as an underlying principle subject.This means that one half of

Quantum versus Nuclear Discourse

217

the metaphor is by its very nature absent, thus preventing the mutually constituting relationship between tenor and vehicle. All metaphorical representation of quantum phenomena will miss the mark, because the vehicle ends up performing the constitutive act of bringing into being a tenor that is not present in any discernable form. Nuclear discourse also expresses a skewed relationship between vehicle and tenor that suggests an analogous dynamic. Nuclear discourse—at least in its initial incarnation—reveals an equal but opposite failure of metaphor: it proves to be all tenor and no vehicle, such that the tenor (the atomic detonation) overwhelms the vehicle. In the metaphor “mushroom cloud,” for example, the relevant aspects of the vehicle (the shape of a mushroom, the consistency of a cloud) seem utterly impoverished and entirely unequal to the task of representing the awesome spectacle and unprecedented implications of the nuclear blast. If quantum phenomena were inexpressible because they escaped sense perception, the blast was inconceivable because it overwhelmed sense perception and threatened to leave its witnesses speechless. While the scale of the quantum is too vanishingly small to be apprehended (there is no tenor to describe), the scale of the blast—and its implications—were too large to be absorbed (no vehicle is sufficient to describe it). If, in quantum discourse, the unreality of subatomic behavior thwarted language, it was the overwhelming “reality” of that first nuclear blast that defied representation. One of the rhetorical strategies undertaken to express the overwhelming effect of the first test detonation—referred to as Trinity—and the unimaginable power that it suggested was to adopt the language of the sublime, which is used to describe an experience of such immensity that it exists beyond all calculation and measurement. More than any other modality, the language of the sublime is designed to express an inexpressible experience or, more accurately, to express the very inexpressibility of an experience. Primarily an account of the emotions surrounding an experience, the sublime is typically associated with feelings of awe, wonder, horror, and terror. In a letter penned in 1693, John Dennis introduced the terms of the sublime when he wrote to a friend giving an account of crossing the Alps. He described the journey as being at once a pleasure to the eye, and “mingled with Horrours, and sometimes almost with despair.”13 For Dennis, beauty and horror were complementary terms of the sublime experience.

218

Chapter 5

A little more than fifty years later, Edmund Burke would write a definitive account of the sublime, which he described as an experience wherein the imagination is at once moved to awe and instilled with horror. Burke stressed in particular the association of the sublime with terror; in A Philosophical Inquiry into the Origins of Our Ideas of the Sublime and the Beautiful he wrote, “Terror is in all cases whatsoever, either more openly or latently, the ruling principle of the sublime.”14 Burke’s other relevant contribution to the notion of the sublime, in this context, lies in his emphasis on its physiological effects.15 This association of the sublime with the physiological suggests that the sublime expresses itself through embodied, sensate experience. Unlike Dennis, Burke viewed the sublime, in its association with vastness and terror, as distinct from the beautiful, which he associated with qualities such as balance and color. According to Burke, “we love the beautiful as what submits to us, while we fear the sublime as what we must submit to.”16 William Laurence, the official journalist for the Manhattan Project, nuclear testing, and the detonations over Hiroshima and Nagasaki—and their preeminent public narrator—wrote of the Trinity test, “the lighting effects beggared description. … Words are inadequate tools for the job of acquainting those not present with the physical, mental and psychological effects. It had to be witnessed to be realized.”17 For Laurence, the blast could not be mediated by language; it had to be experienced directly. Using similar language, Laurence described his feelings at beholding the Hanford plutonium plants: To behold these atomic power plants standing in their primeval majesty is one of the most terrifying and awe-inspiring spectacles on the earth today.There is not a sign, not the slightest hint, that within these huge man-made blocks titanic fires are raging such as have never raged on earth in its present form. One stands before them as though beholding the realization of a vision such as Michelangelo might have had of a world yet to be, as indescribable as the Grand Canyon of Arizona, Beethoven’s ninth symphony or the presence “whose dwelling is the light of the setting suns.”18

The sublime nature of atomic power as both overwhelming and terrifying is also expressed in a comment made by General Thomas Farrell, chief of field operations at the Manhattan Project, after witnessing the Trinity blast. In his official report on the test, Farrell wrote:

Quantum versus Nuclear Discourse

219

The effect could well be called unprecedented, magnificent, beautiful, stupendous, and terrifying. … The lighting effects beggared description. The whole country was lighted by a searing light with the intensity many times that of the midday sun. It was golden, purple, violet, gray and blue. It lighted every peak, crevasse and ridge of the nearby mountain range with a clarity and beauty that cannot be described but must be imagined.”19

Awe and terror—these are precisely the terms through which the sublime is expressed. And yet, there is clearly an aesthetic element to both of these descriptions that departs from Burke’s distinction between the sublime and the beautiful. Farrell describes in detail the colors of the blast, and makes specific reference to its “clarity and beauty.” Laurence compares the “titanic fires” that produce plutonium to a premonitional vision that Michelangelo might have had, and invokes Beethoven’s Ninth Symphony to emphasize the indescribability of these raging fires. Here, references to the sublime nature of the release of atomic power return to John Dennis’s original configuration of the sublime, which combined beauty and terror. This is perhaps not surprising, since the power released by splitting the atom is both natural (a potential already within the atom) and artificial (a power that is released through human intervention). This power both submits to us (we create, control, and contain the power of the atom), and demands our submission (that power, emerging directly from nature, overwhelms us). Thus, atomic power is both beautiful and terrifying: it is a pleasure to the eye, and mingled with Horrours. RELIGIOUS AND MYTHIC RHETORIC IN NUCLEAR DISCOURSE

In these Promethean structures, that may well stand as eternal monuments to American genius and enterprise, heralding the new age of Atomic Power, as well as to the Spirit of Man Challenging Nature, mighty cosmic forces are at work. —William Lawrence

The mythic in quantum discourse was created long after the science, and was expressed in two related ways. On the one hand, it was reflected in the New Age invitation to understand the mystical that was present in the physics of the cosmos, and our place within the universe. In quantum

220

Chapter 5

physics, mythic associations emerged around Bohmian notions of a universal cosmic totality, and in the references to Eastern mysticism that fostered the notion of quantum consciousness and supported the claims of quantum gurus and distant quantum healing. On the other hand, a mythos surrounding the quantum was cultivated by benevolent quantum “savants” who sought to seduce consumers with the promise of secret knowledge. In nuclear discourse, both the mythic and the religious registers became associated with atomic energy. A divine sense of mastery, of ultimate control over the material world and the promise of immortality, is a common trope in nuclear discourse, and sets it in opposition to quantum discourse.While quantum physics later accrued mystical associations, it is fundamentally about giving up control over the material world. The bomb was not about sharing power with the cosmic forces of the universe. Instead, it was about the precise control of those forces, for only in complete mastery over them could a sustained yet delimited chain reaction be ensured. That precise control, combined with the overwhelming spectacle of the blast, opened the door to religious associations. According to Chilton, “religious vocabulary and phrasing … are typical of the way the politicians and the press spoke [after the bomb was dropped]. In religious cultures the awful and the anomalous are allied with the supernatural, and the supernatural is both dangerous and sacred. Such familiar patterns of thought and talk somehow seem to have made the bomb both conceivable and acceptable.”20 Chilton suggests that, precisely because there existed no point of reference in everyday experience from which to describe the immensity of the bomb and its implications, people were drawn to religious rhetoric. Myth, both related to and distinct from religion, provides a means to order and classify our experiences of the world according to universal and archetypal structures of human experience that tend to be consistent across cultures. Where specific references to the mythic or divine appear in nuclear discourse, these references are less about the bomb itself as sacred and supernatural, and more about how control over the power contained in the bomb elevates humans to divine or mythic status by virtue of their godlike mastery over the material forces of the universe. In his August 5, 1946 speech informing the public of the detonation of an atomic bomb over Japan, President Harry Truman said,“It is an atomic

Quantum versus Nuclear Discourse

221

bomb. It is a harnessing of the basic power of the universe.The force from which the sun draws its power has been loosed against those who brought war to the Far East.”21 Truman’s reference to harnessing the sun’s power suggests an association with a central figure in the pantheon of ancient gods.The sun god, seen as the bestower of light and life to the cosmos, the source of enlightenment, and a strict guarantor of justice, was a paramount deity in several ancient religions. In suggesting that his country had harnessed the force that powers the sun,Truman was claiming a position equal or superior to the sun god, whose power America now controlled. At the same time, by saying America had “loosed the power” of the sun against those who engaged in an “illegitimate” war, Truman established America as the “strict guarantor of justice.” Truman’s rhetorical strategy, the goal of which was to establish America’s unassailable dominance, emphasized “power over” rather than “power to.” Truman’s emphasis on the “unleashing” of the bomb’s destructive potential refigured the sun as life-giving into the sun as an instrument of righteous vengeance. Like quantum discourse, a sense of secrecy was expressed in nuclear discourse, but this time the savants were the scientists themselves. Both a mythologizing and an ambivalence about the scientists manifested itself in two recurring images: the scientist as Prometheus, the powerful Titan deity who stole fire from the Olympian gods and gave it to humans, and the scientist as an alchemist in possession of secret and dangerous knowledge. Paul Saffo describes the Trinity site as a place where “men wielded terrible knowledge, transmuting equations into energy stolen from the atomic heart of matter. For an instant, men became like Gods.” 22 The reference to scientists as having “stolen” atomic energy evokes the myth of Prometheus, which was encouraged by Arthur Compton, a leading physicist in the Manhattan Project, who specifically compared his group to Prometheus.The myth was reinforced by Life magazine’s proclamation on August 20, 1945, that the world should take comfort in the fact that “Prometheus, the subtle artificer and friend to man, is still an American citizen.”23 Some Promethean associations even cast the anthropomorphized atom itself as beneficent god, extending the strategy of transforming deadly atomic power into a life-giving companion.24 Laurence, for example, reassuringly stated that “[t]he electron is … indeed, our friend. It began serving us millions of years before man appeared on earth. It

222

Chapter 5

gives us fire and it gives us life. It was the electron bursting with sunlight that Prometheus brought down to man from Olympus.”25 While the associations of American scientists and politicians with gods suggests hubris or blasphemy, their association with Prometheus transforms them into friends of humanity, who took upon themselves a great risk in order to bestow upon humankind a great boon. Specifically Christian references accompanied mythic ones to reinforce the momentousness of the bomb’s arrival. The expression of righteous vengeance, this time in Christian terms, is evident in a quote from Churchill’s speech at Potsdam: “What was gunpowder? Trivial. What was electricity? Meaningless. This atomic bomb is … the second coming in wrath.” Churchill’s words are ambiguous: is he referring to vengeance upon the enemy, or vengeance upon humanity? Because his nation lacked bomb technology, Churchill may not have felt Truman’s sense of mastery; instead, his more ominous language suggests the arrival of the biblical apocalypse. Writing to his editor from Los Alamos after witnessing the Trinity test, Laurence professed, “The story is much bigger than I could imagine, when it breaks it will be an eighth-day wonder, a sort of Second Coming of Christ yarn.”26 Laurence’s construction is curious here: he elevates the story (that he will break) to the “eighth day wonder,” then diminishes it by calling it a “yarn.” Like Churchill, he represents the bomb’s detonation as the Second Coming of Christ; it may be that Laurence used the word “yarn” because he did not wish to go so far as to elevate himself to the level of the bomb’s prophet. A decade and a half after the Trinity detonation, Laurence recalled,“[h]ere, for the first time in history, man stands in the presence of the very act of elemental creation of matter. Here in the great silences … new elements are being born, a phenomenon that, as far as man knows, has not happened since Genesis.”27 In creating these new elements, then, the scientists’ accomplishment was equal to God’s greatest achievement: the creation of the world and everything upon it. Laurence’s romanticism and religious hyperbole is typical of his idealization of atomic power, which was enabled by the fact that his relationship to the bomb was at once intimate and removed: as the only journalist allowed to witness the blasts, he stood in close proximity to it. Because he was merely a witness and not one of the creators, however, he was free of accountability.

Quantum versus Nuclear Discourse

223

Another recurring association in the mythologizing of the scientists was the way they were cast as mysterious, powerful priests or conjurers who could work mighty transformations and who were in possession of a great and terrible power. According to Sheldon Ungar, “[i]nitially, the scientists were heroes, the demigods who inaugurated the new world. A Life reporter wrote that they had donned the ‘tunic of supermen’ and ‘stood in the spotlight of a thousand suns.’ Oppenheimer assumed a public stature that was only surpassed by that of Einstein.”28 But if initially the public saw the scientists as gods, the growing consciousness of the devastating force that they unleashed transformed them from gods into “wizards of death.”29 Scientists, it was said, “had tampered with the unknown and released a forbidden secret,” and their hubris had created an unstable world.30 Public perception was influenced by the initiation of a negative campaign against the scientists, one that had its origins in the political class. In his 1966 analysis of Congressional hearings that took place in 1946, sociologist Harry S. Hall observed: “Politicians seemed to regard scientists in much the same way that primitive men regard their magician-priests. That is to say, Congressmen perceived scientists as being in touch with a supernatural world of mysterious and awesome forces whose terrible power they alone could control.”31 This perception did not work in the scientists’ favor. The fact that scientists exclusively possessed and controlled this knowledge eventually led politicians to see them as members of a cabal from which everyone else was excluded. During the 1953 hearings on the Bureau of Standards’ rejection of the claims made for a battery additive, Senator Edward Thye exclaimed, “That is where you have always got us as a scientist because you can get into that technical field and we are left behind in a daze: we are not sure whether we dare challenge you or not.”32 Politicians also began to resent the fact that atomic scientists appeared to be in control of America’s political destiny, effacing the central role that the government and the military had in developing the bomb, and they accused scientists of having made their world insecure and unstable. Fear and resentment combined with the perception that these scientist/alchemists were of a separate class that understood little about the political and social realities of the day. In his 1945 testimony on the National Science Foundation, mechanical engineer Morris L. Cooke

224

Chapter 5

said,“[i]n the days when science was persecuted, the scientist was a recluse living as far as possible from the haunts of men. He still lives a life apart sharing almost not at all in our common activities and assuming no responsibility for the conduct of affairs outside the narrow confines of his own professional interest.”33 Politicians combined this assumption with a suspicion of the internationalism of scientific research, and concluded that scientists had little national allegiance. Hall explains: As politicians saw it, the mobility of scientists and the ease with which they could transfer from one country to another were mainly due to their having weaker sentiments of loyalty and citizenship. Scientists lacked that visceral feeling of rootedness, of being part of a land and a kinship group that true citizens had. As a consequence, they were able to leave their native land to go and work elsewhere without much feeling of regret or loss. They were never really citizens in the first place.34

Through this process of associations, scientists went from gods to witch doctors, from citizens central to America’s victory of Japan to secretive adepts who lacked allegiance to anything beyond their own work. So removed were these “alchemists” from America’s social mores and ethical commitments that they no longer even qualified as American citizens. Paul Chilton observes that “some of the scientists may indeed have been mythologizing themselves.”35 He goes on to quote Wiseman’s observation that “‘what they really were doing, and they must have been aware of it as they were doing it, was challenging the whole system of God and the whole of the Judaeo-Christian morality that up to then said certain things are prohibited by God.”36 Some scientists who had participated in the construction of the bomb—most notably Oppenheimer and Einstein—were ambivalent about violating that prohibition. Oppenheimer tended to express in specifically religious terms his guilt about his participation in creating the bomb and his fear of what might come of his creation. In a 1947 lecture at MIT, Oppenheimer confessed what he saw as the scientists’ implication in the devastation that wreaked havoc on Japan: Despite the vision and farseeing wisdom of our wartime heads of state, the physicists have felt the peculiarly intimate responsibility for suggesting, for supporting, and in the end, in large measure, for achieving the realization of

Quantum versus Nuclear Discourse

225

atomic weapons. Nor can we forget that these weapons as they were in fact used dramatized so mercilessly the inhumanity and evil of modern war. In some sort of crude sense which no vulgarity, no humor, no overstatement can quite extinguish, the physicists have known sin; and this is a knowledge which they cannot lose.37

Here, Oppenheimer implicitly compares the knowledge gained by the physicists to the knowledge that led to the expulsion of Adam and Eve from Eden and the beginning of great trials for humankind. In 1964, Oppenheimer referred to “our rather blasphemous sense of omnipotency.”38 Oppenheimer offered his darkest vision when he recalled in a 1965 documentary, The Decision to Drop the Bomb, that, upon beholding the mushroom cloud, his mind turned to the words spoken by the god Krishna in the Bhagavad Gita: “I am become Death, the destroyer of worlds.”39 Oppenheimer follows this now-famous quotation with the statement, “I suppose we all thought that, one way or another.” Oppenheimer’s face is presented in extreme close-up throughout his reflections; his head is slightly bowed, and his eyes, which appear half-closed, are looking down and away from the camera. His words are halting and spoken in monotone, as if he were struggling to get them out. The overall impression is of someone reluctantly revealing a shameful truth about himself. Guilt could also come in the form of a postwar zeal to compensate for the destructive power that the physicists’ creation had “loosed upon the world.” In his 1947 article “The Atomic Crusade and Its Social Implications,” Compton suggests that the scientists’ immersion in the Manhattan Project was analogous to Jesus’s time in the desert, the culmination of which would be an “atomic crusade” for peace. He stated: “The years they spent at making atomic bombs prepared those who were making them to burst into a vast missionary call for peace as soon as the war was won. … The whole world shall have peace and, as far as the new advances of science and technology can bring it, prosperity and a more complete life. It is this great goal that the atomists hold before them.”40 Compton’s logic here is curious. Like Laurence when he proclaimed that the development of the bomb was really about the pursuit of knowledge that would cure an ailing world, Compton constructs the building of the bomb as mere preparation for the pursuit of peace—as if the first had not made the second a pressing imperative. Compton also situates the

226

Chapter 5

missionary pursuit of peace within the framework of the medieval crusades (elsewhere he uses the term “holy wars”) undertaken to spread the Christian doctrine. His use of the term “holy war,” combined with the fact that the crusades were fundamentally a military attack on Islam, exposes the contradiction in Compton’s rhetoric, and one that reflects the tension within his own contradictory position as the author of the bomb and the messenger of peace. IMAGE AND SPECTACLE

Direct images of nuclear destruction are rare, but high frequency may be unnecessary for effect if the association is already present in the schemata of many viewers and the images are sufficiently vivid and evocative. —William Gamson and Andre Modigliani

The question of visualization in quantum physics was a fraught one, and all agreed that the phenomena described by the quantum interpretation were not easily visualizable. Where visualization does enter descriptions of quantum phenomena in the form of classical notions such as waves and particles, even Schrödinger, who argued that visualizability was a requirement for any legitimate representation of nature, saw the difficulty of reconciling it with atomic theory. The Uncertainty Principle dictated that detecting a particle in space precluded detecting its speed and direction. For this and other reasons I have detailed elsewhere, insofar as quantum theory defied efforts to locate it at a definite point in time and space, it dispensed with the concept of an entity, of a “thing.” The concept of probability further complicates the matter, because it defines particles as a tendency toward being, and thus as possessing only a tendency toward being seen. Any observation thus produces only a symbolic, mediated image, and any attempt at seeing the thing itself already leads to misdirection. Like the mental pictures from which quantum physics could not escape, the images surrounding the splitting of the atom and its culmination in the detonations over Japan were distortionary; the difference lies in the fact that these distortions were as much a product of a purposeful misdirection and intentional distraction as they were attempts to communicate the phenomenon. Gamson and Modigliani observe that, in the United States, “public awareness begins with the images of sudden,

Quantum versus Nuclear Discourse

227

enormous destruction, symbolizing the rising mushroom cloud of a nuclear bomb blast.”41 While it remains true that the most enduring and iconic visual imprint to emerge from the nuclear holocaust is the image of the mushroom cloud, the images that emerged immediately after Hiroshima and Nagasaki are strikingly detached from the blast. In the week following the bombings, neither the cover of Time nor the cover of Life carried images depicting the explosions or the resulting devastation. The August 14, 1945 cover of Time carries a simple image of the Japanese flag with an “X” through it, while the August 13 cover of Life contains a decontextualized image of a generic airplane suspended against a gray background, accompanied by the simple title,“Jet Plane.” Page thirty-four of the same issue displays a grisly set of multipanel images that show a Japanese citizen being attacked by a flamethrower and then, in stages, burning to death. While the panels portray the burning in excruciating detail, the image is notable for the fact that it depicts the use of a conventional weapon (ironically, it was displayed opposite an ad for mushrooms “fresh from the hothouse”). The issue contains no reference to the dropping of nuclear bombs. In place of the detonation, Newsweek, Time, and Life all ran a picture released by the Associated Press—not of war, destruction, or bombs, but rather of a “huge Japanese magical symbol carved in the sand at Honshu.”42 The only connection offered between the image and the atomic bomb is in the brief captions underneath. The Time caption reads: “Jap charm in the sand, a match for atomic bombs?”43 Newsweek writes, “As the bomb enters the war, the Japs leave this ancient talisman, which is supposed to ward off evil spirits, carved on a Honshu beach.”44 Life makes no mention of the bomb whatsoever with its caption: “Enemy tries to ward off attacks with drawing of good luck charm.”45 From these captions, it would be difficult to infer that atomic bombs had already been dropped on Japan; instead, they read as preemptory efforts to ward off the bomb’s use. Furthermore, as Anthony Leo observes, “the photographs implied that the weapon was … either a cataclysmic natural event or a supernatural one.”46 The image of this magical symbol does imply Japanese fear, even terror of some impending event—but it offers no human context for that fear. Because the symbol is associated with the supernatural, it also implicitly elevates above humanity those beings who have the power to instill this kind of abject terror.The images obscure the bomb as

228

Chapter 5

well as the human agency behind its use, replacing them with something much more powerful—the suggestion that those who inspire such terror possess godlike status. While the images are strikingly detached, it turns out that the editors were not intentionally censoring images of the mushroom cloud; in fact, the magazines, much more than newspapers, were feeling the pressure to provide specific pictures associated with the bombing. They were thwarted, however, by the fact that the Truman government did not make available any photographs of the nuclear blast before the magazines had to be ready to hit the stands. Neither the Nagasaki photographs nor the earlier Trinity photographs (which had remained shrouded in secrecy) were made available until August 11, when Truman announced the surrender of Japan. After the images of the detonation clouds were made available and began to circulate widely in the press, often they were not accompanied by any rhetorical context and in fact seemed to stand in place of language or more detailed and context-specific imagery. As Leo points out: “the focus of the pictorial activity on the physical event and the means of representing it combined to privilege this physical event at the expense of historical contextualization. … The photographs made over Hiroshima and those made in imitation over Nagasaki were almost entirely free of historical complexity: in fact they were free of almost any content at all.”47 With all evidence of the context erased, we are left with a detached cloud that appears as a natural, if curiously shaped, phenomenon, resonating more with a phenomenon such as a tornado than with the horrific physical effects on the ground. The shape of the mushroom cloud was simple, uncomplicated, and easily recognizable, which accounts for how quickly it achieved iconic status. Icons, however, are by nature simplifications—connotative and indirect. To understand the referential “load” of an icon, one must have some awareness of the dynamic context out of which it emerges. This context—military, political, and devastating—was initially absent, making the mushroom cloud an image devoid of content, referring to nothing but itself. The detached presentation of the mushroom cloud stood in stark contrast to prior visual coverage of the air war, which included photographs of conventional bombs being released, and aerial photographs of bombedout cities.48 In its August 20, 1945 issue, Life became the only major

Quantum versus Nuclear Discourse

229

periodical to connect the mushroom bomb to effects on the ground, with seven full-page illustrations. The cover displays a head-and-torso picture of General Tooey Spaatz, the director of the bombing of Japan, looking directly at the camera as he hauls on a cigarette.The bomb spread inside spans pages twenty-five to thirty-one, beginning with two images of the mushroom cloud. These are followed by a double-page spread of the conventional firebombing of Tokyo, and the series ends with “before” and “after” aerial views of the atomic bombing of Hiroshima.49 The before and after aerial images of Hiroshima, however, contain little visual information; looking at those images, it is difficult to discern any real difference between Hiroshima before and after the atomic bomb was dropped, and one might easily assume that the bomb had very little impact. In fact, the closer aerial images of Tokyo after the conventional firebombing reveal much more obvious devastation. Although quantum experiments provided actual pictures of, for example, wave versus particle photon behavior, visualization in quantum physics was primarily associated with the use of classical language. Classical language can be translated into mental formations, and in the case of wave/particle duality, the distortions that the classical terms introduce are a function of how they recall conventional mental formations (of waves, with their “troughs,”“undertows,” and “tides”; of particles, as separate and bounded). The distortion is produced in the mutually reinforcing relationship between classical language and conventional imagery, a relationship that ended up producing a host of associations through the process of metonymic association—both in literary criticism and in popular culture more generally. In quantum physics, then, this relationship led to a kind of creeping distortion wherein the original rhetorical contextualization of visualizable language was lost, eventually resulting in the misleading application of the concepts. In the initial representations of the atomic blast, in contrast, the immediacy of actual pictures stood in place of language, and distortions were introduced through the withholding of rhetorical context rather than the erosion of it over time. Nevertheless, the influence of conventional associations played a significant part in the obfuscation of meaning. In the case of the blast, imagery’s intuitive nature (that which led Schrödinger and Bohr to argue for the retention of the classical concepts) diverted or suppressed consideration of the bomb’s human cost. This process was

230

Chapter 5

achieved in two ways: first, by replacing the effects of the blast with images of conventional war that, while sometimes shocking, called up existing mental formations of war; second, by presenting pictures that either implied some other context (the Japanese coin in the sand) or were empty of context and thus defaulted to conventional mental formations (the mushroom cloud). All of these images obscured the human agency, and thus accountability, behind the blast. CELEBRITY NUCLEARISM

Nuclear weapons have become a screen onto which we have projected some of our deepest conflicts—as an endless stream of “escapist” books and movies shows. —Bryan C.Taylor

Existing imagery was also activated strategically within nuclear discourse in another way: by tapping into popular culture notions of heroes and glamour. The more mythic—as opposed to graphic—imagery surrounding the bomb also served to detach it from the “total reality” of its effects, and appears in its most contrived form in what I call “celebrity nuclearism.” Celebrity nuclearism dominated when the bombs were in place and ready to drop, and appeared again when the soldiers returned after the war’s conclusion. By associating the bomb with celebrity-style images, its destructive potential and American accountability for dropping it were replaced by familiar and comforting references drawn from popular culture. The type, location, and use of this imagery was tightly controlled by U.S. government military officials and its staging began with the film crews that were dispatched to record the immediate preparations for the Hiroshima bombing. Recalling his interaction with the film crews at the time that the Enola Gay was being loaded with the first atomic bomb, pilot Paul Tibbets reflected,“This was full scale Hollywood premier treatment.”50 One might venture to say that never before or since have those engaged in a military exercise been so purposefully crafted by an Administration in the image of movie stars. Several exemplary photographs played a key role in attaching celebrity status to the crew that would fly the Enola Gay over Japan. The first shows Tibbets, navigator Theodore “Dutch” Van Kirk, and bombardier

Quantum versus Nuclear Discourse

231

Tom Ferebee posing beside Enola Gay.51 All three men are looking off to their right, a pose that shows to advantage Ferebee’s rugged good looks. Ferebee has one hand in his pocket; the other is reaching up and leaning against the plane, and his slightly bent leg is resting on the wheel. Tibbets and Van Kirk have their hands in their pockets and are leaning casually against the lower exterior of the plane. The men have clearly been posed, and in such a manner that the plane appears no more than a prop. The second photograph presents Tibbets apparently waving in celebrity fashion from the cockpit of Enola Gay.52 Both of these pictures were later autographed, reflecting and reinforcing the flight crew’s status as celebrities.53 A third photograph shows Tibbets posing beside Enola Gay with his entire crew. The smiling figures appear happy and relaxed, and several crew members are dressed in shorts, as if ready for the beach. Tibbets stands in the middle, legs apart, hands in his pockets and a pipe in his mouth. Tibbets dominates the picture, and his pose, complete with pipe, cannot help but recall General Douglas MacArthur— Supreme Commander of the Southwest Pacific Arena, heroic defender of the Philippines against the Japanese, and recipient of the Medal of Honor. The visual staging of the Enola Gay crew, which cast them in a conventional Hollywood military drama—a narrative form that had been aggressively nurtured over the course of the war—was familiar and accessible. After the United States entered the war and instituted conscription in 1941, Hollywood supported the effort by mass-producing war films, with every major studio participating. With the 1941 release of Sergeant York, a biography of the most decorated American WWI soldier, actor Gary Cooper in the lead role became the prototypical image of the ideal WWII hero. In the film as in real life, the highly religious York goes from pacifist to dedicated commander, and during the Battle of the Argonne ends up single-handedly capturing a regiment of 132 German soldiers. York returned to the United States to be greeted with a tickertape parade and the offer of a key to the city. World War II aviation hero films also abounded, including: Ships with Wings (1941, starring James Clements), A Yank in the RAF (1941, starring Tyrone Power), Flying Tigers (1942, starring John Wayne), and Thirty Seconds Over Tokyo (1944, starring Spencer Tracy). All of these films depict brave and dedicated American pilots triumphing over the enemy; inserting Tibbets and his crew into this

232

Chapter 5

narrative capitalized on the patriotism, confidence, and righteousness that the films inspired. Celebrity nuclearism emerged again with the testing of two nuclear bombs at Bikini Atoll in July 1946. The Bikini tests represented the first time that the dropping of a nuclear bomb was publicly announced beforehand, and the first time it was observed by a large press corps. The first of the two bombs dropped at Bikini was christened “Gilda,” after Rita Hayworth’s notorious but ultimately tamed femme fatale character from the movie of the same name. Attached to the bomb was a pin-up photo of Hayworth that had circulated widely among soldiers during the war. Feminizing the bomb and associating it with a highly sexualized, fetishized image had the advantage of emphasizing masculine control over it, while at the same time the picture represented a pin-up that had transported thousands of soldiers surrounded by war’s violence to a less threatening and more familiar place. An iconic image captures the bomb’s historical moment, condensing the context of that moment into a single, representative form.The power of the iconic image is its ability not to invoke the historical details of that moment, but rather the associated affect. Iconic images also constrain and manipulate meaning; they are not meant to express the full historical complexity of the moment that they portray.The active creation of Hollywood iconography to frame both the characters involved in deploying the atomic bomb and the bomb itself not only directed the affect surrounding it, but also ensured that its detonation would be seen solely from America’s point of view.The perspective of the victims was overrun by the massive, patriotic Hollywood machine that provided the formal models for the images and the lens through which they would be interpreted. Through a process of decontextualization, naturalization, and redirection, the images around the bomb fulfilled the politically motivated goal to sever it from its victims and efface American accountability for their suffering. ANTHROPOMORPHISM AND THE BIRTH OF THE BOMB

And now, Ladies and Gentlemen—so that you can see what an atom is, and how it looks—the Bumsteads have volunteered to let me reduce them to the size of atoms so we can see what wonderful companions atoms are. —Mandrake the Magician

Quantum versus Nuclear Discourse

233

For both the founders and the later popularizers of the quantum concepts, the goal was to increase public engagement either by disseminating the initial theory, applying the metaphors to another field of study, or later, with the intent of profiting from quantum’s association with cutting-edge science. Any distortion of the original concepts was unintentional, and in the case of the founders of the quantum interpretation, viewed as highly problematic. In chapter 1, I discussed how the anthropomorphism of electrons humanized quantum phenomena by constructing metaphorical conceits drawn from familiar experience. I situated this rhetorical strategy in the context of Lakoff and Johnson’s observation that an enormous number of elements in our conceptual system are oriented with respect to whether or not they are similar to the properties of the prototypical person, which allows us to make sense of phenomena in the world in human terms that we can understand on the basis of our own motivations and goals. Both Schrödinger and Bohr attempted to address the more philosophical and epistemological implications of quantum physics using language that would be accessible to the layperson, and this attempt frequently led them to use humanizing metaphorical conceits that cast subatomic phenomena as familiar characters. In his discussion of the process of entropy, for example, Schrödinger contrasts the individual and the collective, with the individual particle as rogue, and the collective as statistical and indefinite. Schrödinger warns us, however, to be wary of developing a misconception regarding the individual particles, emphasizing that the elementary particle is not an individual—and he reminds us to avoid the assumption that it is merely mass “crowding” of individual particles that makes it difficult to distinguish one particle from another. According to this faulty logic, Schrödinger points out, if we could just separate the individual/particle out from the mass then we could “see” it. In general, while Schrödinger and Bohr use anthropomorphizing metaphors to make concepts meaningful, they demonstrate an acute awareness of the distortions that these metaphors inevitably import, and where they present a metaphorical conceit, they immediately put it under erasure by explaining its misleading nature. Like quantum discourse, anthropomorphism in early nuclear discourse was designed to domesticate the object and make it familiar; the difference lay in the agenda and the fact that the metaphors were not thematized or exposed as distortionary. With the bomb, the American

234

Chapter 5

government had a specifically political agenda: allaying public fear, diverting public attention away from the human cost of the atomic bomb, and preempting public resistance to its continuing development. Distortion was not only welcome, it was the whole point. It is toward this specifically political end that anthropomorphism was used to domesticate, diminish, and cast the atomic bomb in a familiar and nonthreatening form. With specific reference to nuclear rhetorical strategies, Antony Rowland argues that one of the ways in which the referent can be “silenced” is via euphemism, which “obfuscates reality by splitting the signified and referent apart.”54 Rowland offers as an example the U.S. military’s use of the term “Little Boy” to describe the bomb: “the referent is the bomb dropped on Hiroshima, compounding twelve pounds of Uranium 235 in an eight thousand pound projectile, but the signified is a small, innocent, harmless human being.”55 The absurdity at the heart of the euphemism (in this case, the image of a child being dropped from a U.S. bomber) remains latent: the signifier overwhelms the referent and succeeds in repressing its core qualities.This differs from the initial Trinity detonation, where the “total reality” of the referent overwhelmed the signifier, and left those who actually witnessed it hard-pressed to find the appropriate language to describe it. While more aggressive and elaborate diversionary rhetoric would be required in the global political arena, the story of the bomb’s creation could be refigured by associating it with the origins of life. Reflecting after the war on the elemental process through which the triggering isotope uranium is produced, Laurence wrote: The process of this elemental creation starts in the atomic pile, in which U.238 is mated with U.235. As often happens in the act of propagation of species, the male element, U.235, is destroyed in the consummation of this marriage of the elements. By its very act of death, the U.235 liberates the fertilizing agent, the neutron, [which] penetrates the U.238 nucleus. When this happens the U.238 also goes out of existence, and a new Isotope of uranium is born, containing 147, instead of 146, neutrons.56

Here, the procreative intercourse between U.238 and U.235 produces not the means of destruction, but rather the spark of life. During this process, the feminine principle dominates, while the masculine principle is sacrificed (although the masculine is implicitly resurrected in the depiction of it “penetrating” the U.238 nucleus). The mechanism of the bomb, then,

Quantum versus Nuclear Discourse

235

has as its internal logic and tendency toward the production—not the destruction—of life. Analyzing the rhetorical strategy of associating of the bomb with procreation, Glenn Hook observes: In the case of nuclear weapon development … the metaphors of birth highlight the element of human creativity and discovery involved in the process of developing the first atomic bombs, but at the expense of obfuscating the underlying moral problem concerned with death—the use of the first atomic bombs for indiscriminate killing. By structuring reality through metaphors of life, obscuring questions of death, the perspective of the victims, consciously or unconsciously, was excluded in favor of the perspective of the users of the bomb.57

Hook raises an important point: not only does the association with birth transform death into life, detaching the bomb from its reality as an instrument of mass destruction, it also skews the perspective in favor of those who had created the bomb, and effaces the victims’ point of view. Associating the bomb’s creation with the joyful experience of bringing new life into the world humanizes its creators. At the same time, it naturalizes an artificially produced process, thus further dissociating atomic power from intention and accountability. Rhetoric requires at least the perception of control over, and affective removal from, the object of representation.This sense of control had teetered dangerously for the witnesses of that first Trinity test. Of that fateful moment, Michael McNay writes: “When the scientists exploded the first atom bomb on a tower in the desert of New Mexico and the fire ball grew bigger and bigger and the hills and the sky turned white, a soldier shouted:‘The long-haired boys have lost control.’” For those who did not directly witness the blast, generating a rhetorical framework for it was less of a problem, and gender supplied this framework. If creating the elemental forces animating the bomb was associated with feminine conception, its detonation was associated with the birth of a male. Directly after the Trinity blast,Truman’s secretary of war Henry L. Stimson received a cable from George L. Harrison, his special consultant:“Doctor has just returned most enthusiastic and confident that the little boy is as husky as his big brother. The light in his eyes discernible from here to Highhold and I could have heard his screams from here to my farm.”58 While the blasts are clearly male, Harrison diminishes their masculinity by casting them as babies—their fearsome power is transposed into something that is not

236

Chapter 5

only nonthreatening, but so powerless that it requires care and nurturing in order to survive. Masculine power and infant vulnerability appear to be in contradiction, but in the context of wanting to control and maintain the secrecy surrounding the blast, the wire represents the dual agenda: to establish the bomb as male, and to diminish its “masculine” power and ensure that, should Harrison’s cable be intercepted or leaked, it would not cause alarm. Once the bombs were ready, they were further anthropomorphized through the act of naming them—“Little Boy” was the uranium bomb headed for Hiroshima, and “Fat Man” was the plutonium bomb destined for Nagasaki. Paul Chilton describes the logic behind this act of naming: “atomic and nuclear weapons were perceived as awesome and incomprehensible. A slot in our classification of reality had to be found for them, and to christen them was the first step. Rites of naming are rites of incorporation into social life.”59 It is notable that this act of christening did not take place until the Hiroshima and Nagasaki bombs were assembled and ready for mobilization. Naming is a proclamation of ownership. The bombs were about to be sent out into the world; it was important that they bear the stamp of the United States, and even be established as American citizens. At the same time, the names “Little Boy” and “Fat Man” not only familiarized the bombs as affable stereotypes, they also reflected a particularly American propensity for nicknames. Nicknames are a sign of affection and inclusion, and function to establish someone as an accepted member of a group. The bombs were one of the gang, and that gang was distinctively American. Nuclear anthropomorphism used the same strategy: we are encouraged to see the atom as just like us—in fact, to see the atom as representing the best of us in its playfulness and desire for human connection. In this sense, the methodologies and goals are the same. An early popularization of the atom featured illustrations by the then eighteen-year-old Maurice Sendak, who, at the request of his high school physics teacher, illustrated the 1946 textbook Atomics for the Millions—originally titled Atomic Adventures. One illustration featured atoms as dancing boys and girls pairing off to form molecules.60 In the picture, three shy-looking girls (sodium atoms) meet up with three boastful-looking boys (chlorine atoms). As they pair off and begin dancing a kind of exaggerated jig together, each pair becomes a molecule of salt. Here, atomic activity is

Quantum versus Nuclear Discourse

237

figured as a romantic fusion brought about by young love and good times. The product—salt—is harmless, ordinary, and essential for our bodies to survive. One year later, the Atomic Energy Commission (AEC) organized an Atomic Energy Week in Hyattsville, Maryland, a pilot project intended to forge peaceful, civilian associations with the atom. As part of the celebrations, high school students put on a play that featured such characters as Miss Molecule and Mr. Atomic Energy. That same year, John W. Campbell wrote The Atomic Story, which offers a collection of cute subatomic characters; the proton, for example, is described as a “plump, positive fellow.”61 In an effort to subdue the threatening power inside of the atom, these people associated it with the optimism and innocence of youth, and it was constructed as friendly, playful, and harmlessly endearing. In 1949, General Leslie Groves, who led the Manhattan Project during the war, developed an idea for presenting the atomic story using familiar cartoon characters. The resulting book, which was distributed at atomic exhibitions, was produced by King Features Syndicate in consultation with the AEC, and General Groves himself wrote the preface. In the preface, General Groves writes: To those who will read it carefully, this pamphlet will bring a clearer understanding of atomic energy. Many will understand what has formerly confused them. Mere words need not frighten them in the future—words such as fission, isotope, proton, chain reaction and atom bomb. This book will reassure the fearful that the future can be made bright.”62 Here, Groves suggests that fear of nuclear bombs is the product of a misunderstanding, rather than the devastating effects of its detonation. The comic rests on the initial premise that fear is a product of ignorance; once the public knows how the bomb works, it will no longer be perceived as a threat. Informing this premise is faith in the reassuring effect of a sense of control through knowledge and understanding, a type of understanding that will empty atomic power of its threatening associations. Since the most popular comic strip stars during that era were the Bumsteads, they play the leads in the story.The story begins with Blondie and Dagwood, accompanied by the rest of the King characters, attending a public lecture given by Mandrake the Magician. Mandrake tells people what atoms are made of, how they get pulled apart, and what causes them

238

Chapter 5

to detonate. To dramatize this information, Mandrake reduces the Bumsteads to atomic size so they can follow the activity of various neutrons, protons, and electrons. Once Dagwood, his family, and other King characters are reduced in size, they encounter molecules of water, which they use to quench their thirst.63 Introducing the reader to hydrogen via a water molecule (which consists of the harmless combination of two hydrogen atoms with one oxygen atom) creates a benign and life-giving context for the primary element of the nuclear bomb. After the notion of isotopes is introduced to the characters (over a series of six pages that involve much hijinks and frivolity), the narrative introduces them to the isotopes of uranium, including uranium 235—the isotope used in atomic bombs.64 Dagwood and Popeye spend some time fruitlessly trying to split the nucleus of U-235, before Mandrake gives Dagwood a bazooka loaded with a single neutron. Dagwood shoots the neutron at the nucleus and then shouts, “BLONDIE—OH BOY! I DID IT! I GOT TH’ TOUCH! IT’S SPLITTING!” He then picks up his dog, gives him a smack on the lips, and exclaims, “CAN YOU TIE THAT? I MADE A HIT WITH ONE NEUTRON, AND AFTER IT SPLIT, THERE ARE NOW THREE OF ’EM!”65 Here, the splitting of the atom that creates the chain reaction is refigured as a carny game. Dagwood’s innocent delight at winning the game makes us forget the brutality of how the real game was won—by obliterating two cities, vaporizing living beings, and causing the slow, painful death by radiation of countless others. As the chain reaction expands, Dagwood scoops up his daughter and tells her that she will have to watch from the audience. While Dagwood flees, Mandrake observes, “That was an awful big noise, eh, Dagwood? It took power to make that noise—all wasted. Now the job is to find out how that power can be used, not wasted.”66 In the middle of the panel is a scalloped-edged orange circle surrounded by a border of gray clouds, with the word “BANG” in the middle. Mandrake’s reference to an “awful big noise” empties the detonation of destructive power and reduces it to the level of a firecracker.The “bang” image in the center similarly reduces the blast by situating it in the context of cartoon fights expressed by words such as “pow!” “biff!” “wham!” Around the perimeter of the page are oval pictures designated “agriculture,” “medicine,” “transportation,” “power,” and “industry.”The final few pages enlarge these images and are

Quantum versus Nuclear Discourse

239

combined with text that explains all of the beneficial ways in which atomic power can be used. Groves chose the most accessible and depoliticized form of popular culture to empty nuclear discourse of its threatening associations, and to replace the existing keywords associated with atomic power with positive images. Secure in the fact that nothing really bad ever happens to the eternally unchanging cartoon characters, the public can assume that the processes to which they are in close proximity must be harmless.The text that runs underneath the cartoon panels is somewhat more complex, and acknowledges the force of the bomb. It quickly shifts to the future benefits of atomic power, however. At the end of the book, the reader is told by World War II presidential advisor Bernard Baruch how he or she can engage with atomic energy: What can you do about atomic energy? Well, what can you do about playgrounds, health conditions, books—what can you do about anything? You want to improve something, so you study what is wrong and you suggest solutions; you talk with local officials, you write letters, for you’ve learned that constructive effort may bring results. You can do this about atomic energy: first, find out more about it—what it is and what it is not, what it should do and what it shouldn’t. Then you can talk with other people about it and gain their interest. The full use of atomic energy for good purposes should be your goal. If you and thousands like you make atomic energy your own possession and useful tool, that goal will be won.67

The profound naïveté of this statement—that the public can have the same input about atomic energy as they do about their neighborhood playground—is characteristic of the immediate postwar years. At this point, there still existed the belief that atomic power was a public trust; only in the ensuing years did it become obvious that citizens and scientists were to be shut out of the developing military-industrial complex. Many of the motivations behind the anthropomorphism of the bomb and the atom are shared with the use of anthropomorphism in quantum discourse. Both cases reflect a desire to translate something incomprehensible and intimidating into familiar and accessible language. Both aim to humanize the object by using metaphors associated with a “typical” person. Both reflect an investment in popularizing their object in a manner that makes them approachable. The crucial difference lies in the fact that

240

Chapter 5

the anthropomorphism surrounding the splitting of the atom is expressly political: its metaphors are designed to obscure and redirect the public away from an uncomfortable and overwhelming reality, rather than to compensate for something that remains intransigently “unreal.” If quantum anthropomorphism stands in the place of a vanishing reality, nuclear discourse is designed to make that reality vanish and replace it with a convenient fantasy. DUALISTIC DISCOURSE: UTOPIA VERSUS APOCALYPSE

To every man is given the key to the gates of heaven; the same key opens the gates of hell. —Richard Feynman

The potential for a global nuclear holocaust was inexpressible as itself; it could not be rendered in descriptive, referential terms. What it would look like, how it would feel, what kind of human toll it would take— how it would be experienced—were unthinkable both in the sense that such a holocaust was beyond the horizon of the thinkable, and in the sense that, in its horror, it could not be faced. Because, as Lakoff and Johnson observe, “no metaphor can ever be comprehended or even adequately represented independently of its experiential basis,” the apocalyptic implications of the nuclear blast could only be expressed inadequately through language.68 We cannot, for example, represent a global nuclear holocaust because nothing even remotely like it has ever happened—there exists no correlative experience to ground the language. As Jacques Derrida observes, “Unlike the other wars … nuclear war has no precedent. It has never occurred, itself; it is a non-event.”69 Derrida designates the nuclear referent as the “absolute referent”—it is, he argues, pure signifier without any existing signified.70 Uniquely selfreferential, nuclear war can be discussed only in terms of itself—“the terrifying reality of the nuclear conflict can only be the signified referent, never the real referent (past or present) of a discourse or text.”71 In the absence of a referent, “‘rhetorical forms” can only refer back to other rhetorical forms, in an inescapable chain of signification. Since this is the foundation of Derrida’s general theory of language, then the nuclear holocaust becomes the ultimate expression of that theory—we have only

Quantum versus Nuclear Discourse

241

an imagined referent, a chimera that does not have even an illusory relation to the real. The notion of nuclear war as a nonevent recalls quantum phenomena, where, as I pointed out in chapter 1, one might question the extent to which the notion of an event is appropriate to quantum phenomena, given the fact that they disrupt the causality and temporal/spatial relations upon which the concept of “events” rests. Part of why quantum physics undermines the notion of event concerns the fundamental role that probabilities play. In this sense, nuclear war and quantum phenomena share the same logic: they are suspended in a state of “not yet, maybe not ever” that can only be resolved through direct observation. But direct observation of quantum phenomena is impossible—one can only describe the behavior after the fact—and direct observation of a nuclear holocaust is impossible because the moment of observation would also be the moment of obliteration, leaving no one left to describe it. Derrida’s association of nuclear war with the “absence of a referent,” where “rhetorical forms can only refer back to other rhetorical forms” recalls Christine Froula’s comparison between quantum physics and Derrida’s “trace.” In chapter 3, I summarized Froula’s description of how the trace undermines “being” as “presence,” which is replaced by a freeplay of meanings wherein the sign never leads to the extralinguistic “thing,” but rather to another sign. In the same way, there is no subatomic “thing” at the quantum level, only an irreducible uncertainty. It turns out, then, that both the nuclear holocaust and quantum phenomena are paradigmatic expressions of Derrida’s theory of the relationship between the absent referent and rhetorical form. While nuclear war remains completely unrepresentable, the unprecedented experience of and the previously unimagined release of the immense power at the heart of the atom witnessed at the initial test blast pushed language to its extremes. In this respect, the bomb exceeded all existing social, political, and cultural contexts. At the same time, however, even from before its creation, the bomb “inhere[d] in a configuration of actual political and technological circumstances along with the specific potentialities that such circumstances bear.”72 These circumstances and potentialities, which included the military use of the bomb as well as the implications and future directions of atomic power, were of particular concern where they pertained to the U.S. Administration’s standing both

242

Chapter 5

at home and abroad. From the Administration’s point of view, then, atomic power had to be managed, integrated, and domesticated on both the national and international stages. The bombing of Hiroshima and Nagasaki became part of an American techno-political matrix that was characterized by rhetorical expansion. Redirection was central to how both the bomb and its implications were figured, and efforts to deflect consideration of the brutal path that atomic power had taken were expressed not through the absence of words, but rather in the proliferation of them. One such effort took the form of transforming the splitting of the atom from dystopian present to utopian future. In By the Bomb’s Early Light: American Thought and Culture at the Dawn of the Atomic Age, Paul Boyer observes, “speculation about atomic energy’s glorious promise enabled Americans to turn from the immediate reality of its military use, and even to view that use as a necessary stage in a larger, beneficent process.”73 Consider, for example, a New York Times editorial written within days of the war’s end: In this shock that ran like an earthquake around the world, there is room for hope, room for dreams of a nobler future for mankind.The atomic bomb was perfected for war, but the knowledge which made it possible came out of … the deathless yearning to know and to use the gifts of nature for the common good. … This new knowledge can bring to this earth not death but life, not tyranny and cruelty, but divine freedom.What dazzling gifts the science which split the atom can offer the heavily laden everywhere! … What cannot this science do for the millions of China and India, bound for so many ages in sweat and hunger to the wheel of material existence.74

In the preceding passage, the creation of the bomb is defined as incidental, an exceptional circumstance that emerged from a unique moment in history.The real motivation for splitting the atom was not the devastation of Hiroshima and Nagasaki, but rather the desire for knowledge that would benefit humankind. Introducing this alternative motivation represses the fact that the massive scientific resources made available during the Manhattan Project were applied with the express purpose of creating a nuclear bomb before the enemy did. Instead, something more profound and noble lay behind this monumental effort: the deathless yearning to know and to use the gifts of nature for the common good. The destructive goal of obliteration and total surrender was transmuted into the constructive goal of saving the entire world from suffering and hunger—to

Quantum versus Nuclear Discourse

243

free them from their material bonds and bestow upon them a “divine freedom.” The bomb was also refigured by integrating its development into a utopian journey narrative. In chapter 1, I introduced the concept of a journey with Kai Mikkonen’s observation that “the notion of travel is prone to give identity and narrativity to a series of events since it ‘humanizes’ the experience of time and space.”75 I showed how Bohr, Schrödinger, and Heisenberg draw on the metaphor of journey to describe the nature of advances in theory. By presenting themselves and their work in this manner, they socialize their efforts and refigure their highly abstract and abstruse methods as a universal experience. In nuclear discourse, the journey narrative is also about the process of scientific discovery, but its ultimate goal is to idealize atomic power by associating it with a future utopia of peace and plenty—a future with which the public could not only safely identify, but that they also would actively desire. In the way that it was invoked, the journey metaphor assigned both a higher purpose and a broader significance to the research that culminated in the detonation of the bombs. In 1946, Laurence wrote: Man’s eternal quest for the philosophers’ stone and the elixir of life was actually a manifestation of his quest for means to conquer space and time. The philosophers’ stone would make him lord of the material world in space; the elixir of life was to give him mastery over time and death. After five hundred thousand to a million years of his existence on this earth, during which his quest had appeared under many metamorphoses, man at last is within striking distance of his goal.76

In this extended conceit, scientific advances are cast as part of a larger quest, and the culmination of that quest is not the bomb, but rather that greatest and most elusive of prizes: the philosopher’s stone. Laurence’s depiction of the philosopher’s stone combines both of its traditional associations: gaining ultimate control of the material world, and, through that, achieving immortality. It turns out, then, that in developing the bomb, the scientists bequeathed upon humankind, not death, but rather life. Not only was death not the point of their research—in gaining mastery of the energetic forces of the atom, their efforts meant that death could be eternally vanquished. Laurence performs his own alchemy here: death becomes life, and the possibility that humanity faced the end of time becomes total mastery over time.

244

Chapter 5

Directly after the preceding passage, Laurence offers another journey metaphor, this time less mythically and more historically resonant: “It will be remembered,” he writes, “that the whole story of fission came as the direct result of an attempt to create an element beyond uranium. It was another case in which the quest of a passage to the Indies led to the discovery of a new continent.”77 As in the passage from the New York Times, the goal of developing the atomic bomb is transformed into an accidental, if opportune, discovery—now represented as a new continent upon which scientists stumbled while seeking a passage to something else. From the postcolonial perspective of the twenty-first century, what strikes one here is the unreflective allusion to the historical process of colonization, with all its forms of brutality and cultural erasure. This unintended subtext forges an association between the obliterating force of the bomb and the obliterating process of colonization. Laurence effaces colonial history, just as he effaces the effects of the bomb on its victims, by focusing on the moment of discovery—that moment prior to the violent act of domination. His quest ends with that epiphanic moment of revelation, a moment suspended in time and severed from its inhuman entailments. If quantum physics could be reduced to one gesture in the popular imagination, it is the replacement of the logic of either/or with the logic of both/and. References to this new dynamic introduced by quantum physics is particularly prevalent with respect to the Principle of Complementarity, where subatomic phenomena behave as both particles and waves. This both/and logic has been invoked as a model for a participatory ethos to replace a dominating Western political structure based on hierarchy and mastery. It has been cited as a model for notions of race and ethnicity that challenge single racial categories and allow for the adoption of hybrid identities. It has been presented as a model for an inclusive female subjectivity that breaks down boundaries between self and other. It has been used as a model for metaphors designed to evoke a participatory universe in which human consciousness blends with a cosmic consciousness in a unifying totality, and for a mode of textuality wherein the reader, author, narrator, and characters are bound up, in a process of co-creation, with one another and with the narrative. Examining nuclear discourse, one would be hard pressed to find metaphors or images based on both/and. One might find it in visions of the

Quantum versus Nuclear Discourse

245

mushroom cloud as beautiful and horrifying, magnificent and terrible, or seek it in the representation of physicist/alchemist as both savant and sorcerer. In the end, however, a logic of either/or dominates nuclear discourse, particularly in the opposition between atomic power as the herald of peace and plenty, and atomic power as the ultimate destroyer. In the transformation of atomic energy from destructive to benign, dystopian to utopian, there exists an implicit dualism. What remains striking about much of the rhetoric of what William Gamson and Andre Modigliani call “nuclear dualism,” however, is the way in which utopia and Armageddon are so often juxtaposed within a single utterance. In “Media Discourse and Public Opinion on Nuclear Power: A Constructionist Approach,” Gamson and Modigliani argue that nuclear dualism is structured around the opposing images of unimaginable destruction and unprecedented possibility, and that they emerged simultaneously with nuclearism’s entry into the American imagination.78 Raymond Swing, for example—one of the most influential radio broadcasters of his time—offered a stark set of alternatives on August 13, 1945: “The choice before men is simple: … either live fabulously well, or … commit suicide as a race.”79 In a 1948 speech titled “The Atom: Death—or the Life Abundant?”William Foulkes wrote: You see, it is a matter of “Death—or Life Abundant.” There is no middle ground. It’s one thing, or the other. Either we control it, or we don’t. If we can’t control it, it is extermination. But if we do control it, its fantastic abundance is as real, and just as certain. It’s one or the other. America’s voice may be decisive. And it’s your voice that will decide for America. The controlled release of atomic energy is the most momentous fact ever to confront the human race. Giving man at one and the same time the power to destroy himself or to enter into a life of unparalleled abundance, it is either the ultimate, grim, sardonic jest of the gods, or the final, great material blessing of God. It’s up to you.80

Laurence, who of all the commentators on the bomb was the most utopian in his outlook, nevertheless expressed a similar dualism in some of his proclamations. In his 1946 book Dawn over Zero, Laurence wrote: For all man’s recent ills have been brought about by the misuse of the products of man’s genius. If mankind sees to it that atomic energy is properly controlled, this new cosmic fire can bring him new light and new warmth and

246

Chapter 5

new freedom. If he allows it to get out of control as was the case with man’s other inventions, it will mean a conflagration that will engulf the earth and all its inhabitants, leaving abysmal chaos and ruin.81

Two years later, in “Paradise or Doomsday?,” Laurence returned to one of his favorite themes—the philosopher’s stone—this time contrasting his utopian vision to a cataclysmic one when he wrote, “provided man manages to avert the disaster that could transform him and all his works into a cloud of atomic dust, he has in atomic energy a very philosopher’s stone. With it he will be able in the not too distant future to remold the world in his heart’s desire.”82 All of these dualistic pronouncements share a key characteristic that is not present in the images around the bomb, the domestication of atomic power, or the magnificent and terrifying image of the mushroom cloud: they foreground human agency. If everything to which I have referred up until this point has detached the bomb from American actions and accountability, dualistic discourse specifically references the people’s power to choose between “deliverance or doom.” In some cases, the choice is referred generally to be in the hands of “mankind,” but in many of these early proclamations, the choice is laid on the shoulders of the public. In his 1946 opening speech to the United Nations Atomic Energy Commission, Bernard Baruch declared: “We are here to make a choice between the quick and the dead. That is our business. Behind the black portent of the new atomic age lies a hope which, seized upon with faith, can work our salvation. If we fail, then we have damned every man to be the slave of fear. Let us not deceive ourselves: We must elect world peace or world destruction.”83 The all-or-nothing choice that typifies these pronouncements lends them rhetorical force; however, it leaves little room for those gray areas where negotiations and agreements within government bodies, between citizens and government, and between nation-states actually take place. The people are presented with a choice, but it is between two impossibilities: a new Eden, or the destruction of humankind. In the years directly following the bomb’s detonation, before the Cold War led to runaway proliferation of America’s nuclear arsenal, it had yet to become clear to the public that they would have no voice in the development and deployment of atomic power. Instead, they were encouraged to see themselves as more than citizens or voters: they were

Quantum versus Nuclear Discourse

247

representatives of all humankind. During this brief time, there was little appreciation of the mediating forces and political structures standing between the American public’s ownership of atomic power’s destiny. Also implicit in this call for public action was the erosion of the “mastery complex” that so dominated reflections on splitting the atom. After the bomb’s detonation over Japan, that former conviction that a mighty humanity had harnessed the forces of the cosmos gave way to an increasing anxiety about controlling the future of the bomb. Once that god-like sense of mastery eroded, the rhetoric became much more about how human action was going to shepherd the use of atomic power. One of the metaphors expressing the choice between paradise and doomsday was that of standing at a crossroads. Laurence, for example, wrote “[t]oday we are standing at a major crossroads. One fork of the road has a signpost inscribed with the magic word, ‘Paradise’; the other fork also has a signpost bearing the word, ‘Doomsday.’”84 In mid-1946, the United States conducted two nuclear weapon tests at Bikini Atoll—the first such tests since Trinity in July 1945, and the first detonations of nuclear devices since the atomic bombing of Nagasaki. The tests were given the code name “Operation Crossroads.” Here, however, we have a different contextualization of standing at a crossroads. No longer about a choice on the part of the American public, this was more a statement by the United States to the rest of the world: America owns the atomic bomb, and nations must either accept American calls for limitations on the development of nuclear bombs by other nations, or face the possibility of global destruction. Even in the midst of the rhetoric of public ownership, then, the U.S. government was already situating the choice in the much larger global arena. Speaking over the radio from Bikini, the prominent American political journalist and peace activist Norman Cousins observed: The real issue [at Bikini] is not whether an atomic bomb can sink a battleship, but whether the peoples of the world can prevent an atomic war. And so we have today two contrasting acts in the biggest drama of all time. … In a way these two acts seem to symbolize the choice before us. If we go one way, the way of the American [U.N.] proposals, we make a good beginning in the struggle for world law. … But if we go the other way it means that sooner or later other nations are going to have their own Bikinis.85

248

Chapter 5

While Cousins’s observation expresses the same dualism as the previous passages I have cited, it also suggests another emerging context: U.S. national interests. If Cousins begins with a general reference to preventing atomic war, his final sentence betrays his real anxiety, that “sooner or later other nations would have their own Bikinis.” Despite Cousins’s manifest goal of persuading the world to accept American UN proposals, his real point is that the only way to avoid atomic war is to preserve the United States as the sole possessor of nuclear weapons. Cousins begins by dismissing the immediate context—America’s pursuit of bigger and better nuclear bombs. In the process, he preempts consideration of American accountability for its use and for further development of the bomb. He then refocuses attention on global responsibility for averting its proliferation. Cousins implicitly introduces another aspect of nuclear dualism, one that, in its maturity, would come to represent its greatest irony. According to Cousins, the Bikini tests represented not further provocation for proliferation, but rather inspiration for the controls that would prevent proliferation. By reminding the world of the horrors of nuclear war, the spectacle of nuclear testing would hasten the acceptance of controls, and thus prevent the apocalypse. Raymond Swing was blunter about the motivation behind Operation Crossroads: “the first of the atomic era war games … is a notice served to the world that we have the power and intend to be heeded.”86 Cousins and Swing express here, in incipient fashion, the logic of deterrence wherein the United States’ ability to flex its nuclear muscles was cast, in one of the greatest and most incoherent contradictions of the century, as the ultimate guarantor of world peace. Once the question of the bomb reached the international stage, the rhetoric of deliverance or doom was replaced by the dualistic logic of “us versus them,” articulated initially around the short-lived secret of nuclear technology. In Truman’s August 6, 1945 statement informing the nation about Hiroshima, he stresses the importance of holding on to the secret of atomic technology:“under the present circumstances it is not intended to divulge the technical processes of production or all the military applications. Pending further examination of possible methods of protecting us and the rest of the world from the danger of sudden destruction.”87 In the same rhetorical move made by the alchemists of the past, Truman constructs the “recipe” for the bomb as too dangerous to be disclosed to

Quantum versus Nuclear Discourse

249

the uninitiated. In the same speech, Truman boasts that “the battle of the laboratories held fateful risks for us as well as the battles of the air, land, and sea, and we have now won the battle of the laboratories as we have won the other battles.”88 His rhetoric here introduces the logic wherein American control of the technology of the bomb is construed as evidence of American exceptionalism—an exceptionalism that came to be linked to the historical destiny of the United States.Thus, as Ungar points out, an affinity was forged between America’s technological superiority and “all of the other ways in which the United States presented itself as the superior model for the rest of humankind”—including the superiority of American liberal democracy and capitalism.89 Through a series of associations, control over nuclear technology became a cipher for the superiority of the American way of life. American control of the bomb was fleeting, however, and once it became clear that the United States would no longer be able to keep the technology of the atomic bomb a secret, a new rhetorical strategy was required to maintain control over the atom. This strategy reached its maturity in Eisenhower’s celebrated “Atoms for Peace” speech, delivered to the United Nations on December 8, 1953.90 By the time Eisenhower delivered his speech, the Soviets had already developed their own nuclear arsenal, and simply blocking proliferation was no longer a feasible strategy for maintaining American nuclear dominance. To compensate for the loss of its technological monopoly, the American administration attempted to foreclose the military option altogether by shifting attention away from proliferation and onto peacetime use of nuclear power. More than simply allowing the United States to reclaim ownership of nuclear power, this strategy allowed the United States to dodge domestic and international scrutiny of the moral and ethical implications of its own use of the bomb. Eisenhower’s speech is a study in contradiction that displays the same sort of dualism that I examined earlier. Fully half of his speech is devoted to establishing in great detail the still-overwhelming nuclear armaments dominance of the United States, as well as stressing the devastating consequences to be suffered by any nation that presumed to use a nuclear device against the United States. Eisenhower then includes a transitional paragraph that primes the audience for a shift from the militaristic rhetoric of the first half of the speech. He goes on to proclaim,

250

Chapter 5

“my country’s purpose is to help us to move out of the dark chamber of horrors into the light, to find a way by which the minds of men, the hopes of men, the souls of men everywhere, can move forward towards peace and happiness and well-being.”91 In his speech, Eisenhower at once establishes the military dominance of the United States, and suggests that military use of the bomb, which the United States no longer controls, is obsolete. In an ironic argument for America’s unique insight into the peaceful benefits of atomic power, given the fact that it was the only nation ever to have used it destructively, Eisenhower states, “The United States knows that if the fearful trend of atomic military build-up can be reversed, this greatest of destructive forces can be developed into a great boon, for the benefit of all mankind.”92 Eisenstein concludes with the heartfelt promise that the United States will “devote its entire heart and mind to finding the way by which the miraculous inventiveness of man shall not be dedicated to his death, but consecrated to his life.”93 By focusing on both war and peace, apocalyptic and utopian, Eisenhower expertly combines the assertion of U.S. military dominance and its power to threaten other nations with the anticipatory claim as leader of the peaceful exploitation of nuclear power. With this rhetorical move, Eisenhower elaborates America’s postwar strategy of seamlessly linking its own national interests with global interests, and finally with the interests of all humankind. Delivered at a moment when the Cold War had become an established fact, Eisenhower’s speech helped create an image of the United States as visionary, peace-loving, and cooperative, while the Soviet Union was cast as “obstructionist and militaristic.”94 Boyer argues that American faith expressed in the atom’s peacetime promise was “part of the process by which the nation muted its awareness of Hiroshima and Nagasaki and of even more frightening future prospects.”95 In this political context, however, it was not primarily its own awareness that needed muting; it was crucial that the United States divert the world’s attention away from the destruction that it had already wrought using atomic power. The second half of Eisenhower’s speech was thus part of a more general and strategic rhetorical about-face, wherein a peace-loving America would make the atom, as W. W. Waymack claimed in a 1947 letter, “a benevolent servant” to produce for humankind “more comforts, more leisure, better health,

Quantum versus Nuclear Discourse

251

more of real freedom [and] a much happier life.”96 Having used nuclear power for its own wartime advantage, the United States sought to close off that avenue for others by raising up the specter of a nuclear holocaust that its own use of the bomb had introduced into the global imagination. Reexternalized and detached from human agency, the bomb took on the status of a disembodied threat lacking any origin or attachment. Accountability for its prior use was supplanted by American leadership in preventing its use, and the United States went from ethically dubious nation to global benefactor whose mission was to tame the atom and produce bounty and comfort for all. CONCLUSION

In the constant interaction of its composite parts and the immense amount of energy it contains, the atom is exceptionally dynamic. Attempting to understand how the atomic components operate at the quantum level leads one down a path toward contradictions, paradoxes, and ultimately, the abandonment of our fundamental assumptions about experience, the power of language to represent the world around us, and our very relationship with the material world.Those who seek a window into quantum behavior will find themselves creating the very tendencies that they wish only to observe. Any attempt to represent them in language similarly constitutes this behavior: metaphors emerge that create “things” where there are none, and oppositions where they do not really exist. Already highly metaphorical—precisely because it is disconnected from lived experience—quantum language becomes doubly metaphorical in the hands of popularizers, leading to all sorts of slippages that compound the distortions already reflected in the original language. Quantum discourse, which even at its inception was only loosely connected to what it attempted to represent, thus proves exceptionally malleable in the hands of those who wish to use it as a scaffolding for claims about distant healing; cosmic consciousness; participatory, collective, or personal politics; and literary form. Inadvertently influencing the behavior of an atom through observation is one thing, launching an assault on it with the goal of gaining mastery over all of that energy is another. If the atom is lively, it is also potentially more deadly than humankind ever imagined.When the power

252

Chapter 5

inside the atom was initially released, the effect confounded the senses, overwhelmed lived experience, and, like quantum phenomena, challenged the boundaries of representation. No wonder, then, that people returned again and again to the indescribable nature of the atomic blast, and that they turned to the language of religion, myth, and the sublime in their attempts to articulate what they had witnessed. Nuclear discourse was malleable, too—and for the same reason: the thing itself had from the beginning been constructed through metaphor and analogy, and thus was open to reconstruction. This time, in the hands of the media and politicians instead of popularizers, subatomic particles became, variously, comic book characters, the elixir of life, harbingers of doom, and guarantors of peace and plenty. As with the various incarnations of quantum discourse, the rhetorical forms through which nuclear discourse was expressed were wedded to their historical moment and to the agendas of those who employed them. In its magnitude and utter novelty, the Trinity test inspired feelings of awe and terror that only the language of the sublime could express, and then only indirectly. Once the bomb itself was born, and nuclear discourse became both politicized and militarized, the indirect language became purposeful, and euphemism, gender coding, anthropomorphism, and stylization all functioned to allay public fear and repress blame. The fundamental character of mature nuclear discourse is its dualistic structure. It is already incipient in the religious and mythic rhetoric, which sometimes celebrates humanity’s ascension to godlike status, and sometimes condemns humans as “destroyers of worlds,” and scientists as blasphemous “wizards of death.” It reaches its maturity, however, when this dualism is expressed in the same utterance, when deliverance and doom are directly juxtaposed as two opposing choices facing humankind. Concurrent with the development of this structure is the language of choice: which way will we decide to go? Toward salvation, or toward obliteration? There is a certain impoverished, stunted nature to this rhetorical form—it lacks nuance and depth, it recalls the simplistic manner in which children see the world as black or white. The somewhat primitive nature of nuclear dualism is not a rhetorical reflection of the political dualism of East versus West, democracy versus communism that defined its historical moment. Rather, it represents a foundational logic that subsequently was reified in political dualism. Ultimately, then, the intentional

Quantum versus Nuclear Discourse

253

splitting of the atom sponsored the prevailing dualistic rhetoric, just as the interactive, inconclusive, both/and nature of quantum phenomena sponsored the rhetoric of holism and indeterminacy. From the same basic material, then, two opposing relations to the natural world emerge: participation versus mastery, humility versus aggression.What we try to do to the atomic world generates how we perceive our relationship to it and, in turn, our understanding of objectivity, being, and truth.

CONCLUSION

In the process of writing this book, I have discovered that one of the most difficult tasks has not been to understand the concepts themselves— challenging though they may be—but rather to find the best language to represent how others describe and use them. It is impossible to talk about the constituent “parts” of the quantum realm without using the terms “electron” or “proton” or “particle,” and impossible to talk about behavior within that realm without falling back on words such as “state” or “phenomenon” or “system” or “process.” And yet none of these terms, which are drawn from our experience of the macrocosmic world, do justice to the contingent and probabilistic nature of what goes on in this realm. I have found that it is the nominal bias in our language that proves most problematic when one simply wants to get on with the task, for example, of evaluating the application of quantum concepts in contexts other than physics. At times, I have elected to put the term “particle” in quotation marks, but to do this throughout with every misleading nominal term would be exceedingly distracting. Even to say “an electron is not a thing” or “a proton is not a thing” is already to mire oneself in a logical paradox; how can they not be things, when you have just named them? To say that a subatomic particle is more a “tendency toward being” or a “probability wave” than an entity is perhaps better, but one must still begin by naming it, so that the terms “tendency” and “probability,” which one wishes to foreground as the main subject, remain qualifiers of the entity “particle.” Even “field of probability” is inadequate, since, as Heisenberg points out, the probability function does not in itself represent a course of events in the course of time, but rather a tendency for events and our knowledge of events. The only solution is constantly to put

256

Conclusion

the terms under erasure even as one uses them, but this undertaking could then become the entire point, and analyzing what critics or philosophers or quantum gurus did with these terms would retreat to the background. Arthur Eddington asserted that “no familiar conception can be woven around the electron,” and that “it belongs to the waiting list.”1 After concluding that our theory of the electron amounts to the not particularly illuminating statement that “something unknown is doing we don’t know what,” Eddington observes I have read something like it elsewhere— The slithy toves Did gyre and gimble in the wabe. There is the same suggestion of activity. There is the same indefiniteness as to the nature of the activity and of what it is that is acting.2

Eddington quotes Lewis Carroll’s poem “Jabberwocky” here in an attempt to convey the elusiveness of electron behavior. As any first-year English literature student knows, however, part of the point of the poem lies in the fact that these lines remain intelligible because they retain the standard subject-verb-object grammatical structure of the English language.We have a main subject, “toves,” modified by the adjective “slithy”; we have two main verbs, “did gyre” and “did gimble”; and we have a prepositional phrase,“in the wabe,” with “wabe” as the object of the preposition “in.” The same goes for the sentence “something [subject] unknown [adjective] is doing [verb] we don’t know what [object].” I parse this “Jabberwocky” excerpt here to demonstrate how difficult it is to invent a new mode of narration appropriate to quantum physics when it is not just a matter of finding the right terms, but rather of discovering a new grammar. Eddington comes closer to a more “quantumfriendly” grammar (although it is still based on the subject-verb-object structure) when he quotes the drunken jester Trinculo from Shakespeare’s The Tempest: “This is the tune of our catch, played by the picture of nobody.”3 Trinculo, however, is not known for his clarity of speech, and it seems that the more we chase a quantum grammar, the more we have to give up intelligibility. Nevertheless, this example suggests that it may be in poetical language, more than any other form, that an

Conclusion

257

appropriate form of expression could exist, and this may explain why the critics of poetry came closest to using quantum physics effectively and accurately. But poetry does not bear the same burden of communication as scholarly writing, and can support a higher level of connotation, irresolution, and indeterminate meaning. At the same time, The Tempest’s general tendency toward self-reflexivity about its own status as a theatrical work suggests the strategy of Schrödinger, Bohr, and Heisenberg that I highlighted in chapter 1: to lay bare the limitations of classical terms even as they are forced to use them. Immersion in their continual practice of explaining the inadequacy of their own terms of expression is what enables the fullest understanding of quantum concepts and their contexts.This also explains why the other groups of literary critics to use quantum concepts successfully were those who foregrounded the problem of language. I turn now to another foundational conundrum in quantum physics, and one that is related to the “entity/nominal” problem in quantum physics: the problem of epistemology versus ontology. I have focused a great deal on the following tightly connected questions introduced by quantum physics: How do we typically speak about our world? What are the limitations of what we can say about this world? What are the limitations that constrain what we can know about this world? Finally, how are all three questions related? Quantum mathematics, which is defined as a priori knowledge derived from reason, uncovers specific limitations to our a posteriori knowledge derived from experience. In expanding our knowledge about the material world, then, quantum mathematics proves that aspects of this material world are unknowable. These insights concern an absolute lack of knowledge; there do not exist, nor can there exist, any conditions under which we could overcome our ignorance.4 The question, then, is whether this limitation is epistemological or ontological. In the philosophy of science, the distinction between reality and knowledge is understood as a distinction between the ontic and the epistemic. The indeterminism of subatomic phenomena, for example, may be explained as an existing uncertainty that is “encoded” within matter (ontic), or it may be explained as a limitation of human observation and perception (epistemic). If some aspects of the quantum world can never be known, and this unknowability is not a function of any sort

258

Conclusion

of limiting factor such as insufficient intelligence or lack of technological sophistication, then it would seem that this unknowability is encoded within its nature—that it is ontological.This unknowability concerns the fact that atomic behavior on the quantum level cannot be separated from the observing instruments that measure it. If the act of observation and the behavior of quantum phenomena are necessarily part of a single system, a system that is made up of the same substance, then their conjoined nature is an ontological matter. The ontological approach is overwhelmingly favored by New and post-New Age popularizers of quantum concepts who argue, for example, that uncertainty or entanglement are inherent qualities of the universe, and that therefore potentiality and connection rather than determinism and isolation are fundamental qualities of the universe. The ontological argument that the observer is either part of the observed or actively creates the observed has proved attractive for quantum popularizers who argue for a holistic, participatory relationship between human and world, consciousness and cosmos.These popularizers claim that consciousness stretches over and between individuals and the cosmos, and that we and everything in the universe are interconnected and mutually affecting. Seen as an ontological actuality, the observer effect can be interpreted as enabling us to apply our will to produce both physical and psychical changes. While many literary critics and political theorists also assume that quantum concepts refer to an actually existing reality, their goal is to develop interpretive models; as such, they apply the concepts metaphorically as models for analysis or action. Despite the tendency of popularizers to view the Uncertainty Principle, the Principle of Complementarity, and the observer effect as referring to ontological realities, for the founders of quantum physics, questions of ontology inevitably folded back into questions of epistemology and representation. In chapter 1, I quoted Heisenberg’s reflection on the impossibility of using existing language to describe quantum physics. In Physics and Philosophy, Heisenberg offers two postulates: communication can only take place through language; language is, by its very nature, limited to the representation of “the concepts of daily life,” to which the concepts of classical physics are tied.5 Heisenberg further points out that quantum physics cannot satisfy the demands of traditional materialist ontology, where the scientist is asked to describe what really happens. Since

Conclusion

259

the concepts of daily life do not pertain in quantum physics, and since the words “description” and “really” and “happens” can only refer to the concepts of daily life, “[t]he demand to ‘describe what happens’ in the quantum-theoretical process between two successive observations represents an irresolvable contradiction.”6 Bohr extrapolates from Heisenberg’s argument to arrive at the following conclusion: “There is no quantum world.There is only an abstract quantum physical description. It is wrong to think that the task of physics is to find out how nature is. Physics concerns what we can say about nature.”7 While Heisenberg admits that the aim of physics is still to represent nature as it really is (although he sees this as an impossibility vis a vis quantum physics), Bohr adopts a more instrumentalist approach: nature is nothing more or less than what we can say about it. For these men, questions of what we can know and how we can communicate it end up overriding questions concerning the “reality” of quantum phenomena. Most adaptations of quantum concepts do not take up the provisionality of quantum language, nor do they recognize how central the question of representation is in quantum physics. As a result, terms such as “wave/particle duality,” “complementarity,” “observer effect,” “uncertainty,” and “entanglement” tend to be interpreted as referring unproblematically to an actually existing reality. These terms, interpreted incorrectly, lead to metaphorical distortions of what are already (unrecognized) metaphorical distortions. I have already observed that their highly metaphorical nature is what accounts for the portability of quantum concepts, and what exaggerates their misapplication. Nevertheless, these misapplications are extremely instructive, for in the manner and context of their reintroduction, they reveal a host of social, political, cultural, and epistemological priorities. It is this productive misapplication that provided one of my primary motivations for exploring the many modalities and contexts in which quantum concepts are invoked. PostNew Age use of the concepts of quantum jumping and the observer effect, for example, uncover a form of hypercapitalism combined with scientism that commodifies personal healing and spiritual awakening, and capitalizes on the desire for “easy money,” at the same time that it cloaks itself in Eastern philosophies and practices. The secondary or implicit dynamics that quantum post-New Age practices reflect include a retreat into individual subjectivity, an obsession with personal

260

Conclusion

happiness and well-being, the erosion of the public sphere, and a desire for structured, empirically proven guidance as a means toward a renewed sense of control over one’s life. Any historical moment contains multiple tendencies, however, and at the same time that quantum concepts are enlisted to support the individual’s preoccupation with self-betterment, they are also enlisted to imagine a new form of politics that defines itself against liberal individualism and seeks new conceptual models for resolving social fragmentation, alienation, and cultural conflict. Often, quantum politics focuses on the observer effect as a conceptual ground for participatory democracy. Through a series of metonymic associations, political theorists argue variously that quantum politics emphasizes relational, interdependent, and collectivity-oriented engagement rather than atomic, isolated individualism; that it promotes direct democracy; and that the dynamics of social systems and the understanding of social “realities” are relative to the priorities and knowledge of the observer. The “both/and” quality of the wave/particle duality phenomenon provides the metaphor for pluralism, which is said to involve a constant, enriching negotiation between individuals or discrete groups, and larger collectivities. Some translate the Uncertainty Principle generally as a mode of self-questioning—meaning, not being so “certain” of one’s point of view—that fosters respect for and a degree of deference toward other peoples’ observations and ideas. For those who presume that the Uncertainty Principle claims that nothing is “fixed” or certain, the door is left open for new and better political formations. While I find revealing the manner in which quantum concepts are appropriated, I have also, throughout the book, been critical of their inaccurate and uninformed application and have consistently pointed out the manner in which they have been misapplied.Through repeated clarification in a number of contexts, I have tried to enhance incrementally an understanding of these concepts and the radical scientific paradigm shift that they represent. In the end, however, I have criticized their misrepresentation, not because I wish to protect the “purity” of quantum concepts (they are already impure, or compromised, once they enter into language), but rather partly in defense of a certain standard of knowledge production. Admittedly, where post-New Age quantum healing and quantum get-rich schemes are concerned, a campaign for upholding these

Conclusion

261

standards is largely irrelevant, since quantum rhetoric is simply being used as a marketing tool, and misleading or empty claims about one’s product are more or less the norm. For this reason, while I have pointed out the distortions in the quantum claims made by these practitioners, I have been equally interested in why these distortions take the form that they do, and what that says about individual and social priorities. When quantum concepts are used in the interest of advancing knowledge, however—as in the fields of quantum politics or literary criticism, or even the New Age argument that quantum physics and Eastern mysticism profess the same thing—I have devoted more time to pointing out their misuse. To take the example of literary criticism, I have identified where their use is entirely unnecessary, except as a means to dress up familiar concepts such as indeterminacy or fragmentation. I have identified where critics have left out important details about quantum concepts, associated two disparate quantum phenomena, or applied the wrong quantum concept to the wrong phenomenon. Similarly, I have critiqued generalized references to “the quantum world” in the interest of making overarching claims about, for example, postmodern literature or the role of the reader in relation to all literary texts.When they engage in these practices, I have argued, critics dilute the usefulness of quantum physics as an interpretive tool, build in inaccuracies that threaten to make incoherent their interpretations of the text, and are in danger of abandoning any nuanced sense of the distinctiveness of individual texts and the varied manner in which readers interact with them. For me, there is more at stake here than identifying instances of bad literary criticism. When literary critics promulgate misleading and inaccurate representations of science, or use it in an uninformed way, they undermine efforts in the humanities to broaden their interdisciplinary reach and engage with the growing field of science studies. As I argued in chapter 3, this kind of casual, ill-informed use of the sciences opens up the discipline to ridicule by scientists like Alan Sokal. More than this, it is to divest ourselves of the responsibility, as members of a broader intellectual community, to share informed knowledge that is presented in a rigorous, thoughtful, and genuinely productive manner. In the quantum world, there are no subatomic entities, and probabilities and uncertainties rule. In the practical world, however—that space where we manipulate matter toward certain ends—electrons, protons,

262

Conclusion

neutrons, and so on do and must exist in identifiable form. In this world, subatomic particles have definitive properties: a proton has a positive charge, an electron has a negative charge, and a neutron has no charge at all and a slightly larger mass than a proton. Elements are defined by the number of protons, electrons, bonded neutrons, and so on that they possess, and we can predict with a high degree of accuracy how an element will behave. Uranium 238 has 92 protons, 92 electrons, and 146 neutrons.8 Its isotope, Uranium 235, has 92 protons, 92 electrons, and 143 neutrons. If we direct a free neutron to split the atom of Uranium 235, it will release two or more neutrons.These neutrons in turn will split more atoms, creating an exponential increase in the number of escaping neutrons, and multiplying the number of fissions to create a self-sustaining chain reaction. Uranium 238 is not likely to blow up anything. Uranium 235 can level an entire city. Where the bomb is concerned, then, atomic processes are sufficiently macrocosmic that they can be described in relation to the concepts of everyday life.To say that atoms “split” or neutrons “escape” is a fair representation of what actually happens, although the terms have their own metaphorical load (neutrons are said to “escape,” not “leave”; they are said to “bind with” protons, not “join” them). And yet, the fundamental problem of language with respect to the atomic bomb and the origins of that problem are not unlike those attached to quantum physics. Like quantum physics, for example, that first nuclear test blast could not—at least initially—be chronicled and, like quantum physics, those who initially sought to represent it were overtly conscious about the limits of language in communicating what they had apprehended. Those who first attempted to describe the blast felt as though they could never quite capture the thing itself, and the origin of this failure was likewise in the fact that the blast could not be integrated into the “concepts of everyday life,” with their ground in ordinary sense perception. In their similarities, these two atomic discourses offer complementary insights into the limits of Indo-European language, with its origin in sense perception, visuality, and entity-based nominalism. Their similarities also suggest more universal tendencies: the tendency to express the ineffable in mystical and religious terms, the tendency to domesticate and make familiar the inexplicable by way of humanizing, anthropic narratives, and the tendency to understand and represent that

Conclusion

263

which is external to us in terms of our own embodied and sensate nature. Quantum and nuclear discourse—and the relationship between the two—are so instructive because they both occupy a “threshold” position (as both a point of origin and an outside limit) beyond which lies the inexpressible. In short, they are what one might call “threshold discourses.” In education, a threshold concept refers to knowledge that is counterintuitive or alien to the student, where learning is excursive and may lead to unexpected outcomes, and where aspects of a subject are brought together that previously appeared to be unrelated. To use the term “threshold discourse,” then, is to emphasize the conceptually alien and the utterly novel, and to suggest how various iterations of a phenomenon can combine disciplines, methodologies, worldviews, and cultural contexts in unexpected ways. At the same time, “threshold discourse” refers to the process of negotiating the borders between experience, language, and the material world. On an even more fundamental level, threshold discourse can be understood as a category defined by selfconscious references to the limits of language itself with respect to the material world, and beyond that, the absolute horizon of the knowable. Where it imposes these limits, the atom demands of us a certain degree of humility. In Possible Worlds and Other Papers, J. B. S. Haldane wrote, “the Universe is not only queerer than we suppose, but queerer than we can suppose.”9 For his part, Richard Feynman once posed the following question: if some cataclysm destroyed all scientific knowledge, and only one sentence was passed on to the next generation, which sentence would contain the most information in the fewest words? This was his answer: “all things are made of atoms—little particles that move around in perpetual motion, attracting each other when they are a little distance apart, but repelling upon being squeezed into one another.”10 In the final instance, perhaps it is the atom that provokes us to consider the sum of our knowledge, and the atom that forces us to acknowledge its limits.

NOTES

INTRODUCTION

1. The double-slit experiment that demonstrated single photon interference was performed in 1986 by Philippe Grangier, Gérard Roger, and Alain Aspect, and published as “Experimental Evidence for a Photon Anticorrelation Effect on a Beam Splitter: A New Light on Single-photon Interferences,” Europhysics Letters 1, no. 4 (1986): 17–79. 2. In his original 1927 paper, “Über de anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik,” Heisenberg used the word “Ungenauigkeit” (“indeterminacy”) to describe his basic theoretical principle. Only in the endnote did he switch to the word “Unsicherheit” (“uncertainty”). When the English translation of Heisenberg’s textbook, The Physical Principles of the Quantum Theory, was published in 1949, however, the word “uncertainty” was used throughout, and it became the commonly used term in English. 3. More sophisticated measuring devices have shown that the amount of uncertainty is less than Heisenberg predicted. Nevertheless, Heisenberg’s mathematical results remain valid, and the fundamental uncertainty resulting from the act of measurement persists. 4. Bohr invoked complementarity as a model, variously, for: wave mechanics and matrix mechanics as being equally “true”; the reciprocal nature of conjugate variables such as momentum/position; the tension between observation and the refractory effect of the experimental arrangement; quantum systems vs. classical apparatus; quantum evidence and observation vs. classical description/ terminology; spacetime acausal discontinuity vs. dynamic causal continuity. In later years Bohr came to think that complementarity was important in philosophy and many other fields: psycho-physical parallelism; the mind–body relation; the Eastern philosophy of yin and yang (the Taoist yin/yang symbol is on Bohr’s gravestone). 5. Heisenberg introduced the phrase “Copenhagen Interpretation” in a 1955 lecture titled “The Copenhagen Interpretation of Quantum Theory,”

266

Notes

subsequently published in Physics and Philosophy (1958). Before the book was released, Heisenberg privately said that he regretted using the term because it suggested that there were other legitimate interpretations, and he considered any other interpretation “nonsense.” The full quotation reads: “I avow that the term ‘Copenhagen interpretation’ is not happy since it could suggest that there are other interpretations, like Bohm assumes. We agree, of course, that the other interpretations are nonsense, and I believe that this is clear in my book, and in previous papers. Anyway, I cannot now, unfortunately, change the book since the printing began enough time ago.” Heisenberg, quoted in Olival Freire, Jr., “Science and Exile: David Bohm, the Hot Times of the Cold War, and His Struggle for a New Interpretation of Quantum Mechanics,” Historical Studies on the Physical and Biological Sciences 36, no. 1 (2005): 31–35. 6. Both the wavefunction collapse and the Uncertainty Principle concern interference by an observing mechanism; however, there is an important distinction between the two. The Uncertainty Principle concerns how accurately we can measure the state of a subatomic particle, while the wavefunction collapse concerns the act of “altering” the nature of subatomic “particles.” Only the wavefunction collapse is properly associated with the observer effect. 7. More specifically, the Schrödinger’s cat thought experiment consisted of the following: suppose we put a cat in a box with a nuclear sample, a Geiger counter, and an apparatus that would result in poison being released to kill the cat if the Geiger counter detected any nuclear decay.The half-life of this particular nuclear sample is such that there is a 50 percent chance of decay occurring in an hour. If states really exist in superpositions until an observation is made to collapse the wavefunction, then the sample is both decayed and undecayed after that hour. The Geiger counter is both triggered and untriggered, the poison is both released and unreleased, and the cat is both alive and dead. This state of superposition lasts until an observer opens the box and checks to see whether the cat is dead. 8. See Hugh Everett, “‘Relative State’ Formulation of Quantum Mechanics,” Review of Modern Physics 20, no. 3 (1957): 454–462. 9. For an exploration of the Many-Worlds Interpretation, see The Many Worlds Interpretation of Quantum Mechanics, ed. Bryce DeWitt and Neill Graham (Princeton: Princeton University Press, 1973). 10. For the full article, see Alfred Einstein, Boris Podolsky, and Nathan Rosen, “Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?” Physical Review 47, no. 10 (1935): 777–780. 11. Like “wave,” “particle,” and “observer,” the term “spin” is misleading, and does not refer to our conventional notion of something “spinning” on an axis. 12. Einstein remained throughout his career critical of the concept of quantum nonlocality because it disrupted classic notions of causality, and meant that

Notes

267

information was traveling faster than the speed of light, which is a violation of relativity. He famously called the phenomenon of nonlocality “spooky action at a distance.” 13. The sources that I draw from to discuss those aspects of Conceptual Metaphor Theory and that I find relevant to my argument include George Lakoff, Women, Fire, and Dangerous Things: What Categories Reveal About the Mind (Chicago: University of Chicago Press, 1987); Lakoff, “The Contemporary Theory of Metaphor,” in The Cognitive Linguistics Reader, ed. Vyvyan Evans, Benjamin Bergen, and Jörg Zinken, 264–316 (London: Equinox Publishing, 2007); George Lakoff and Mark Johnson, Metaphors We Live By (Chicago: University of Chicago Press, 1980); Mark Johnson, The Body in the Mind:The Bodily Basis of Meaning, Imagination, and Reason (Chicago: The University of Chicago Press, 1987); George Lakoff and Mark Turner, More Than Cool Reason: A Field Guide to Poetic Metaphor (Chicago: University of Chicago Press, 1989); Hanna Pulaczewska, Aspects of Metaphor in Experience: Examples and Case Studies (Berlin: Walter de Gruyter Press, 1999); Liliane Papin, “This Is Not a Universe: Metaphor, Language, and Representation,” Publications of the Modern Language Association of America 107, no. 5 (1992): 1253–1265;Vyvyan Evans, Benjamin Bergen, and Jörg Zinken, eds., The Cognitive Linguistics Reader (London: Equinox Publishing, 2007); Kai Mikkonen, “The ‘Narrative Is Travel’ Metaphor: Between Spatial Sequence and Open Consequence,” Narrative 15, no. 3 (2007): 286–305; Theodore Brown, Making Truth: Metaphor in Science (Champaign: University of Illinois Press, 2008); Lina Hellsten, “Popular Metaphors of Bioscience: Bridges over Time?,” Configurations 16, no. 1 (2008): 11–32; and Penny Tompkins and James Lawley, “Embodied Schema: The Basis of Embodied Cognition,” notes first presented at The Developing Group, June 6, 2009, accessed December 14, 2015, http://www.cleanlanguage.co.uk/articles/articles/245/1/Embodied -Schema-The-basis-of-Embodied-Cognition. CHAPTER 1

1. Charles Bazerman, “Emerging Perspectives on the Many Dimensions of Scientific Discourse,” in Reading Science: Critical and Functional Perspectives on Discourses of Science, ed. J. R. Martin and Robert Vee1 (London: Routledge, 1998), 15. 2. Ibid. 3. James Edie, Speaking and Meaning:The Phenomenology of Language (Bloomington: Indiana University Press, 1976), 162–163. 4. Liliane Papin, “This Is Not a Universe: Metaphor, Language, and Representation,” Publications of the Modern Language Association of America 107, no. 5 (1992): 1253.

268

Notes

5. Theodore Brown, Making Truth: Metaphor in Science (Champaign: University of Illinois Press, 2008), 41–42. 6. Lina Hellsten, “Popular Metaphors of Bioscience: Bridges Over Time?” Configurations 16, no. 1 (2008): 14. 7. George Lakoff and Mark Johnson, Metaphors We Live By (Chicago: University of Chicago Press, 1980), 18–19. 8. Jörg Zinken, “Introduction to Part Four: Metaphor, Metonymy, and Blending,” in The Cognitive Linguistics Reader, ed.Vyvyan Evans, Benjamin Bergen, and Jörg Zinken (London: Equinox, 2007), 264. 9. Lakoff and Johnson, xiv, xvi. 10. Mark Johnson, The Body in the Mind:The Bodily Basis of Meaning, Imagination, and Reason (Chicago:The University of Chicago Press, 1987), 102. 11. Ibid., 29. 12. Tayebeh Asgari, “The Study of Image Schemas in Hafez Poems: Cognitive Perspective,” International Journal of Language and Linguistics 1, no. 4 (2013): 183. 13. Niels Bohr, The Philosophical Writings of Niels Bohr, vol. I: Atomic Theory and the Description of Nature (Woodbridge, CT: Ox Bow Press, 1987), 5. 14. Niels Bohr, “Natural Philosophy and Human Cultures,” Nature 143 (1939): 269. 15. Quoted in Arthur I. Miller, “Visualization Lost and Regained: The Genesis of the Quantum Theory in the Period 1913–1927,” in On Aesthetics in Science, ed. Judith Wechsler (Boston: Birkhäuser, 1988), 93. 16. Vyvyan Evans, A Glossary of Cognitive Linguistics (Edinburgh: Edinburgh University Press, 2007), 106. 17. Lakoff and Johnson, Metaphors We Live By, 25. 18. Johnson, The Body in the Mind, 102. 19. Isaac Newton, The Mathematical Principles of Natural Philosophy Book 3 (London: H. D. Symonds, 1803), 161–162. 20. Niels Bohr, “Causality and Complementarity,” Philosophy of Science 4, no. 3 (1937): 293. 21. In Erwin Schrödinger, Science, Theory, and Man (Crow’s Nest, AU: G. Allen and Unwin, 1957), 216. 22. Ibid., 71. 23. Ibid., 91. 24. Werner Heisenberg, Physics and Philosophy: The Revolution in Modern Science (Amherst, NY: Prometheus Books, 1999), 84.

Notes

269

25. George Lakoff, Women, Fire, and Dangerous Things: What Categories Reveal About the Mind (Chicago: University of Chicago Press, 1987), 55. 26. Ibid., 54. 27. Ibid., 55. 28. Penny Tompkins and James Lawley, “Embodied Schema: The Basis of Embodied Cognition,” notes first presented at The Developing Group, June 6, 2009, accessed November 30, 2016, http://www.cleanlanguage.co.uk/articles/ articles/245/1/Embodied-Schema-The-basis-of-Embodied-Cognition/Page1 .html. 29. Johnson, The Body in the Mind, 113–114. 30. Mark Johnson discusses the STATES ARE LOCATIONS metaphor in The Body in the Mind, 114. 31. Heisenberg gives a detailed account of this visit and the “Copenhagen debates” in his 1971 collection Physics and Beyond: Encounters and Conversations (London: George Allen and Unwin Ltd.), which contains transcripts of a number of conversations among leading atomic scientists of the time, including some in which he participated. 32. Erwin Schrödinger quoted in Heisenberg, ibid., 73. 33. Schrödinger’s most detailed critique of quantum jumps can be found in “Are There Quantum Jumps? Part I,” British Journal for the Philosophy of Science 3, no. 10 (1952): 109–123; and “Are There Quantum Jumps? Part II,” British Journal for the Philosophy of Science 3, no. 10 (1952): 233–242. 34. Erwin Schrödinger, “Quantisation as a Problem of Proper Values: Part 1,” in Collected Papers on Wave Mechanics, ed. W. M. Deans and J. F. Shearer, trans. J. F. Shearer (Blackie and Son: London, 1927), 10–11. 35. Werner Heisenberg quoted in Miller, “Visualization Lost and Regained,” 93. 36. Papin, “This Is Not a Universe,” 1256. 37. Ibid., 1255. 38. Ibid. 39. Ibid., 1254. 40. Ibid., 1255. 41. Ibid. 42. Heisenberg, Physics and Philosophy, 46. 43. Ibid., 41. 44. Ibid., 144–145. Contradictio in adjecto refers to a contradiction between a noun and its modifying adjective (a commonly cited example is the phrase

270

Notes

“there was a deafening silence in the room”). Heisenberg’s use of the term here is idiosyncratic, since he is positioning “describe what really happens” as the modifying adjective of the “quantum-theoretical process.” The point stands, however: there is an inescapable contradiction within the phrase “describe what really happens in a quantum-theoretical process.” 45. Heisenberg to Pauli, 8 June, 1926. 46. Hanna Pulaczewska, Aspects of Metaphor in Experience: Examples and Case Studies (Berlin:Walter de Gruyter Press, 1999), 201. 47. Schrödinger, Science,Theory, and Man, 202. 48. Pulaczewska, Aspects of Metaphor in Experience, 118. 49. Ibid., 119. 50. Niels Bohr, The Philosophical Writings of Niels Bohr, vol. I: Atomic Theory and the Description of Nature (Woodbridge, CT: Ox Bow Press, 1987), 34–35. 51. Werner Heisenberg, “Development of Concepts in Quantum Theory,” in The Physicists’ Conception of Nature, ed. Jagdish Mehra (Boston: Reidel, 1973), 269. 52. Heisenberg, Physics and Philosophy, 185. 53. Interview of Werner Heisenberg by Thomas S. Kuhn and John Heilbron, Febrary 27, 1963, Niels Bohr Library & Archives, American Institute of Physics, College Park, Maryland, accessed January 19, 2016, https://www.aip.org/ history-programs/oral-histories/search?search_api_views_fulltext=heisenberg. 54. Ibid., February 22, 1963. 55. Werner Heisenberg, The Physical Principles of the Quantum Theory, trans. Carl Eckhart and Frank C. Hoyt (Mineola, NY: Dover Publications, 1949), 11. 56. Interview of Heisenberg by Kuhn and Heilbron, February 27, 1963. 57. Kristian Camilleri, “Heisenberg and the Transformation of Kantian Philosophy,” International Studies in the Philosophy of Science 19, no. 3 (2005): 285. 58. Bohr, “Causality and Complementarity,” 293. 59. Ibid. 60. Niels Bohr, “Discussions with Einstein on Epistemological Problems in Atomic Physics,” 1949, accessed January 25, 2016, https://www.marxists.org/ reference/subject/philosophy/works/dk/bohr.htm. 61. Niels Bohr letter to Erwin Schrödinger, October 26, 1935, quoted in Walter Moore, Schrödinger: Life and Thought (Cambridge, UK: Cambridge University Press,) 113. 62. Bohr, “Discussions with Einstein.” 63. Bohr, “Causality and Complementarity,” 293.

Notes

271

64. Heisenberg, Physics and Philosophy, 168. 65. Ibid., 169. 66. Ibid., 179. 67. Bernard Pullman, The Atom in the History of Human Thought (Oxford: Oxford University Press, 1998), 298. 68. Lakoff and Johnson, Metaphors We Live By, 132. 69. Ibid., 33. 70. Schrödinger, Science,Theory, and Man, 170. 71. Ibid., 172. 72. Ibid., 215. 73. Ibid., 216. 74. Ibid., 42. 75. Ibid., 222. 76. Ibid., 194. 77. Ibid., 222–223. 78. In Bohr, The Philosophical Writings of Niels Bohr, vol. 1: Atomic Theory and the Description of Nature, 66. 79. Ibid., 19. 80. Ibid., 4. 81. George Lakoff, “The Contemporary Theory of Metaphor,” in The Cognitive Linguistics Reader, eds. Vyvyan Evans, Benjamin Bergen, and Jörg Zinken (London: Equinox Publishing, 2007), 282. 82. Brown, Making Truth, 36. 83. Kai Mikkonen, “The ‘Narrative is Travel’ Metaphor: Between Spatial Sequence and Open Consequence,” Narrative 15, no. 3 (2007), 287. 84. Ibid., 286. 85. Ibid., 289. 86. Bohr, The Philosophical Writings of Niels Bohr, vol. 1: Atomic Theory and the Description of Nature, 52–53. 87. Schrödinger, Science,Theory, and Man, 37. 88. Schrödinger, “Are There Quantum Jumps? Part I,” 109–110. 89. Schrödinger, Science,Theory, and Man, 32, 34. 90. Heisenberg, Physics and Beyond, 70. 91. Ibid., 108.

272

Notes

92. Werner Heisenberg, “The Representation of Nature in Contemporary Physics,” Daedalus 87, no. 3 (1958): 104. 93. Bohr, The Philosophical Writings of Niels Bohr, vol. 1: Atomic Theory and the Description of Nature, 101. 94. Ibid., 2. 95. Ibid., 92, 101. 96. Heisenberg, Physics and Philosophy, 31. 97. Heisenberg Physics and Beyond, 80–81. 98. Schrödinger, Science,Theory, and Man, 68. 99. Niels Bohr, “Maxwell and Modern Theoretical Physics,” Nature 128, no. 3234 (1931): 691–692. 100. Niels Bohr, The Philosophical Writings of Niels Bohr, vol. III: Essays 1958– 1962 on Atomic Physics and Human Knowledge (Woodbridge, CT: Ox Bow Press 1987), 12. 101. Ibid, 82. 102. Ibid., 88. 103. Bohr, “Discussions with Einstein.” CHAPTER 2

1. Max Jammer, The Conceptual Development of Quantum Mechanics (New York: McGraw-Hall Book Company, 1966), 166–167, 180. 2. Paul Forman,“Weimar Culture, Causality and Quantum Theory, 1918–27,” in Weimar Culture and Quantum Mechanics: Selected Papers by Paul Forman and Contemporary Perspectives on the Forman Thesis, ed. Cathryn Carson, Alexei Kojevnikov, and Helmuth Trischler (London: Imperial College Press, 2015), 196. 3. Ibid., 194–195. 4. Olival Freire, Jr., The Quantum Dissidents: Rebuilding the Foundations of Quantum Mechanics (1950–1990) (Berlin: Springer, 2015), 19. The Fifth Solvay International Conference on Electrons and Photons held in Brussels in October 1927 offers an example of the heated debated around the new quantum interpretation. 5. Ibid., 21. In “Who Invented the ‘Copenhagen Interpretation’? A Study in Mythology,” Philosophy of Science 71 (2004): 669–682, Don Howard takes aim at the notion that there was anything like a unified “Copenhagen interpretation,” pointing out that Heisenberg created the moniker much later, in 1955, and that it was under this misleading moniker that Heisenberg combined his own form of “subjectivism” with Bohr’s more objectivist theory of complementar-

Notes

273

ity. Howard’s main target is this subjectivism, which he says opened the door to later fallacious and “silly” notions about a knowing observer whose consciousness affects external reality. While many of the later interpretations of Heisenberg’s Uncertainty Principle do distort it by imagining the observer as a human agent, Heisenberg himself was careful to insist that he was talking about the impact of the measuring instrument, and that there was no subjective element at play in quantum indeterminacy. I believe that the term “Copenhagen-Göttingen” group or interpretation more accurately captures combined efforts that went into creating the quantum interpretation; I also believe that it is inaccurate to ascribe to Heisenberg the introduction of subjectivity and consciousness into quantum theory, when he so clearly rejected the notion that subjectivity had any role in quantum effects. 6. Erwin Schrödinger, “An Undulatory Theory of the Mechanics of Atoms and Molecules,” Physical Review 28, no. 6 (1926): 1049–1070. 7. Arthur I. Miller, “Visualization Lost and Regained:The Genesis of the Quantum Theory in the Period 1913–1927.” In On Aesthetics in Science, ed. Judith Wechsler, 73–102 (Boston: Birkhäuser, 1988), 73. 8. Werner Heisenberg expressed his conundrum in a letter to Wolfgang Pauli, 8 June, 1926, in Scientific Correspondence with Bohr, Einstein, Heisenberg, ed. Alan Herman, Karl Von Meyan, and Victor F. Weisskopf (Berlin: Springer, 1985), 26–35. 9. Edward M. MacKinnon, Scientific Explanation and Atomic Physics (Chicago:The University of Chicago. Press, 1982), 246–247. 10. Henk de Regt, “Erwin Schrödinger, Anschaulechkeit, and Quantum Theory,” Studies in the History and Philosophy of Modern Physics 28, no. 4 (1997): 463. 11. For an account of this reception within the physics community, see ibid., 471; see also Mara Beller, “Matrix Theory before Schrödinger: Philosophy, Problems, Consequences,” Isis 74, no. 4 (1983): 470. 12. Wolfgang Pauli to Niels Bohr, December 12, 1924, accessed December 30, 2016, http://cds.cern.ch/search?cc=Pauli+Letter+Collection&ln=fr&jrec=1 1&p=1924. 13. Paul Forman,“Kausalitiit, Anschaulichkeit, and Individualitiit, or How Cultural Values Prescribed the Character and the Lessons Ascribed to Quantum Mechanics,” in Society and Knowledge, ed. N. Stehr and V. Meja (New Brunswick:Transaction Books, 1984), 336. 14. Arthur I. Miller, Imagery in Scientific Thought: Creating 20th Century Physics (Boston: Birkhäuser, 1984), 128. 15. Miller, “Visualization Lost and Regained,” 74, 93.

274

Notes

16. Ibid., 96. 17. Ibid., 93. 18. Henk de Regt, “Spacetime Visualisation and the Intelligibility of Physical Theories,” Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 32, no. 2 (2001): 243–265. 19. Walter Moore, Schrödinger: Life and Thought (Cambridge: Cambridge University Press, 1992), 228. 20. Ibid. 21. Frances Galton, Inquiries into Human Faculty and Its Development, ed. Gavan Tredoux (2001), 60, accessed November 16, 2016, http://galton.org/books/ human-faculty/text/human-faculty.pdf. 22. Ibid., 70. 23. See Erwin Schrödinger, “On the Relation Between the Quantum Mechanics of Heisenberg, Born, and Jordan, and that of Schrödinger,” in Collected Papers on Wave Mechanics, ed. W. M. Deans and J. F. Shearer, trans. J. F. Shearer, 45–61. London: Blackie and Son, 1928. 24. Beller, “Matrix Theory Before Schrödinger, 470–471. 25. Manjit Kumar, Quantum: Einstein, Bohr and the Great Debate About the Nature of Reality (New York:W.W. Norton and Company, 2008), 193. 26. Mara Beller, “Matrix Theory Before Schrödinger, 471. 27. Ibid., 472. 28. Ibid., 471. 29. Ibid., 474. 30. Erwin Schrödinger, “Quantisation as a Problem of Proper Values: Part 1,” in Collected Papers on Wave Mechanics, ed. W. M. Deans and J. F. Shearer, trans. J. F. Shearer (Blackie and Son: London, 1927), 1–12. Originally published as “Quantisierung als Eigenwertproblem (Erste Metteilung),” in Annalen der Physik 384 (4) (January 1926): 361–376. 31. Schrödinger, “On the Relation Between the Quantum Mechanics of Heisenberg, Born, and Jordan, and that of Schrödinger,” 46. 32. Ibid., 45. 33. Ibid., 46. 34. Interview of Werner Heisenberg by Thomas S. Kuhn and John Heilbron, February 22, 1963, Niels Bohr Library & Archives, American Institute of Physics, College Park, Maryland, accessed January 19, 2016, https://www.aip.org/ history-programs/oral-histories/search?search_api_views_fulltext=heisenberg.

Notes

275

35. Werner Heisenberg to Wolfgang Pauli, 8 June, 1926, in Scientific Correspondence with Bohr, Einstein, Heisenberg, ed. Alan Herman, Karl Von Meyan, and Victor F.Weisskopf (Berlin: Springer, 1985), 26–35. 36. Max Born, “On the Quantum Mechanics of Collision,” in Quantum Theory and Measurement, ed. J. A.Wheeler and W. H. Zurek (Princeton: Princeton University Press, 1983): 54. 37. Ibid., 52. 38. Miller, “Visualization Lost and Regained,” 76. 39. Ibid., 474. 40. Quoted in Arwin Hermann, “Erwin Schrodinger,” in Dictionary of Scientific Biography, vol.12, ed. Charles Coulston Gillispie (New York: Charles Scribner & Sons, 1975), 221. “The Significance of Wave Mechanics” was first published as “La signification de la mecanique ondulatoire” in Louis de Broglie, Physicien et Penseur, ed. André George (Paris: Albin Michel, 1953), 3. 41. Erwin Schrödinger, “Are There Quantum Jumps? Part II,” British Journal for the Philosophy of Science 3, no. 10 (1952): 240. 42. Ibid., 242. 43. Ibid. 44. Interview of Heisenberg by Kuhn and Heilbron, February 22, 1963. 45. Arthur I. Miller, Imagery in Scientific Thought: Creating 20th Century Physics (Boston: Birkhäuser, 1984), 147. 46. Ibid., 75. 47. Schrödinger, “Are There Quantum Jumps? Part II,” 242. 48. Beller, “Matrix Theory Before Schrödinger,” 481. 49. Werner Heisenberg to Wolfgang Pauli, 23 February, 1927, in Scientific Correspondence with Bohr, Einstein, Heisenberg, ed. Alan Herman, Karl Von Meyan, and Victor F.Weisskopf (Berlin: Springer, 1985), 26–35. 50. Beller, “Matrix Theory Before Schrödinger,” 475. 51. Werner Heisenberg, Physics and Philosophy: The Revolution in Modern Science (Amherst, NY: Prometheus Books, 1999), 84. 52. Niels Bohr, “Atomic Theory and Mechanics,” in The Philosophical Writings of Niels Bohr, vol. I: Atomic Theory and the Description of Nature (Woodbridge, CT: Ox Bow Press, 1987), 34–35. 53. Niels Bohr quoted in Werner Heisenberg, Physics and Beyond (London: George Allen and Unwin Ltd., 1971), 74. 54. Niels Bohr, “The Atomic Theory and the Fundamental Principles Underlying the Description of Nature,” in The Philosophical Writings of Niels Bohr, vol. I:

276

Notes

Atomic Theory and the Description of Nature (Woodbridge, CT: Ox Bow Press, 1987), 114. 55. Niels Bohr, “The Quantum Postulate and the Recent Development of Atomic Theory.” Supplement to Nature 21, no. 3050 (1928): 580. 56. Dugald R. Murdoch, Niels Bohr’s Philosophy of Physics (Cambridge, UK: Cambridge University Press, 1989), 221. 57. Ibid., 584. 58. In Niels Bohr, The Philosophical Writings of Niels Bohr, vol. III.: Essays 1958– 1962 on Atomic Physics and Human Knowledge (Woodbridge, CT: Ox Bow Press, 1987), 3. 59. Ibid., 290. 60. Miller, “Visualization Lost and Regained,” 88; Hendrik Lorentz, quoted in Miller, Imagery in Scientific Thought, 144. 61. Niels Bohr quoted in Miller, “Visualization Lost and Regained,” 89. 62. Erwin Schrödinger, Science, Theory, and Man (Crow’s Nest, AU: G. Allen and Unwin, 1957), 161. 63. Ibid., 204. 64. “What Does Anschauung Mean?” The Monist 2, no. 4 (1892): 530. 65. Ibid., 130. 66. Miller, “Visualization Lost and Regained,” 73. 67. Schrödinger, “On the Relation Between the Quantum Mechanics of Heisenberg, Born, and Jordan, and that of Schrodinger,” 45. 68. Ibid., 46–57. 69. Ibid., 57. 70. De Regt, “Schrödinger, Anschaulichkeit, and Quantum Theory,” 470–471. 71. Ibid., 59. 72. Schrödinger, “On the Relation Between the Quantum Mechanics of Heisenberg, Born, and Jordan, and that of Schrödinger,” 59. 73. Ibid., 59. For the original German see Erwin Schrödinger, “Über das verhältnis der Heisenberg-Born-Jordanschen quantenmechanik zu der meinem,” Annalen der Physik 384, no. 8 (1926): 752. 74. Ibid., 58. 75. Ibid. 76. Ibid. Schrödinger uses the phrase anschaulichkeit fröhnende in “Über das verhältnis,” 751. Here, anschaulichkeit is translated as “intuition.”

Notes

277

77. Michael Cuffaro, “Studies in History and Philosophy of Science Part B,” Studies in History and Philosophy of Modern Physics 41, no. 4 (2010): 331. 78. Bohr, “The Quantum Postulate,” 581. 79. Niels Bohr, “Causality and Complementarity,” Philosophy of Science 4, no. 3 (1937): 293. 80. Bohr, “The Quantum Postulate,” 580. As I stated earlier with reference to Bohr’s use of the term, “definition” here should be understood not in the sense of giving meaning to, but rather in the sense of clearly outlining the situation or phenomenon. 81. Bohr, “The Quantum Postulate,” 590. 82. Bohr, “Causality and Complementarity,” 293–294. 83. Bohr, “Light and Life,” 423. 84. Ibid., 364. 85. Arnold Sommerfeld, quoted in Forman, “Kausalitit, Anschaulichkeit and Individualitit, or How Cultural Values Prescribed the Character and the Lessons Ascribed to Quantum Mechanics,” 209. 86. In several other English translations, the original “anschaulich” in the title is translated as “intuitive” rather than “actual.” 87. Werner Heisenberg, “The Actual Content of Quantum Theoretical Kinematics and Mechanics,” National Aeronautics and Space Administration, translator of original 1927 German publication unknown, Washington, DC, 1983, accessed January 19, 2016, http://ntrs.nasa.gov/archive/nasa/casi.ntrs .nasa.gov/19840008978.pdf. 88. Ibid., 1–2. For the original German publication, see Werner Heisenberg, “Über den anschaulichen inhalt der quantentheoretischen kinematik und mechanik,” Zeitschrift fur Physik 43, no. 3–4 (1926): 172. 89. Ibid., 22. 90. Ibid., 30. 91. Interview of Heisenberg by Kuhn and Heilbron, February 19, 1963. 92. Interview of Heisenberg by Kuhn and Heilbron, February 25, 1963. 93. Werner Heisenberg, transcript of audiotape titled “The Development of the Uncertainty Principle,” produced by Spring Green Multimedia in the UniConcept Scientist Tapes series, 1974, accessed January 19, 2016, https:// www.aip.org/history/exhibits/heisenberg/voice1.htm. 94. Ibid. 95. Ibid. 96. Ibid.

278

Notes

97. Ibid. 98. Werner Heisenberg, The Physical Principles of The Quantum Theory,” trans. Carl Eckart and Frank C. Scott (Mineola, NY: Dover, 1949), iii. 99. Henk de Regt, “Spacetime Visualisation and the Intelligibility of Physical Theories,” 262. 100. Heisenberg, Physics and Philosophy, 92. 101. Ibid., 90–91. 102. Kristian Camilleri, “Heisenberg and the Transformation of Kantian Philosophy,” International Studies in the Philosophy of Science 19, no. 3 (2005): 271. 103. Interview of Heisenberg by Kuhn and Heilbron, February 22, 1963. 104. Ibid., 3. 105. Schrödinger, Science,Theory, and Man, 30. 106. Ibid., 107–108. 107. Ibid., 108–109. 108. Ibid., 112. 109. Ibid. 110. Schrödinger, “Are There Quantum Jumps? Part I,” 110. 111. Ibid., 110. 112. Ibid., 111. 113. Catherine Chevalley, “Physics as Art: The German Tradition and the Symbolic Turn in Philosophy, History of Art and Natural Science in the 1920s,” in The Elusive Synthesis: Aesthetics and Science, ed. A. I. Tauber, 227–249 (Dordrecht, the Netherlands: Kluwer Academic Publishers, 1996), 232–233. 114. Ibid., 241. 115. Ibid. 116. Ibid., 241–242. It is unclear how Chevalley is using “objectivate” here. Earlier she seems to use it in the Kantian sense, but it is possible that here she is using it according to a meaning that is specific to quantum physics, in which case it should be understood to mean the interpretation of a quantum mechanical concept in terms of classical physics. 117. Bohr, “The Quantum Postulate,” 580. 118. Niels Bohr, “Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?,” in The Dreams that Stuff Is Made Of: The Most Astounding Papers of Quantum Physics—And How They Shook the World, ed. Stephen Hawking (Philadelphia: Running Press, 2011), 482.

Notes

279

119. Niels Bohr, “Discussions with Einstein on Epistemological Problems in Atomic Physics,” 1949, accessed January 25, 2016, https://www.marxists.org/ reference/subject/philosophy/works/dk/bohr.htm. 120. Bohr, “Discussions with Einstein.” 121. Niels Bohr, “Light and Life: Part I,” Nature 131 (1933): 421. 122. Werner Heisenberg to Einstein, in Why Beauty Is Truth: The History of Symmetry, ed. Ian Stewart (New York: Basic Books, 2007), 278. 123. Werner Heisenberg, Across the Frontiers (Woodbridge, CT: Ox Bow Press, 1990), 175. 124. Ibid., 170. 125. Ibid., 169. 126. Ibid., 171. 127. Ibid., 174. 128. Ibid. 129. Heisenberg, Across the Frontiers,” 175. 130. Ibid., 176. 131. Ibid., 170–171. 132. Ibid., 171. 133. Ibid., 170. 134. Johannes Kepler quoted in Heisenberg, Across the Frontiers, 178. 135. Ibid., 179. 136. Ibid., 180. CHAPTER 3

1. Robert Nadeau, Readings from the New Book of Nature: Physics and Metaphysics in the Modern Novel (Amherst: University of Massachusetts Press, 1981), 63–64. 2. Ibid., 14. 3. Ibid., 11. 4. Sue Sun Yom, “Biography and the Quantum Leap: Waves, Particles, and Light as a Theory of Writing the Human Life,” in Virginia Woolf: Texts and Contexts, ed. Beth Rigel Daugherty and Eileen Barrett (New York: Pace University Press, 1996), 145. 5. Gillian Beer, Virginia Woolf: The Common Ground (Ann Arbor: University of Michigan Press, 1996).

280

Notes

6. Holly Henry, Virginia Woolf and the Discourse of Science: The Aesthetics of Astronomy (Cambridge, UK: Cambridge University Press, 1983), 48, 58, 143. 7. Ibid., 7. 8. Ibid., 6. 9. Ibid., 57. 10. Nadeau, Readings from the New Book of Nature, 151. 11. Ibid. 12. Thomas Bohnenkamp, “Post-Einsteinian Physics and Literature: Toward a New Poetics,” Mosaic 22, no. 3 (1989): 19. 13. N. Katherine Hayles, “Gender Encoding in Fluid Mechanics: Masculine Channels and Feminine Flows,” differences: A Journal of Feminist Cultural Studies 4, no. 2 (1992): 22. 14. Yom, “Biography and the Quantum Leap,” 145. 15. Miriam Marty Clark, “Consciousness, Stream and Quanta in To The Lighthouse,” Studies in the Novel 21, no. 4 (1989): 416. 16. Ian Ettinger, “Relativity and Quantum Theory in Virginia Woolf ’s The Waves,” Zeteo:The Journal of Interdisciplinary Writing (2012): 4, accessed January 12, 2016, http://zeteojournal.com/2012/04/11/relativity-and-quantum-theory -in-the-waves-2/. 17. Gillian Beer, “Wave Theory and the Rise of Modernism,” in Realism and Representation: Essays on the Problem of Realism in Relation to Science, Literature, and Culture, ed. George Levine (Madison: University of Wisconsin Press, 1993), 121. 18. Clark, “Consciousness, Stream and Quanta,” 415–416. 19. Ibid., 417. 20. Paul Tolliver Brown, “Relativity, Quantum Physics, and Consciousness in Virginia Woolf ’s To the Lighthouse,” Journal of Modern Literature 32, no. 3 (2009): 52. 21. See, for example,Winnifred B. Cutler et al.,“Lunar Influences on the Reproductive Cycle in Women,” Human Biology 59, no. 6 (1987): 959–972; Sung Ping Law, “The Regulation of Menstrual Cycle and Its Relationship to the Moon,” Acta Obstetricia et Gynecologica Scandinavica 65, no. 1 (1986): 45–48. 22. Clark does acknowledge that Lily Briscoe sees the world in its “wavelikeness” as well as in its “particularity,” but uses this to support existing arguments that Lily is an androgynous character. It appears, then, that one may possess wave-like and particle-like qualities simultaneously only if one exists outside of gender categories altogether, and that to see something a certain way is to be a certain way.

Notes

281

23. Virginia Woolf, To the Lighthouse (Orlando: Harcourt and Brace, 1927), 7. 24. Ettinger, “Relativity and Quantum Theory,” 7. 25. Ibid., 11. 26. Renee Benham, “Gender Duality and Particle Physics in Virginia Woolf's Orlando,” February 22, 2013, accessed August 5, 2014, http://prezi.com/iedxk3x _clwv/gender-duality-and-particle-physics-in-virginia-woolfs-orlando/. See also Victoria Smith, “‘Ransacking the Language’: Finding the Missing Goods in Virginia Woolf ’s Orlando,” Journal of Modern Literature 29, no.4 (2006): 57–75. 27. Benham, “Gender Duality and Particle Physics in Virginia Woolf's Orlando.” 28. Daniel Albright, Quantum poetics:Yeats, Pound, Eliot, and the Science of Modernism (Cambridge, UK: Cambridge University Press), 1997. 29. Ibid., 17–18. 30. Ibid., 24–25. Albright mashes up the metaphor of wave/particle duality with, variously, gravitational pull, electromagnetic attraction, electron donation, a consideration of crystalline structures—a strategy that introduces an excess of meaning that dilutes and distracts. I have avoided these references in an effort to uncover the aspects of his model that more or less cohere into an identifiable interpretive approach. 31. Ibid., 19. 32. Ibid. 33. Ibid., 21. 34. W. John Coletta and David. H. Tamres, “Robert Frost and the Poetry of Physics,” The Physics Teacher 30, no. 360 (1992): 364. 35. Ibid. 36. Ibid. 37. Burt Kimmelman,“George Oppen’s Silence and the Role of Uncertainty in Post-war American Avant-Garde Poetry,” Mosaic 36, no. 2 (2003): 145. 38. Ibid., 151. 39. Ibid. 40. Ibid. 41. Guy Rotella also uses Bohr’s principle successfully to interpret poetry in “Comparing Conceptions: Frost and Eddington, Heisenberg, and Bohr,” American Literature 59 (1987):167–189. 42. Nick Herbert, quoted in Rebekka Edlund,“Carnival and Quantum Theory: Metaphors of Identity in Wilson Harris’s The Carnival Trilogy,” The Society for Caribbean Studies Annual Conference Papers 7, ed. Sandra Courtman (2006), accessed

282

Notes

April 29, 2014, http://community-languages.org.uk/SCS-Papers/olvol7p1 .PDF. 43. Edlund, “Carnival and Quantum Theory,” 2–3. 44. Ibid., 3. 45. Ibid. 46. Ibid. 47. Ibid. 48. Ibid., 4. 49. Kimmelman, “George Oppen’s Silence,” 158–159. 50. Ibid., 150. 51. Ibid., 158. 52. Wayne Kobylinski, “Getting to X: Paul Muldoon’s Quantum Poetics.” (PhD diss., University of North Carolina, Chapel Hill, 2005). Proquest (3170473). 53. Ibid., 23. 54. Ibid., 24. 55. J. D. O’Hara, “Donald Barthelme, The Art of Fiction No. 66,” Paris Review no. 80, 1981, accessed April 25, 2014, http://www.theparisreview.org/ interviews/3228/the-art-of-fiction-no-66-donald-barthelme. 56. Susan Strehle, Fiction in the Quantum Universe (Chapel Hill: University of North Carolina Press, 1992), 199. 57. Ibid., 195. 58. Ibid., 196. 59. See Sean Kinch,“Quantum Mechanics as Critical Model: Reading Nicholas Mosley’s Hopeful Monsters,” Studies in Contemporary Fiction 47, no. 3 (2006): 292. 60. James Fisher Solomon, Discourse and Reference in the Nuclear Age (London: University of Oklahoma Press, 1988), 75. 61. Ibid., 88. 62. Ibid., 75. 63. Ibid. 64. Christine Froula, “Quantum Physics/Postmodern Metaphysics: The Nature of Jacques Derrida,” Western Humanities Review 39, no. 4 (1985): 287–313. 65. Froula effectively engages the wave/particle “paradox,” the Uncertainty Principle, the dissolution of subject/object duality, and the phenomenon of action at a distance to show how all of them participate in undermining linear writing and the authority of Western logos.

Notes

283

66. Froula, “Quantum Physics/Postmodern Metaphysics, 296. 67. Ibid., 288. 68. Ibid., 292. 69. Ibid., 298. 70. Ibid. 71. Ibid., 295. 72. Kobylinski, “Getting to X,” 10–11. 73. M. Keith Booker, “Joyce, Planck, Einstein, and Heisenberg: A Relativistic Quantum Mechanical Discussion of Ulysses,” James Joyce Quarterly 27 (1990): 581. 74. Jason C. Smith, “Schrödinger’s Cat and Sarah’s Child: John Fowles’s ‘Quantum’ Narrative,” Mosaic 32, no. 2 (1999): 102–103. 75. Steven Carter, “‘A Place to Step Further’: Jack Spicer’s Quantum Poetics,” in Literature and Science as Modes of Expression, ed. Frederick Marine (Boston: Kluwer, 1989), 177–188. 76. Carter, “A Place to Step Further,” 47. 77. Ibid., 40. 78. Ibid., 45. 79. Smith, “Schrödinger’s Cat,” 101. 80. Ibid., 100. 81. Ibid., 101. 82. Ibid. Smith is referring to Heisenberg’s gamma ray microscope, which shoots a photon at the electron, either determining its position at the expense of its momentum, or vice versa. This is the foundation of the Uncertainty Principle, and of the assertion that the measuring apparatus is inevitably a part of the outcome. 83. Ettinger, “Relativity and Quantum Theory,” 8. 84. Their goal was to prove that it was indeed possible to measure conjugate qualities of particles because the measurement of one (which caused the wavefunction collapse into a single state) also caused the instantaneous collapse of the other particle, which remained undisturbed by the measuring apparatus—but appears to “know” that the other has been measured. Einstein remained unsatisfied with this result, because it still meant that information was traveling faster than the speed of light, which is a violation of the proven principles of relativity. 85. Ettinger, “Relativity and Quantum Theory,” 8.

284

Notes

86. Ibid. 87. Ibid. 88. Samuel Chase Coale, “Quantum Flux and Narrative Flow: Don DeLillo’s Entanglements with Quantum Theory,” Papers on Language and Literature 47, no. 3 (2011): 262. 89. Ibid., 269. 90. Ibid. 91. “Quantum fluctuation” describes the temporary appearance of energetic particles out of empty space, as consistent with the Uncertainty Principle. It may also be associated with an uncertainty with respect to time. 92. Coale, “Quantum Flux and Narrative Flow,” 269. 93. Ibid., 276. 94. Ibid., 290. 95. Alan Sokal,“Transgressing the Boundaries:Toward a Transformative Hermeneutics of Quantum Gravity,” Social Text no. 46/47 (1996): 217–252. 96. Alan Sokal and Jean Bricmont, Fashionable Nonsense: Postmodern Intellectuals’ Abuse of Science (New York: Picador USA, 1998). 97. Kinch, “Quantum Mechanics as Critical Model,” 301–302. 98. Ibid., 303. 99. Ibid. 100. Ibid. 101. Ibid., 306. 102. Bohm and Peat, Science, Order, and Creativity, 158. 103. Ibid., 164. 104. Mary Ellen Pitts,“The Holographic Paradigm: A New Model for the Study of Literature and Science,” Modern Language Studies 20, no. 4 (1990): 85. 105. Ibid., 86. 106. Mark Hussey, “To the Lighthouse and Physics: The Cosmology of David Bohm and Virginia Woolf,” in New Essays on Virginia Woolf, ed. Helen Wussow (Dallas: Contemporary Research Press, 1995), 85. 107. Ibid. 108. Ibid. 109. Ibid., 87. 110. Ibid., 85. 111. Ibid., 81.

Notes

285

112. Kinch, “Quantum Mechanics as Critical Model,” 295. 113. Ibid., 293. 114. Kinch offers the following examples: Carole Maso, Defiance (New York: Plume, 1999); H. R. McGregor, Schrödinger’s Baby (Lake Havascu City: London Bridge Publishers, 1998); Anna McGrail, Mrs. Einstein (New York: W. W. Norton & Company, 1998); David Ambrose, The Man Who Turned into Himself (London: Simon & Schuster UK, 2013). 115. Kinch, “Quantum Mechanics as Critical Mode,” 306. CHAPTER 4

1. Fritjof Capra, The Tao of Physics: An Exploration of the Parallels Between Modern Physics and Eastern Mysticism (London: Flamingo Press, 1983), 142. 2. Ibid., 146, 151. 3. Ibid., 150. 4. Ibid. 5. Ibid. 6. Werner Heisenberg, Physics and Philosophy: The Revolution in Modern Science (Amherst, NY: Prometheus Books, 1999), 54. 7. Victor J. Stenger, “The Myth of Quantum Consciousness,” Humanist 53, no. 3 (1992): 15. 8. Capra, The Tao of Physics, 151. 9. Ibid., 166. 10. Ibid. 11. Ibid., 166, 167. 12. Heisenberg, Physics and Philosophy, 15. 13. Robert Oppenheimer quoted in Capra, The Tao of Physics, 166. 14. Robert Oppenheimer, Science and the Common Understanding (Oxford: Oxford University Press, 1954), 9. 15. Capra, The Tao of Physics, 167. 16. Sal Restivo, “Parallels and Paradoxes in Modern Physics and Eastern Mysticism—A Critical Reconnaissance,” Social Studies of Science 8, no. 2 (1978): 148. 17. Ibid., 151. 18. Gary Zukav, The Dancing Wu Li Masters: An Overview of the New Physics (New York:William Morrow and Company, 2001), 5, 7.

286

Notes

19. Ibid., 49. 20. Ibid., 51. 21. Ibid. 22. Ibid., 71. 23. Ibid., 105–106. 24. Ibid., 240. 25. Ibid., 240–241. 26. Ibid., 228–229. 27. Ibid. 28. Eugene Wigner, “The Unreasonable Effectiveness of Mathematics in the Natural Sciences,” Communications in Pure and Applied Mathematics 13, no. 1 (1960): 1–14. 29. Eugene Wigner, “Remarks on the Mind-Body Problem,” in The Scientist Speculates, ed. By I. J. Good (London: William Heinemann, Ltd., 1961; New York: Basic Books, 1962), accessed March 14, 2017, http://www .informationphilosopher.com/solutions/scientists/wigner/Wigner_Remarks .pdf. 30. Popular accounts of “Wigner’s friend” usually describe it in terms of Schrödinger’s live and dead cat;Wigner actually used the example of photons. 31. Eugene Wigner, Symmetries and Reflections: Scientific Essays (Cambridge, MA: MIT Press, 1970). 32. David Kaiser, How the Hippies Saved Physics: Science, Counterculture, and the Quantum Revival (New York:W.W. Norton and Company, 2011), 75. 33. C. M Patton and John Wheeler, “Is Physics Legislated by Cosmology?” in Quantum Gravity: An Oxford Symposium, ed. C. J. Isham, Roger Penrose, and D.W. Sciama, 538–605. Oxford: Clarendon, 1975. 34. The Anthropic Universe,” Science Show, February 18, 2006. 35. Kaiser, How the Hippies Saved Physics, 80. 36. Ibid., 114. 37. See J. S. Bell, “On the Einstein-Podolsky-Rosen Paradox,” Physics 1 (1964): 195–200. 38. Kaiser, How the Hippies Saved Physics, 275. 39. Ibid., xvii. 40. Kaiser incorrectly attributes the term “entanglement” to Bell here (it was first used by Schrödinger in a letter to Einstein in 1935), but he is correct in the sense that Bell’s theorem describes many forms of entangled quantum states.

Notes

287

41. Kaiser, How the Hippies Saved Physics, 270. 42. In actual fact, nonmembers William Wootters, Wojciech Zurek, and Dennis Dieks published the no-cloning theorem, but its development was in part prompted by a proposal of Nick Herbert’s—a Fundamental Fysiks Group member—for a superluminal communication device using quantum entanglement. For an account of Herbert’s influence, see Asher Peres, “How the No-Cloning Theorem Got Its Name,” Fortschritte der Physik 51, no. 45 (2003): 458–461. 43. Kaiser, How the Hippies Saved Physics, 269, xii, xvi. 44. Ibid., 268. 45. Ibid., 269. 46. Ibid. 47. Ibid., 270. 48. Ibid., 193. 49. Ibid., 265. 50. Danah Zohar, “Forces of Reaction,” Sunday Times, February 6 1994, 14. 51. Zohar incorrectly states that physicists also call this a “wave packet,” which refers to the entirely distinct notion of a “probability wave” that describes the likelihood that a particle or particles will be measured to have a given position and momentum. 52. Zohar, “Forces of Reaction,” 14. 53. Ibid. 54. Ibid. 55. Ibid. 56. Danah Zohar and Ian Marshall, Quantum Society: Mind, Physics, and a New Social Vision (London: Harper Perennial, 1995), 43. 57. Ibid. 58. Ibid. 59. Ibid., 45. 60. Dawne C. McCance, “Physics, Buddhism, and Postmodernism,” Zygon 21, no. 3 (1986): 288. 61. Ibid., 289, 295. 62. Ibid. 298. 63. Ibid., 295. 64. Ibid., 288.

288

Notes

65. “The Quantum Society,” KirkusReviews.com, February 15, 1994, accessed May 28, 2014, https://www.kirkusreviews.com/book-reviews/danah-zohar/ the-quantum-society/. 66. Asghar Kazemi, “Quantum Politics: New Methodological Perspective,” E-Journal of International Law and Political Science, February 13, 2011, accessed October 15, 2016, http://scholarforum.blogspot.ca/2011/02/quantum-politics -new-methodological.html. 67. Ted Becker, ed. Quantum Politics:Applying Quantum Theory to Political Phenomena (Westport: Praeger Publishers, 1991). 68. Ibid., 58–59. 69. Ibid., xv. In a curious association, Becker argues that the “political quantum” provides the scientific basis for selecting representatives by random sampling. 70. Flora Lewis,“The Quantum Mechanics of Politics,” NewYork Times, November 6, 1983, 116. 71. Ibid., 99. 72. Ibid., 106. 73. Ibid. 74. Ibid., 99. 75. Ibid. 76. “Quantum Prosperity Program,” accessed September 11, 2014, http://www .quantumprosperityprogram.com (site discontinued). 77. For the collapse of the New Left, see Gerald Graff, “Co-Optation,” in The New Historicism, ed. Harold A.Veeser (New York: Routledge, 1989), 168. For the depoliticization of identity politics, see L. A. Kauffman, “The Anti-Politics of Identity,” Socialist Review 20, no. 1 (1990): 67–80. For the collapse of second wave feminism, see Susan Faludi, Backlash: The Undeclared War Against American Women (New York: Anchor, 1991); Deborah Siegel, Sisterhood, Interrupted: From Radical Women to Girls Gone Wild (New York: Palgrave Macmillan, 2007). 78. For examples of these new metrics, see Robert Costanza, S. Fisher, C. Ali, L. Beer, R. Bond et al., “Quality of Life: An Approach Integrating Opportunities, Human Needs, and Subjective Well-being,” Ecological Economics 61 (2008): 267– 276; Robert Costanza, M. Hart, S. Posner, and J.Talberth, Beyond GDP:The Need for New Measures of Progress (Boston: Boston University Press, 2009); E. Diener, “Subjective Well-being: The Science of Happiness and a Proposal for a National Index,” American Psychologist 55 no. 1 (2000): 34–43. 79. The title of the film is also written as What tΗē βLεεP Dθ wΣ (k)πow!? 80. What the Bleep Do We Know? (2004) directed by Betsy Chasse, Mark Vicente, and William Arntz (United States: Fox Video, 2006), DVD.

Notes

289

81. Jeremy Campbell, “Observer and Object, Reader and Text: Some Parallel Themes in Modern Science and Literature,” in Beyond the Two Cultures: Essays on Science,Technology, and Literature, ed. Joseph W. Slade and Judith Yaross Lee (Ames: Iowa State University Press, 1990), 23–37. 82. Deepak Chopra, Quantum Healing: Exploring the Frontiers of Mind/Body Medicine (New York: Bantam Books, 1990), 118–119. 83. Ibid., 241. 84. As biologists become more adept at observing phenomena on a nanoscale, there is a growing belief that key influences on gene expression and function may emanate from subatomic or quantum dynamics—in other words, that some form of the Uncertainty Principle may operate at the genetic level. Seeing DNA from a quantum perspective, they propose, offers a new way to understand how information could travel backward from the environment to DNA to produce adaptive mutations. This novel approach to explaining how mutations operate at the level of DNA, however, is distinct from the explicitly causal relation that Chopra constructs between the human will and DNA. 85. Chopra, Quantum Healing, 100. 86. Ibid. 87. Ibid., 97. 88. James A. Putnam,“Human Intelligence,” 2003, accessed November 25, 2016, http://newphysicstheory.com/New%20Physics%20Theory%202014/Human _Intelligence.pdf. 89. Richard Gordon, “Resonance, Life-Force, and the Principles of Quantum Touch,” May 2001, accessed July 24, 2017, https://www.quantumtouch.com/ en/about-quantum-touch/healing-stories/2283-resonance-life-force-and-the -principals-of-quantum-touch, 2017. 90. Jack Lyons, Quantum Touch, accessed March 13, 2016, https://www .quantumtouch.com/en/instructors/sch-instructors/instructor-835. 91. Robert Anthony, Robert Anthony Online, accessed May 15, 2014, http:// www.robertanthonyonline.com/. 92. Ibid. 93. Ibid. 94. The fact that the law of attraction has been debunked and replaced by quantum field theory is too obscure to challenge the rhetorical linking of the two. 95. Over one hundred sites associate the application of the law of attraction with getting rich.The use of the term “law of attraction” as a social concept dates back to the New Thought (spiritual) Movement at the beginning of the nineteenth

290

Notes

century. The first association of the term with prosperity can be found in Bruce MacLelland’s 1907 prosperity theology book, Prosperity Through Thought Force, which is based on the principle “you are what you think.”The contemporary addition of the quantum has gained a significant amount of ground, however, with dozens of sites linking success, quantum theory, and the law of attraction. 96. Anthony, Robert Anthony Online. 97. Bert Goldman, Quantum Jumping, 2017, accessed May 15, 2014, http:// www.quantumjumping.com/. CHAPTER 5

1. Ernest Rutherford, speech to the British Association for the Advancement of Science, London, 1932, reprinted as “Breaking Down the Atom,” The Times, September 12, 1933, 6. 2. Richard Rhodes, The Making of the Atomic Bomb (New York: Simon and Schuster, 1986), 292–293. 3. Sarah Henstra, The Counter-Memorial Impulse in Twentieth-Century English Fiction (New York: Palgrave Macmillan, 2009), 164. 4. Charles E. Gannon, Rumors of War and Infernal Machines:Technomilitary Agenda Setting in American and British Speculative Fiction (Lanham: Rowman and Littlefield, 2005), 129. 5. Ibid., 168. 6. Paul Chilton, “Nukespeak: Nuclear Language, Culture, and Propaganda,” in Nukespeak: The Media and the Bomb, ed. Aubrey Crispin (London: Comedia, 1982), 95. 7. I. A. Richards, The Philosophy of Rhetoric (London: Oxford University Press, 1935). 8. Ibid., 100. 9. Ibid. 10. Ibid., 117. 11. Ibid., 97. 12. Elizabeth Leane, “Knowing Quanta: The Ambiguous Metaphors of Popular Physics,” The Review of English Studies 52, no. 207 (2001): 420. 13. John Dennis quoted in Marjorie Hope Nicolson, “Sublime in External Nature,” Dictionary of the History of Ideas: Studies of Selected Pivotal Ideas, vol. 4, ed. Philip P.Wiener (New York: Charles Scribner’s Sons, 1968), 335. 14. Edmund Burke, A Philosophical Inquiry into the Origins of Our Ideas of the Sublime and the Beautiful (London: R. and J. Dodsley, 1757), 32.

Notes

291

15. For a discussion of Burke’s association of the sublime with the physiological, see:Vanessa L. Ryan, “The Physiological Sublime: Burke’s Critique of Reason,” Journal of History of Ideas 62, no. 1 (2001): 265–279. 16. Frances Ferguson, “The Nuclear Sublime,” Diacritics 14, no. 2 (1984): 6. 17. William L. Laurence, Men and Atoms: The Discovery, the Uses and the Future of Atomic Energy (New York: Simon and Schuster, 1959), 194. 18. William Laurence quoted in Sheldon Ungar, The Rise and Fall of Nuclearism: Faith and Fear as Determinants of the Arms Race (Philadelphia: Pennsylvania State University Press, 1992), 36. 19. Quoted in Margot A. Henriksen, Dr. Strangelove's America: Society and Culture in the Atomic Age (Oakland, CA: University of California Press, 1997), xv. 20. Chilton, “Nukespeak,” 97. 21. Harry S. Truman, “Statement by the President of the United States,” speech to the American public,Washington, DC, August 6, 1945, accessed May 16, 2014, https://www.trumanlibrary.org/publicpapers/index.php?pid=100&st=sun+dr aws&st1=. 22. Paul Saffo, “The Road from Trinity: Reflections on the Atom Bomb,” accessed May 16, 2014, http://www.saffo.com/essays/the-road-from-trinity -reflections-on-the-atom-bomb/. 23. Arthur H. Compton, “The Atomic Crusade and Its Social Implications,” Annals of the American Academy of Political and Social Science 249 (1947): 10; “The Atomic Age,” Life, August 20, 1945, 32. 24. Paul Boyer, By the Bomb’s Early Light: American Thought and Culture at the Dawn of the Atomic Age (New York: Pantheon, 1985), 9. 25. Laurence, Men and Atoms, 288. 26. Robert J. Oppenheimer quoted in Francis Donald Grabau, “The Christ Bomb at Trinity Site, New Mexico: An Occult Astrological Look at the World’s First Nuclear Bomb,” accessed June 24, 2014, http://starpathvisions.com/Faust -Christ-Bomb.htm. 27. Laurence, Men and Atoms, 98. 28. Ungar, The Rise and Fall of Nuclearism, 61. 29. Ibid. 30. Spencer R. Weart, “The Heyday of Myth and Cliché,” Bulletin of the Atomic Scientists 41, no. 7 (1985): 41. 31. Harry S. Hall, “Scientists and Politicians,” Bulletin of the Atomic Scientists 12, no. 2 (1956): 47.

292

Notes

32. Battery AD-X2: Hearings before the Select Committee on Small Business, U.S. Senate, 83rd Congress, 1st Session (U.S. Government Printing Office, Washington, DC, 1953), 304. 33. Science Legislation: Hearings before the Subcommittee of the Committee on Military Affairs, U.S. Senate, 79th Congress, 1st Session (U.S. Government Printing Office,Washington, DC, 1945), 1003. 34. Hall, “Scientists and Politicians,” 51. 35. Chilton, “Nukespeak,” 98. 36. Ibid., 98. 37. Robert J. Oppenheimer, “Physics in the Contemporary World,” Arthur D. Little Memorial Lecture at M.I.T., November 25, 1947. 38. Quoted in Ungar, The Rise and Fall of Nuclearism, 35. 39. Len Giovannitti, The Decision to Drop the Bomb, produced by Fred Freed (1965), DVD: Wilmette, IL: NBC, Encyclopaedia Britannica Films Inc., 2010. Oppenheimer mistakenly attributes the quote to the god Vishnu. Oppenheimer had read the original text in Sanskrit, and the translation was his own. The quotation was first printed in “The Eternal Apprentice,” Time, November 8, 1948, where the word “destroyer” was replaced with “shatterer.” This revision became standard when reproducing Oppenheimer’s statement. 40. Compton, “The Atomic Crusade and Its Social Implications,” 11. 41. William A. Gamson and Andre Modigliani, “Media Discourse and Public Opinion on Nuclear Power: A Constructionist Approach,” American Journal of Sociology 95 (1989): 12. 42. Ibid., 32. 43. “The Battle of Japan,” Time, August 13, 1945, 24. 44. “Awesome Force of Atom Bomb Loosed to Hasten Jap Surrender,” Newsweek, August 13, 1945, 30. 45. “The Jap Homeland,” Life, August 13, 1945, 32. 46. Vincent Anthony Leo, “The Mushroom Cloud: From Photographic Fact to Cultural Symbol” (MA thesis, Ohio State University, 1984), accessed May 16, 2014, https://www.ohiolink.edu/. 47. Ibid., 33. 48. Ibid., 35–36. 49. “War’s Ending: Atomic Bomb and Soviet Entry Bring Jap Surrender Offer,” Life, August 20, 1945, 30–31. 50. Leo, “The Mushroom Cloud,” 24–25.

Notes

293

51. Enola Gay was named after Tibbets’s mother. She was known as a gentle person, and Tibbets was very fond of her. Associating the plane with a maternal presence personalized it as a member of the crew’s “family,” softened the aura of aggression and hypermasculinity around its mission, and suggested that the plane would be “watching over” the men. 52. In reality,Tibbets was waving off the public so that he could take off, but his smile and wave easily fit into celebrity iconography, and the autograph merely consolidated this iconography.The picture nicely illustrates a comment by Spencer Weart: “a mechanism that is central to any history of images: associations already in the mind can creep into the picture that people think they perceive.” In Weart, Nuclear Fear: A History of Images (Cambridge, MA: Harvard University Press, 1988), 8. 53. An 8x10 reproduction of the autographed flight crew picture currently sells for £295.00 (roughly $485.00 USD). 54. Antony Rowland, “Silence and Awkwardness in Nuclear Discourse,” English 43, no. 176 (1994): 156. 55. Ibid., 165. 56. William L. Laurence, Dawn over Zero:The Story of the Atomic Bomb (New York: Knopf, 1946), 147. 57. Glenn D. Hook, “Making Nuclear Weapons Easier to Live With: The Political Role of Language in Nuclearization,” Bulletin of Peace Proposals 16, no. 1 (1985): 69. 58. Richard G. Hewlett and Oscar E. Anderson Jr., The New World, 1939–1946: A History of the Atomic Energy Commission, 3 vols. (University Park: Pennsylvania State University Press, 1962), 386. 59. Chilton, “Nukespeak,” 103. 60. Maxwell Leigh Eidinoff and Hyman Ruchlis, Atomics for the Millions (New York: McGraw Hill Book Company, 1947), 13. 61. John W. Campbell Jr., The Atomic Story (New York: Henry Holt and Company, 1947), vii. 62. In Joe Musial, Learn How Dagwood Splits the Atom (New York: King Features Syndicate, 1949. 63. Ibid., 8. 64. Ibid., 19. 65. Ibid., 22–23. 66. Ibid., 26. 67. Ibid., 35.

294

Notes

68. George Lakoff and Mark Johnson, Metaphors We Live By (Chicago: University of Chicago Press, 1989), 19. 69. Jacques Derrida, “No Apocalypse, Not Now (Full Speed Ahead: Seven Missiles, Seven Missives),” trans. C. Porter and P. Lewis. Diacritics 14, no. 2 (1984): 23. 70. Ibid. 71. Ibid. 72. Solomon, Discourse and Reference in the Nuclear Age (London: University of Oklahoma Press, 1988), 52. 73. Boyer, By the Bomb’s Early Light, 124. 74. “One Victory Not Yet Won,” New York Times, August 12, 1945. 75. Kai Mikkonen, “The ‘Narrative is Travel’ Metaphor: Between Spatial Sequence and Open Consequence,” Narrative 15, no. 3 (2007): 287. 76. Laurence, Dawn over Zero, 254. 77. Ibid., 149. 78. Gamson and Modigliani,“Media Discourse and Public Opinion on Nuclear Power,” 12. 79. Raymond G. Swing, “Commentary on the Atomic Age,” Mutual Broadcasting System, August 13 1945. 80. William R. Foulkes, “The Atom: Death—or the Life Abundant?” Vital Speeches of the Day 14, no. 1 (1948): 349. 81. Laurence, Dawn over Zero, 271. 82. William L. Laurence, “Paradise or Doomsday?,” Woman’s Home Companion, May 1948, 33. 83. Bernard M. Baruch, speech, United Nations Atomic Energy Commission, Hunter College, New York, June 14, 1946.Transcript in New York Times, June 15, 1946, 4. 84. Laurence, “Paradise or Doomsday?,” 33. 85. Norman Cousins, quoted in Bernard Baruch,“General Reaction to the U.S. Plan for Atomic Energy Control; A Survey of Radio Opinion,” June 16, 1946, in “Papers of Bernard Baruch,” Box 64, 138–139, Princeton University, Princeton, NJ. 86. Raymond Swing quoted in Boyer, By the Dawn’s Early Light, 83. 87. Truman, “Statement by the President of the United States.” 88. Ibid. 89. Ungar, The Rise and Fall of Nuclearism, 80, 180.

Notes

295

90. Dwight Eisenhower, “Atoms for Peace,” speech, United Nations General Assembly, New York, December 8, 1953, accessed June 14, 2014, https://www .iaea.org/about/history/atoms-for-peace-speech. 91. Ibid. 92. Ibid. 93. Ibid. 94. Gamson and Modigliani,“Media Discourse and Public Opinion on Nuclear Power,” 13. 95. Boyer, By the Bomb’s Early Light, 13. 96. W. W. Waymack, “A Letter to Judith and Dickie,” National Education Association Journal 36 (1947): 214. CONCLUSION

1. Arthur Eddington, The Nature of the Physical World (New York:The University Press, 1929), 290. 2. Ibid., 291. 3. Ibid., 292. 4. There have been developments around the method of “weak measurement” of a quantum system, which claims that one may gain some information about one property (say, position) without appreciably disturbing the complementary property (momentum). In the experiment, a stream of photons with identical wavefunctions are “sampled.” However, this method does not measure the properties of a single particle, but rather measures properties, believed to be identical, of a large number of particles, and then takes the average of the observations as representative of the properties of a single particle. Here, one can measure multiple properties of many “identical” particles and deduce what the properties of a single one would have been—but this deduction takes place after the event. Therefore, the premise that you cannot measure all properties of a single particle at the same time still holds. For a description of this process, see: Jeff S. Lundeen, Brandon Sutherland, Aabid Patel, Corey Stewart, and Charles Bamber, “Direct Measurement of the Quantum Wavefunction,” Nature 474 (2011): 188–191. 5. Werner Heisenberg, Physics and Philosophy: The Revolution in Modern Science (Amherst, NY: Prometheus books, 1999), 144. 6. Ibid., 144–145. 7. Niels Bohr quoted in Abraham Pais, The Genius of Science: A Portrait Gallery (Oxford: Oxford University Press, 2000), 24.

296

Notes

8. In modern nuclear bombs, U-238 is used as a tamper to increase the efficiency of the weapon by reducing the critical mass required. In thermonuclear weapons, U-238 is used to add more energy and thus more “yield” to the weapon. 9. J. B. S. Haldane, Possible Worlds and Other Papers (New York: Harper and Brothers, 1928), 286. Similar remarks that appear to be misappropriations of Haldane’s words have been attributed to Eddington (“The universe is not only stranger than we imagine, it is stranger than we can imagine”) and to Heisenberg (“Not only is the Universe stranger than we think, it is stranger than we can think”). The Eddington quotation is always unattributed, while the Heisenberg quotation is misattributed to his 1973 book, Across the Frontiers, which contains no such words. In fact, most references to Across the Frontiers on the Internet are attributions of this quotation, and there are hundreds of the same misattribution. This seems to be a case of incorrect information being introduced via the Internet and then cited in exponential proliferation. 10. Richard Feynman, Feynman Lectures on Physics, vol. 1, ed. P. Richard, Robert B. Leighton, and Matthew Feynman Sands (Boston: Addison Wesley, 1964), 1–2.

BIBLIOGRAPHY

Aaserud, Finn. “Niels Bohr’s Mission for an ‘Open World.’” From Proceedings of the 2nd ICESHS: The Global and the Local: The History of Science and the Cultural Integration of Europe, ed. M. Kokowski. Cracow, Poland, September 6–9, 2006, 708. Accessed June 26, 2014. http://www.2iceshs.cyfronet.pl/2ICESHS _Proceedings/Chapter_25/R-17_Aaserud.pdf. Agassi, Joseph. “The Rhetoric of Science.” Philosophy of the Social Sciences 29 (1999): 329–335. Albright, Daniel. Quantum Poetics:Yeats, Pound, Eliot, and the Science of Modernism. Cambridge, UK: Cambridge University Press, 1997. Ambrose, David. The Man Who Turned into Himself. London: Simon and Schuster UK, 2013. Anthony, Robert. Robert Anthony Online. Accessed May 15, 2014. http://www .robertanthonyonline.com/. “The Anthropic Universe.” Science Show. February 18, 2006. Accessed March 13, 2017. http://www.abc.net.au/radionational/programs/scienceshow/the -anthropic-universe/3302686. Asgari, Tayebeh. “The Study of Image Schemas in Hafez Poems: Cognitive Perspective.” International Journal of Language and Linguistics 1 (4) (2013): 182–190. “The Atomic Age.” Life, August 20, 1945, 32. “Awesome Force of Atom Bomb Loosed to Hasten Jap Surrender.” Newsweek, August 13, 1945. Baruch, Bernard. “General Reaction to the U.S. Plan for Atomic Energy Control; A Survey of Radio Opinion,” June 16, 1946. In “Papers of Bernard Baruch,” Box 64, 138–139. Princeton University, Princeton, NJ. Baruch, Bernard M. “Speech at Opening Session of U.N. Atomic Energy Commission.” New York Times, June 15 1946.

298

Bibliography

Baruch, Bernard M. Speech, United Nations Atomic Energy Commission, Hunter College, New York, June 14, 1946.Transcript in New York Times, June 15, 1946, 4. Battery AD-X2: Hearings before the Select Committee on Small Business. U.S. Senate, 83rd Congress, 1st Session. Washington, DC: U.S. Government Printing Office, 1953. “The Battle of Japan.” Time, August 1945, 13. Bazerman, Charles. “Emerging Perspectives on the Many Dimensions of Scientific Discourse.” In Reading Science: Critical and Functional Perspectives on Discourses of Science, ed. J. R. Martin and Robert Vee1, 16–28. London: Routledge, 1998. Becker, Ted. Quantum Politics: Applying Quantum Theory to Political Phenomena. Westport: Praeger Publishers, 1991. Beer, Gillian. Open Fields: Science in Cultural Encounter. London: Oxford University Press, 1986. Beer, Gillian. Virginia Woolf: The Common Ground. Ann Arbor: University of Michigan Press, 1996. Beer, Gillian. “Wave Theory and the Rise of Modernism.” In Realism and Representation: Essays on the Problem of Realism in Relation to Science, Literature, and Culture, ed. George Levine, 193–213. Madison: University of Wisconsin Press, 1993. Bell, J. S. “Bertlmann’s Socks and the Nature of Reality.” In J. S. Bell, Speakable and Unspeakable in Quantum Mechanics: Collected Papers on Quantum Philosophy, 139–158. Cambridge, UK: Cambridge University Press, 2004. Bell, J. S. “On the Einstein-Podolsky-Rosen Paradox.” Physics 1 (1964): 195–200. Beller, Mara. “Matrix Theory before Schrödinger: Philosophy, Problems, Consequences.” Isis 74 (4) (1983): 469–491. Beller, Mara. Quantum Dialogue: The Making of a Revolution. Chicago: University of Chicago Press, 1999. Benham, Renee. “Gender Duality and Particle Physics in Virginia Woolf's Orlando.” February 22, 2013. Accessed August 5, 2014. http://prezi.com/ iedxk3x_clwv/gender-duality-and-particle-physics-in-virginia-woolfs -orlando/. “Black Cross over Orange Circle.” [Cover Drawing.] Time (August) (1945). Bohm, David, and F. David Peat. Science, Order, and Creativity. London: Routledge, 2000. Bohnenkamp, Thomas. “Post-Einsteinian Physics and Literature: Toward a New Poetics.” Mosaic 22 (3) (1989): 19–30.

Bibliography

299

Bohr, Niels. Atomic Theory and Human Knowledge. Mineola, NY: Dover Publications, 2010. Bohr, Niels. “Atomic Theory and Mechanics.” In The Philosophical Writings of Niels Bohr. Vol. I. Atomic Theory and the Description of Nature. Woodbridge, CT: Ox Bow Press, 1987. Bohr, Niels. “The Atomic Theory and the Fundamental Principles Underlying the Description of Nature.” In The Philosophical Writings of Niels Bohr. Vol. I. Atomic Theory and the Description of Nature, 102–119. Woodbridge, CT: Ox Bow Press, 1987. Bohr, Niels. “Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?” In The Dreams that Stuff Is Made Of:The Most Astounding Papers of Quantum Physics—And How They Shook the World, ed. Stephen Hawking, 471–483. Philadelphia: Running Press, 2011. Bohr, Niels. “Causality and Complementarity.” Philosophy of Science 4 (3) (1937): 289–298. Bohr, Niels. “Discussions with Einstein on Epistemological Problems in Atomic Physics,” 1949. Accessed January 25, 2016. https://www.marxists.org/reference/ subject/philosophy/works/dk/bohr.htm. Bohr, Niels. “Light and Life: Part I.” Nature 131 (1933): 421–423. Bohr, Niels. “Maxwell and Modern Theoretical Physics,” Nature 128 (3234) (1931): 691–692. Bohr, Niels. “Natural Philosophy and Human Cultures.” Nature 143 (1939): 268–272. Bohr, Niels. Niels Bohr Collected Works. Vol. III. The Correspondence Principle (1918–1923), ed. J. R. Nielsen. Amsterdam: North-Holland Publishing, 1976. Bohr, Niels. The Philosophical Writings of Niels Bohr.Vol. I. Atomic Theory and the Description of Nature. Woodbridge, CT: Ox Bow Press, 1987. Bohr, Niels. The Philosophical Writings of Niels Bohr. Vol. III. Essays 1958–1962 on Atomic Physics and Human Knowledge. Woodbridge, CT: Ox Bow Press, 1987. Bohr, Niels. “The Quantum Postulate and the Recent Development of Atomic Theory.” Supplement to Nature 21 (3050) (1928): 580–590. Bono, James J. “Science, Discourse, and Literature: The Role/Rule of Metaphor in Science.” In Literature and Science: Theory and Practice, ed. Stuart Peterfreund, 59–89. Lebanon, NH: University Press of New England, 1990. Booker, M. Keith. “Joyce, Planck, Einstein, and Heisenberg: A Relativistic Quantum Mechanical Discussion of Ulysses.” James Joyce Quarterly 27 (1990): 577–658.

300

Bibliography

Born, Max. “On the Quantum Mechanics of Collision.” In Quantum Theory and Measurement, ed. J. A. Wheeler and W. H. Zurek, 52–55. Princeton: Princeton University Press, 1983. Boyer, Paul. By the Bomb’s Early Light: American Thought and Culture at the Dawn of the Atomic Age. New York: Pantheon, 1985. Brooks, Arthur. Gross National Happiness: Why Happiness Matters for America and How We Can Get More of It. New York: Basic Books, 2008. Brown, Paul Tolliver. “Relativity, Quantum Physics, and Consciousness in Virginia Woolf ’s To the Lighthouse.” Journal of Modern Literature 32 (3) (2009): 39–62. Brown, Theodore. Making Truth: Metaphor in Science. Champaign: University of Illinois Press, 2008. Burke, Edmund. A Philosophical Inquiry into the Origins of Our Ideas of the Sublime and the Beautiful. London: R. and J. Dodsley, 1757. Camac, Mary K., and Samuel Glucksberg. “Metaphors Do Not Use Associations Between Concepts, They Are Used to Create Them.” Journal of Psycholinguistic Research 13 (1984): 443–455. Camilleri, Kristian. “Heisenberg and the Transformation of Kantian Philosophy.” International Studies in the Philosophy of Science 19 (3) (2005): 271–287. Campbell, Jeremy. “Observer and Object, Reader and Text: Some Parallel Themes in Modern Science and Literature.” In Beyond the Two Cultures: Essays on Science,Technology, and Literature, ed. Joseph W. Slade and Judith Yaross Lee, 23–37. Ames: Iowa State University Press, 1990. Campbell, John W., Jr. The Atomic Story. New York: Henry Holt and Company, 1947. Campbell, Paul N. “Poetic-Rhetorical, Philosophical, and Scientific Discourse.” Philosophy & Rhetoric 6 (1) (1973): 1–29. Capra, Fritjof. The Tao of Physics: An Exploration of the Parallels Between Modern Physics and Eastern Mysticism. London: Flamingo Press, 1983 [1975]. Capra, Fritjof. The Turning Point: Science, Society, and the Rising Culture. Riverside: Simon and Shuster, 1982. Carrion, Jorge. “The Bicephalous Writer: The Commingling of the Creative Writer and the Critic in a Single Body.” Hispanic Issues On Line (9) (2012): 13–21. Carter, Steven. Bearing Across: Studies in Science and Literature. Lanham: International Scholars Publications, 2002. Carter, Steven. “‘A Place to Step Further’: Jack Spicer’s Quantum Poetics.” In Literature and Science as Modes of Expression, ed. Frederick Marine, 177–188. Boston: Kluwer, 1989.

Bibliography

301

Carter, Steven. “Robert Duncan and Erwin Schrödinger: Esthetics of the Wavefunction.” Studies in the Humanities 17 (1) (1990): 36–48. Ceccarelli, Leah. Shaping Science with Rhetoric. Chicago: University of Chicago Press, 2001. Chevalley, Catherine. “Physics as Art: The German Tradition and the Symbolic Turn in Philosophy, History of Art and Natural Science in the 1920s.” In The Elusive Synthesis: Aesthetics and Science, ed. A. I. Tauber, 227–249. Dordercht, The Netherlands: Kluwer Academic Publishers, 1996. Chilton, Paul. “Nukespeak: Nuclear Language, Culture, and Propaganda.” In Nukespeak: The Media and the Bomb, ed. Aubrey Crispin, 94–112. London: Comedia, 1982. Chopra, Deepak. Quantum Healing: Exploring the Frontiers of Mind/Body Medicine. New York: Bantam Books, 1990. Clark, Miriam Marty. “Consciousness, Stream and Quanta in To The Lighthouse.” Studies in the Novel 21 (4) (1989): 413–424. Coale, Samuel Chase. “Quantum Flux and Narrative Flow: Don DeLillo’s Entanglements with Quantum Theory.” Papers on Language & Literature 47 (3) (2011): 261–294. Cole, Laura. “The Quantum Question: How 20th Century Physical Discovery Shaped Virginia Woolf ’s The Waves.” 2012. Accessed September 1, 2016. http:// tracklaura.com/2012/07/14/the-quantum-question/. Coletta,W. John, and David H.Tamres. “Robert Frost and the Poetry of Physics.” Physics Teacher 30 (360) (1992): 360–365. Compton, Arthur H. “The Atomic Crusade and Its Social Implications.” Annals of the American Academy of Political and Social Science 249 (1947): 9–19. Connor, Steven. “Quantum Writing: Literature and the World of Numbers.” Lecture Given at the Institute of Continuing Education, Cambridge, England, June, 2013. Costanza, Robert, S. Fisher, C. Ali, L. Beer, R. Bond et al. “Quality of Life: An Approach Integrating Opportunities, Human Needs, and Subjective Well-being.” Ecological Economics 61 (2008): 267–276. Costanza, Robert, M. Hart, S. Posner, and J. Talberth. Beyond GDP: The Need for New Measures of Progress. Boston: Boston University Press, 2009. Cuffaro, Michael. “Studies in History and Philosophy of Science Part B.” Studies in History and Philosophy of Modern Physics 41 (4) (2010): 309–317. Cutler, Winnifred B., Wolfgang M. Schleidt, Erika Freidmann, George Preti, and Robert Stine. “Lunar Influences on the Reproductive Cycle in Women.” Human Biology 59 (6) (1987): 959–972.

302

Bibliography

Davies, Paul C.W.“Does Quantum Mechanics Play a Non-Trivial Role in Life?” Bio Systems 78 (2004): 69–79. Davies, Paul C. W. Other Worlds: A Portrait of Nature in Rebellion: Space, Superspace, and the Quantum Universe. New York: Simon and Schuster, 1980. Davies, Paul C.W. Quantum Aspects of Life. London: Imperial College Press, 2008. Dawkins, Richard. The God Delusion. Boston: Mariner Books, 2008. de Regt, Henk W. “Erwin Schrödinger, Anschaulechkeit, and Quantum Theory.” Studies in the History and Philosophy of Modern Physics 28 (4) (1997): 461–481. de Regt, Henk W. “Spacetime Visualisation and the Intelligibility of Physical Theories.” Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 32 (2) (2001): 243–265. Derrida, Jacques.““No Apocalypse, Not Now (Full Speed Ahead: Seven Missiles, Seven Missives),” trans. C. Porter and P. Lewis. Diacritics 14 (2) (1984): 20–31. DeWitt, B., and Neill Graham, eds. The Many Worlds Interpretation of Quantum Mechanics. Princeton: Princeton University Press, 1973. Diener, E. “Subjective Well-being: The Science of Happiness and a Proposal for a National Index.” American Psychologist 55 (1) (2000): 34–43. DuPlessis, Rachel Blau. “‘The Familiar/Becomes Extreme’: George Oppen and Silence.” North Dakota Quarterly 55 (4) (1987): 18–36. Eddington, Arthur. The Nature of the Physical World. New York: The University Press, 1929. Edie, James. Speaking and Meaning:The Phenomenology of Language. Bloomington: Indiana University Press, 1976. Edlund, Rebekka. “Carnival and Quantum Theory: Metaphors of Identity in Wilson Harris’s The Carnival Trilogy.” The Society for Caribbean Studies Annual Conference Papers 7, ed. Sandra Courtman, 2006, 1–4. Accessed April 29, 2014. http://community-languages.org.uk/SCS-Papers/olvol7p1.PDF. Ehrlich, Matthew. “Living with the Bomb: Fred Friendly’s ‘The Quick and the Dead.’” Journalism History 35 (1) (2009): 2–11. Eidinoff, Maxwell Leigh, and Hyman Ruchlis. Atomics for the Millions. New York: McGraw Hill Book Company, 1947. Einstein, Albert, and Leopold Infeld. The Evolution of Physics: From Early Concepts to Relativity and Quanta. London: Simon and Schuster, 1966. Einstein,Albert, Boris Podolsky, and Nathan Rosen.“Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?” Physical Review 47 (10) (1935): 777–780. “Einstein, The Man Who Started It All.” Newsweek, March 10, 1947.

Bibliography

303

Eisenhower, Dwight. “Atoms for Peace.” Speech, United Nations General Assembly, New York, December 8, 1953. Accessed June 15, 2014. https:// www.iaea.org/about/history/atoms-for-peace-speech. Eisenstaedt, Alfred. “V-J Day in Times Square.” Life, August 1945, 27. “The Eternal Apprentice.” Time, November 8, 1948. Ettinger, Ian. “Relativity and Quantum Theory in Virginia Woolf ’s The Waves.” Zeteo: The Journal of Interdisciplinary Writing (2012): 1–19. Accessed January 12, 2016. http://zeteojournal.com/2012/04/11/relativity-and-quantum-theory -in-the-waves-2/. Evans,Vyvyan. A Glossary of Cognitive Linguistics. Edinburgh: Edinburgh University Press, 2007. Evans,Vyvyan, Benjamin Bergen, and Jörg Zinken, eds. The Cognitive Linguistics Reader. London: Equinox Publishing, 2007. Everett, Hugh. “‘Relative State’ Formulation of Quantum Mechanics.” Reviews of Modern Physics 20 (3) (1957): 454–462. Everett, Hugh. “The Theory of the Universal Wavefunction.” In The ManyWorlds Interpretation of Quantum Mechanics, ed. Bryce DeWitt and Neill Graham, 3–140. Princeton: Princeton University Press, 1973. Fahnestock, Jeanne. “Accommodating Science:The Rhetorical Life of Scientific Facts.” Written Communication 3 (3) (1986): 275–296. Fahnestock, Jeanne. Rhetorical Figures in Science. Oxford: Oxford University Press, 2002. Faludi, Susan. Backlash: The Undeclared War Against American Women. New York: Anchor, 1991. Ferguson, Frances. “The Nuclear Sublime.” Diacritics 14 (2) (1984): 4–10. Feynman, Richard P. The Character of Physical Law. New York: Modern Library, 1965. Feynman, Richard P. Feynman Lectures on Physics.Vol. 1, ed. P. Richard, Robert B. Leighton, and Matthew Feynman Sands. Boston: Addison Wesley, 1964. Feynman, Richard P. “Simulating Physics with Computers.” International Journal of Theoretical Physics 21 (6) (1982): 467–488. Forman, Paul. “Kausalitiit, Anschaulichkeit, and Individualitiit, or How Cultural Values Prescribed the Character and the Lessons Ascribed to Quantum Mechanics.” In Society and Knowledge, ed. N. Stehr and V. Meja, 333–347. New Brunswick: Transaction Books, 1984. Forman, Paul. “Weimar Culture, Causality and Quantum Theory, 1918–27.” In Weimar Culture and Quantum Mechanics: Selected Papers by Paul Forman and

304

Bibliography

Contemporary Perspectives on the Forman Thesis, ed. Cathryn Carson, Alexei Kojevnikov, and Helmuth Trischler, 85–203. London: Imperial College Press, 2015. Foulkes, William R. “The Atom: Death—or the Life Abundant?” Vital Speeches of the Day 14 (1) (1948): 345–349. Freire, Olival Jr. The Quantum Dissidents: Rebuilding the Foundations of Quantum Mechanics (1950–1990). Berlin: Springer, 2015. Freire, Olival Jr. “Science and Exile: David Bohm, the Hot Times of the Cold War, and His Struggle for a New Interpretation of Quantum Mechanics.” Historical Studies in the Physical and Biological Sciences 36 (1) (2005): 31–35. Froula, Christine. “Quantum Physics/Postmodern Metaphysics: The Nature of Jacques Derrida.” Western Humanities Review 39 (4) (1985): 287–313. Galton, Frances. Inquiries into Human Faculty and Its Development, ed. Gavan Tredoux. 2001. Accessed November 16, 2016. http://galton.org/books/human -faculty/text/human-faculty.pdf. Gamson, William A., and Andre Modigliani. “Media Discourse and Public Opinion on Nuclear Power: A Constructionist Approach.” American Journal of Sociology 95 (1989): 1–37. Gannon, Charles E. Rumors of War and Infernal Machines: Technomilitary Agenda Setting in American and British Speculative Fiction. Lanham: Rowman and Littlefield, 2005. Gibbs, Raymond W., Jr. “What Do Idioms Really Mean?” Journal of Memory and Language 31 (1992): 485–506. Gibbs, Raymond W., Jr. “Why Idioms Are Not Dead Metaphors.” In Idioms: Processing, Structure, and Interpretation, ed. Christina Cacciari and Patrizia Tabossi, 57–77. Hillsdale, NJ: Erlbaum, 1993. Gibbs, Raymond W., Jr., and Jennifer E. O’Brien. “Idioms and Mental Imagery: The Metaphorical Motivation for Idiomatic Meaning.” Cognition 36 (1990): 35–68. Giovannitti, Len. The Decision to Drop the Bomb. Documentary. Produced by Fred Freed, 1965. DVD: Wilmette: NBC, Encyclopedia Britannica Films, Inc., 2010. Glucksberg, Samuel, Mary Brown, and Matthew McGlone. “Conceptual Metaphors Are Not Automatically Accessed During Idiom Comprehension.” Memory & Cognition 21 (1993): 711–719. Goldman, Bert. “Quantum Jumping.” 2017. Accessed May 15, 2014. http:// www.quantumjumping.com/. Gordon, Richard. “Resonance, Life-Force, and the Principles of Quantum Touch.” May 2001. Accessed July 24, 2017. https://www.quantumtouch.com/

Bibliography

305

en/about-quantum-touch/healing-stories/2283-resonance-life-force-and-the -principals-of-quantum-touch, 2017. Goswami, Amit. The Self-Aware Universe: How Consciousness Creates the Material World. Los Angeles: Tarcher Books, 1975. Grabau, Francis Donald. “The Christ Bomb at Trinity Site, New Mexico: An Occult Astrological Look at the World’s First Nuclear Bomb.” Accessed June 24, 2014. http://starpathvisions.com/Faust-Christ-Bomb.htm. Graff, Gerald. “Co-Optation.” In The New Historicism, ed. Harold A. Veeser, 168–182. New York: Routledge, 1989. Grangier, Philippe, Gérard Roger, and Alain Aspect. “Experimental Evidence for a Photon Anticorrelation Effect on a Beam Splitter: A New Light on Single-photon Interferences.” Europhysics Letters 1 (4) (1986): 17–79. Graves, Heather B. Rhetoric In(to) Science: Style as Invention in Inquiry. Cresskill: Hampton Press, 2005. Graybar, Lloyd J., and Ruth Flint Graybar. “America Faces the Atomic Age: 1946.” Air University Review 35 (2) (1984): 68–77. Gross, Alan. The Rhetoric of Science. Cambridge, MA: Harvard University Press, 1990. Gross, Alan. Starring the Text: The Place of Rhetoric in Science Studies. Carbondale: Southern Illinois University Press, 2006. Haldane, J. B. S. Possible Worlds and Other Papers. New York: Harper and Brothers, 1928. Hall, Harry S. “Scientists and Politicians.” Bulletin of the Atomic Scientists 12 (2) (1956): 46–52. Halloran, Michael S., and Annette N. Bradford. “Figures of Speech in the Rhetoric of Science.” In Essays on Classical Rhetoric and Modern Discourse, ed. Robert J. Connors, Lisa S. Ede, Andrea A. Lunsford, and S. W. Smith, 179–192. Carbondale: Southern Illinois University, 1984. Harris, R. A., ed. Landmark Essays in the Rhetoric of Science. Mahwah, NJ: Hermagoras Press, 1997. Harris, R. Alan. “Knowing, Rhetoric, Science.” In Visions and Revisions: Continuity and Change in Rhetoric and Composition, ed. J. Williams, 163–219. Carbondale: Southern Illinois University Press, 2002. Hayles, N. Katherine. The Cosmic Web. Ithaca: Cornell University Press, 1984. Hayles, N. Katherine. “Gender Encoding in Fluid Mechanics: Masculine Channels and Feminine Flows.” differences: A Journal of Feminist Cultural Studies 4 (3) (1992): 16–44. “H-Bomb Secrecy.” Washington Post, August 22, 1953.

306

Bibliography

Heisenberg, Werner. Across the Frontiers. Woodbridge, CT: Ox Bow Press, 1990. Heisenberg, Werner. “The Actual Content of Quantum Theoretical Kinematics and Mechanics,” National Aeronautics and Space Administration. Translator of original 1927 German publication unknown. Washington, DC, 1983. Accessed January 19, 2016. http://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa .gov/19840008978.pdf. Heisenberg, Werner. “Development of Concepts in Quantum Theory.” In The Physicists’ Conception of Nature, ed. Jagdish Mehra, 264–275. Boston: Reidel, 1973. Heisenberg, Werner. Interview of Werner Heisenberg by Thomas S. Kuhn and John Heilbron. Session dates: November 30, 1962; February 7, 11, 13, 15, 19, 22, 25, 27, 28, 1963; and July 5, 12, 1963. Niels Bohr Library & Archives, American Institute of Physics, College Park, Maryland. Accessed January 19, 2016. https://www.aip.org/history-programs/oral-histories/search?search_api _views_fulltext=heisenberg. Heisenberg,Werner. The Physical Principles of the Quantum Theory, trans. C. Eckart and F. C. Hoyt. Mineola, NY: Dover, 1949 [1930]. Heisenberg, Werner. Physics and Beyond. London: George Allen and Unwin Ltd., 1971. Heisenberg, Werner. Physics and Philosophy: The Revolution in Modern Science. Amherst, NY: Prometheus Books, 1999. Heisenberg, Werner. “Quantenmechanik.” Naturwissenschaften 14 (1926): 899–994. Heisenberg,Werner. “The Representation of Nature in Contemporary Physics.” Daedalus 87 (3) (1958): 95–108. Heisenberg, Werner. Scientific Correspondence with Bohr, Einstein, Heisenberg, ed. Alan Herman, Karl Von Meyan, and Victor F. Weisskopf. Berlin: Springer-Verlag, 1985. Heisenberg, Werner. Transcript of audiotape titled “The Development of the Uncertainty Principle.” Produced by Spring Green Multimedia in the UniConcept Scientist Tapes series, 1974. Accessed January 19, 2016. https:// www.aip.org/history/exhibits/heisenberg/voice1.htm. Heisenberg, Werner. “Über den anschaulichen inhalt der quantentheoretischen kinematic und mechanik.” Zeitschrift fur Physik 43 (3–4) (1926): 172–198. Heisenberg, Werner. “Über quantentheoretishe umdeutung kinematisher und mechanischer beziehungen.” Zeitschrift fur Physik 33 (1925): 879–893. Heisenberg, Werner. Why Beauty Is Truth: The History of Symmetry, ed. Ian Stewart. New York: Basic Books, 2007.

Bibliography

307

Hellsten, Lina. “Popular Metaphors of Bioscience: Bridges over Time?” Configurations 16 (1) (2008): 11–32. Henriksen, Margot A. Dr. Strangelove's America: Society and Culture in the Atomic Age. Oakland, CA: University of California Press, 1997. Henry, Holly. Virginia Woolf and the Discourse of Science:The Aesthetics of Astronomy. Cambridge, UK: Cambridge University Press, 1983. Henstra, Sarah. The Counter-Memorial Impulse in Twentieth-Century English Fiction. New York: Palgrave Macmillan, 2009. Hermann, Armin. “Erwin Schrödinger.” In Dictionary of Scientific Biography. Vol. 12, ed. Charles Coulston Gillispie, 217–223. New York: Charles Scribner & Sons, 1975. Hewlett, Richard G., and Oscar E. Anderson, Jr. The New World, 1939–1946: A History of the Atomic Energy Commission, 3 vols. University Park: Pennsylvania State University Press, 1962. Hook, Glenn D. “Making Nuclear Weapons Easier to Live With: The Political Role of Language in Nuclearization.” Bulletin of Peace Proposals 16 (1) (1985): 67–77. Howard, Don. “Revisiting the Einstein-Bohr Dialogue.” Accessed January 23, 2016. https://www3.nd.edu/~dhoward1/Revisiting%20the%20Einstein -Bohr%20Dialogue.pdf. Howard, Don. “Who Invented the ‘Copenhagen Interpretation’? A Study in Mythology.” Philosophy of Science 71 (2004): 669–682. Hussey, Mark. “To the Lighthouse and Physics: The Cosmology of David Bohm and Virginia Woolf.” In New Essays on Virginia Woolf, ed. Helen Wussow, 79–97. Dallas: Contemporary Research Press, 1995. Jammer, Max. The Conceptual Development of Quantum Mechanics. New York: McGraw-Hall Book Company, 1966. “A Jap Burns.” Photograph. Life, August 1945, 13. “The Jap Homeland.” Life, August 13, 1945. “Jet Plane.” Cover photograph. Life, August 1945, 13. Johnson, Mark. The Body in the Mind:The Bodily Basis of Meaning, Imagination, and Reason. Chicago: The University of Chicago Press, 1987. Kaempffer, Waldemar. Science,Today and Tomorrow. New York:Viking, 1946. Kaiser, David. How the Hippies Saved Physics: Science, Counterculture, and the Quantum Revival. New York: W. W. Norton and Company, 2011. Kauffman, L. A. “The Anti-Politics of Identity.” Socialist Review 20 (1) (1990): 67–80.

308

Bibliography

Kazemi, Asghar. “Quantum Politics: New Methodological Perspective.” EJournal of International Law and Political Science. February 13, 2011. Accessed October 15, 2016. http://scholarforum.blogspot.ca/2011/02/quantum-politics -new-methodological.html. Keller, Evelyn. Refiguring Life: Metaphors of Twentieth-Century Biology. New York: Columbia University Press, 1995. Kimmelman, Burt. “George Oppen’s Silence and the Role of Uncertainty in Post-war American Avant-Garde Poetry.” Mosaic 36 (2) (2003): 145–162. Kinch, Sean. “Quantum Mechanics as Critical Model: Reading Nicholas Mosley’s Hopeful Monsters.” Studies in Contemporary Fiction 47 (3) (2006): 289– 308. Kobylinski,Wayne. “Getting to X: Paul Muldoon’s Quantum Poetics.” PhD diss., University of North Carolina, Chapel Hill, 2005. Proquest (3170473). Kojevnikov, Alexei. “Freedom, Collectivism, and Quasiparticles: Social Metaphors in Quantum Physics.” Historical Studies in the Physical and Biological Sciences 29 (2) (1999): 295–331. Kumar, Manjit. Quantum: Einstein, Bohr and the Great Debate About the Nature of Reality. New York: W. W. Norton and Company, 2008. Lakoff, George. “The Contemporary Theory of Metaphor.” In The Cognitive Linguistics Reader, ed.Vyvyan Evans, Benjamin Bergen, and Jörg Zinken, 264–316. London: Equinox Publishing, 2007. Lakoff, George. Women, Fire, and Dangerous Things: What Categories Reveal About the Mind. Chicago: University of Chicago Press, 1987. Lakoff, George, and Mark Johnson. Metaphors We Live By. Chicago: University of Chicago Press, 1980. Lakoff, George, and Mark Turner. More Than Cool Reason: A Field Guide to Poetic Metaphor. Chicago: University of Chicago Press, 1989. Laurence, William L. Dawn over Zero: The Story of the Atomic Bomb. New York: Knopf, 1946. Laurence, William L. Men and Atoms: The Discovery, the Uses and the Future of Atomic Energy. New York: Simon and Schuster, 1959. Laurence, William L. “Paradise or Doomsday?” Woman’s Home Companion, May 1948, 33. Law, Sung Ping. “The Regulation of Menstrual Cycle and Its Relationship to the Moon.” Acta Obstetricia et Gynecologica Scandinavica 65 (1) (1986): 45–48. Leane, Elizabeth. “Knowing Quanta: The Ambiguous Metaphors of Popular Physics.” Review of English Studies 52 (207) (2001): 411–431.

Bibliography

309

Leo, Vincent Anthony. “The Mushroom Cloud: From Photographic Fact to Cultural Symbol.” MA thesis, Ohio State University, 1984. Accessed May 16, 2014. https://www.ohiolink.edu/. Letherdale, W. H. The Role of Analogy, Model and Metaphor in Science. Oxford: North Holland Publishing Company, 1974. Lewis, Flora. “The Quantum Mechanics of Politics.” New York Times, November 6, 1983, 98. Lightman, Allen P. “Magic on the Mind: Physicists’ Use of Metaphor.” American Scholar 58 (1) (2001): 97–101. Locke, John. An Essay Concerning Human Understanding, ed. Peter H. Nidditch. Oxford: Clarendon Press, 1975. Lukacs, John. “Quantum Mechanics and the End of Scientism.” In Science as Metaphor: The Historical Role of Scientific Theories in Forming Western Culture, ed. Richard Olson, 293–300. Belmont: Wadsworth Publishing Company, 1971. Lundeen, Jeff S., Brandon Sutherland, Aabid Patel, Corey Stewart, and Charles Bamber. “Direct Measurement of the Quantum Wavefunction.” Nature 474 (2011): 188–191. Lyons, Jack. Quantum Touch. Accessed March 13, 2016. https://www .quantumtouch.com/en/instructors/sch-instructors/instructor-835. MacKinnon, Edward M. “The Rise and Fall of the Schrodinger Interpretation.” In Studies in the Foundations of Quantum Mechanics, ed. Patrick Suppes, 1–59. Michigan: Philosophy of Science Association, 1980. MacKinnon, Edward M. Scientific Explanation and Atomic Physics. Chicago: The University of Chicago. Press, 1982. MacLelland, Bruce. Prosperity Through Thought Force. 1907. Accessed June 25, 2015. http://www.brainybetty.com/2007Motivation/Bruce%20MacLelland %20-%20Prosperity%20Through%20Thought.pdf. The Manhattan Project Heritage Preservation Association, Inc. Einstein/Sachs Document Collection. Accessed May 15, 2014. http://www.mphpa.org/. Maso, Carole. Defiance. New York: Plume, 1999. McCance, Dawne C. “Physics, Buddhism, and Postmodernism.” Zygon 21 (3) (1986): 287–296. McGrail, Anna. Mrs. Einstein. New York: W. W. Norton & Company, 1998. McGregor, H. R. Schrödinger’s Baby. Lake Havascu City: London Bridge Publishers, 1998. McNay, Michael. “Blinding Light of Experience.” The Guardian, November 1981.

310

Bibliography

Mikkonen, Kai. “The ‘Narrative Is Travel’ Metaphor: Between Spatial Sequence and Open Consequence.” Narrative 15 (3) (2007): 286–305. Miller, Arthur I. Albert Einstein’s Special Theory of Relativity: Emergence (1905) and Early Interpretation (1905–1911). New York: Springer, 1998. Miller, Arthur I. Imagery in Scientific Thought: Creating 20th Century Physics. Boston: Birkhäuser, 1984. Miller, Arthur I. “Visualization Lost and Regained:The Genesis of the Quantum Theory in the Period 1913–1927.” In On Aesthetics in Science, ed. Judith Wechsler, 73–102. Boston: Birkhäuser, 1988. Moore, Walter. Schrödinger: Life and Thought. Cambridge, UK: Cambridge University Press, 1992. Murdoch, Dugald R. Niels Bohr’s Philosophy of Physics. Cambridge, UK: Cambridge University Press, 1989. Murphy, Gregory. “On Metaphoric Representation.” Cognition 60 (1996): 173–204. “The Mushroom Cloud.” Time, July 17, 1950, 66. Musial, Joe. Learn How Dagwood Splits the Atom. New York: King Features Syndicate, 1949. Nadeau, Robert. Readings from the New Book of Nature: Physics and Metaphysics in the Modern Novel. Amherst: University of Massachusetts Press, 1981. Nagel, Bengt. “Discussion of J. S. Bell’s Paper ‘Six Possible Worlds of Quantum Mechanics.’” Possible Worlds in Humanities, Arts, and Sciences: Proceedings of Nobel Symposium 65 (1988): 388–396. Newton, Isaac. The Mathematical Principles of Natural Philosophy, Book 3. London: H. D. Symonds, 1803. Nicolson, Marjorie Hope. “Sublime in External Nature.” In Dictionary of the History of Ideas: Studies of Selected Pivotal Ideas.Vol. 4, ed. Philip P. Wiener, 333–337. New York: Charles Scribner’s Sons, 1968. Norris, Christopher. Quantum Theory and the Flight from Realism: Philosophical Responses to Quantum Mechanics. New York: Routledge, 2002. O’Hara, J. D. “Donald Barthleme, The Art of Fiction No. 66.” Paris Review, no. 80, 1981. Accessed April 25, 2014. http://www.theparisreview.org/interviews/ 3228/the-art-of-fiction-no-66-donaldbarthelme. “One Victory Not Yet Won.” New York Times, August 12, 1945. Oppenheimer, J. Robert. “The Atom as a Great Force for Peace,” New York Times Magazine, June 9, 1946. Oppenheimer, J. Robert. “Niels Bohr and Atomic Weapons,” New York Review of Books, December 1964, 17.

Bibliography

311

Oppenheimer, J. Robert. “Physics in the Contemporary World.” Arthur D. Little Memorial Lecture at M.I.T., November 25, 1947. Oppenheimer, J. Robert. Science and the Common Understanding. Oxford: Oxford University Press, 1954. Ortony, Andrew. “Are Emotion Metaphors Conceptual or Lexical?” Cognition and Emotion 2 (1988): 95–103. Osborne, Michael M. “Metaphor.” In Encyclopedia of Communication Theory, ed. Stephen W. Littlejohn and Karen A. Foss, 654–657. Beverly Hills: Sage, 2009. Pais, Abraham. The Genius of Science: A Portrait Gallery. Oxford: Oxford University Press, 2000. Pais, Abraham. Inward Bound: Of Matter and Forces in the Physical World. Oxford: Oxford University Press, 1988. Papin, Liliane. “This Is Not a Universe: Metaphor, Language, and Representation.” Publications of the Modern Language Association of America 107 (5) (1992): 1253–1265. Patton, C. M., and John Wheeler. “Is Physics Legislated by Cosmology?” In Quantum Gravity: An Oxford Symposium, ed. C. J. Isham, Roger Penrose, and D. W. Sciama, 538–605. Oxford: Clarendon, 1975. Pauli, Wolfgang. Scientific Correspondence with Bohr, Einstein, Heisenberg, ed. Alan Herman, Karl Von Meyan, and Victor F. Weisskopf. Berlin: Springer Verlag, 1985. Pauli to Niels Bohr. December 12, 1924. Accessed September 11, 2016. http:// cds.cern.ch/search?cc=Pauli+Letter+Collection&ln=fr&jrec=11&p=1924. Penrose, Roger. “Gravitational Collapse and Space-Time Singularities.” Physical Review Letters 14 (3) (1965): 57–59. Pera, M., and W. R. Shea, eds. Persuading Science: The Art of Scientific Rhetoric. Canton, MA: Science History Publications, 1991. Pera, Marcello. The Discourses of Science, trans. C. Botsford. Chicago: University of Chicago Press, 1994. Peres, Asher. “How the No-Cloning Theorem Got Its Name.” Fortschritte der Physik 51 (45) (2003): 458–461. Pitts, Mary Ellen. “The Holographic Paradigm: A New Model for the Study of Literature and Science.” Modern Language Studies 20 (4) (1990): 80–89. Plotinsky, Arkady. Complementarity: Anti-Epistemology after Bohr and Derrida. Durham: Duke University Press, 1994. Prelli, Lawrence J. A Rhetoric of Science: Inventing Scientific Discourse. Columbia: University of South Carolina Press, 1989.

312

Bibliography

Pulaczewska, Hanna. Aspects of Metaphor in Experience: Examples and Case Studies. Berlin: Walter de Gruyter Press, 1999. Pullman, Bernard. The Atom in the History of Human Thought. Oxford: Oxford University Press, 1998. Putnam, James A. “Human Intelligence.” 2003. Accessed November 25, 2016. http://newphysicstheory.com/New%20Physics%20Theory%202014/Human _Intelligence.pdf. “Quantum Prosperity Program.” Accessed September 11, 2014. http://www .quantumprosperityprogram.com (site discontinued). “The Quantum Society.” KirkusReviews.com, February 15, 1994. Accessed May 28, 2014. https://www.kirkusreviews.com/book-reviews/danah-zohar/ the-quantum-society/. Restivo, Sal. “Parallels and Paradoxes in Modern Physics and Eastern Mysticism—A Critical Reconnaissance.” Social Studies of Science 8 (2) (1978): 143–181. Rhodes, Richard. The Making of the Atomic Bomb. New York: Simon and Schuster, 1986. Richards, I. A. The Philosophy of Rhetoric. London: Oxford University Press, 1935. Rotella, Guy. “Comparing Conceptions: Frost and Eddington, Heisenberg, and Bohr.” American Literature 59 (2) (1987): 167–189. Rowland, Antony. “Silence and Awkwardness in Nuclear Discourse.” English 43 (176) (1994): 139–150. Rutherford, Ernest. Speech to the British Association for the Advancement of Science, London, 1932. Reprinted as “Breaking Down the Atom,” The Times, September 12, 1933. Ryan, Vanessa L. “The Physiological Sublime: Burke’s Critique of Reason.” Journal of the History of Ideas 62 (1) (2001): 265–279. Saffo, Paul. “The Road from Trinity: Reflections on the Atom Bomb.” Accessed May 16, 2014. http://www.saffo.com/essays/the-road-from-trinity-reflections -on-the-atom-bomb/. Schrödinger, Erwin. “Are There Quantum Jumps? Part I.” British Journal for the Philosophy of Science 3 (10) (1952): 109–123. Schrödinger, Erwin. “Are There Quantum Jumps? Part II.” British Journal for the Philosophy of Science 3 (10) (1952): 233–242. Schrödinger, Erwin. “Discussion of Probability Relations Between Separated Systems.” Proceedings of the Cambridge Philosophical Society 31 (1935): 555–563; 32 (1936): 446–451.

Bibliography

313

Schröedinger, Erwin. “La signification de la mecanique ondulatoire.” In Louis de Broglie, Physicien at Penseur, ed. André George, 2–32. Paris: Albin Michel, 1953. Schrödinger, Erwin. “On the Relation Between the Quantum Mechanics of Heisenberg, Born, and Jordan, and that of Schrödinger.” In Collected Papers on Wave Mechanics, ed. W. M. Deans and J. F. Shearer, trans. J. F. Shearer, 45–61. London: Blackie and Son, 1928. Schrödinger, Erwin. “Perturbation Theory, With Applications to the Stark Effect of Balmer Lines.” In Collected Papers on Wave Mechanics, ed. W. M. Deans and J.F. Shearer, trans. J. F. Shearer, 62–101. London: Blackie and Son, 1927. Schrödinger, Erwin. “Quantisation as a Problem of Proper Values: Part 1.” In Collected Papers on Wave Mechanics, ed. W. M. Deans and J.F. Shearer, trans. J. F. Shearer, 1–12. London: Blackie and Son, 1927. Schrödinger, Erwin. “Quantisation as a Problem of Proper Values: Part 2.” In Collected Papers on Wave Mechanics, ed. W. M. Deans and J.F. Shearer, trans. J. F. Shearer, 45–61. London: Blackie and Son, 1927. Schrödinger, Erwin. “Quantisation as a Problem of Proper Values: Part 3.” In Collected Papers on Wave Mechanics, ed. W. M. Deans and J.F. Shearer, trans. J. F. Shearer, 62–101. London: Blackie and Son, 1927. Schrödinger, Erwin. “Quantisation as a Problem of Proper Values: Part 4.” In Collected Papers on Wave Mechanics, ed. W. M. Deans and J.F. Shearer, trans. J. F. Shearer, 102–123. London: Blackie and Son, 1927. Schrödinger, Erwin. “Quantisierung als Eigenwertproblem (Erste Metteilung).” Annalen der Physik 384 (4) (January 1926): 361–376. Schrödinger, Erwin. Science, Theory, and Man. Crow’s Nest, AU: G. Allen and Unwin, 1957. Schrödinger, Erwin. “Über das verhältnis der Heisenberg-Born-Jordanschen quantenmechanik zu der meinem.” Annalen der Physik 384 (8) (1926): 734–756. Schrödinger, Erwin. “An Undulatory Theory of the Mechanics of Atoms and Molecules.” Physical Review 28 (6) (1926): 1049–1070. Schrödinger, Erwin. What Is Life? Mind and Matter. Cambridge, UK: Cambridge University Press, 2010 [1944]. Science Legislation: Hearings before the Subcommittee of the Committee on Military Affairs, U.S. Senate, 79th Congress, 1st Session. Washington, DC: U.S. Government Printing Office, 1945. “Scientist Tells of Einstein's A-bomb Regrets.” Philadelphia Bulletin, May 13, 1955. Shils, Edward. “Freedom and Influence: Observations on the Scientists’ Movement in the United States.” Bulletin of the Atomic Scientists 13 (1) (1957): 13–18.

314

Bibliography

Shimony, Abner. “Review of Henry Folse, The Philosophy of Niels Bohr: The Framework of Complementarity.” Physics Today 38 (10) (1985): 108–109. Siegel, Deborah. Sisterhood, Interrupted: From Radical Women to Girls Gone Wild. New York: Palgrave Macmillan, 2007. Sillanpää, Maria. “A New Deal for Sustainable Development in Business: Taking the Social Dimension Seriously at The Body Shop.” In Sustainable Measures: Evaluation and Reporting on Environmental and Social Performance, ed. Martin Benneti and Peter James, 529–551. Sheffield: Greenleaf Publishing, 1999. Smith, Jason C. “Schrödinger’s Cat and Sarah’s Child: John Fowles’s ‘Quantum’ Narrative.” Mosaic 32 (2) (1999): 91–106. Smith, Victoria. “‘Ransacking the Language’: Finding the Missing Goods in Virginia Woolf ’s Orlando.” Journal of Modern Literature 29 (4) (2006): 57–75. Sokal, Alan.“Transgressing the Boundaries:Toward a Transformative Hermeneutics of Quantum Gravity.” Social Text 46/47 (1996): 217–252. Sokal, Alan, and Jean Bricmont. Fashionable Nonsense: Postmodern Intellectuals’ Abuse of Science. New York: Picador USA, 1998. Solomon, James Fisher. Discourse and Reference in the Nuclear Age. London: University of Oklahoma Press, 1988. “The Sound of the Atom.” Newsweek, July 10, 1950, 50–51. Stenger, Victor J. “The Myth of Quantum Consciousness.” Humanist 53 (3) (1992): 13–15. Stenger,Victor. J. “Quantum Quackery.” January/February 1997. Accessed May 15, 2014. http://www.csicop.org/si/show/quantum_quackery. Strehle, Susan. Fiction in the Quantum Universe. Chapel Hill: University of North Carolina Press, 1992. Swing, Raymond G. “Commentary on the Atomic Age.” Radio broadcast. Mutual Broadcasting System, August 13 1945. Taylor, Barbara Brown. “Physics and Faith: The Luminous Web.” Christian Century (June 1999): 612–619. Tompkins, Penny, and James Lawley. “Embodied Schema: The Basis of Embodied Cognition.” Notes first presented at The Developing Group, June 6, 2009. Accessed November 30, 2016. http://www.cleanlanguage.co.uk/articles/ articles/245/1/Embodied-Schema-The-basis-of-Embodied-Cognition/ Page1.html. Truman, Harry S. “Statement by the President of the United States.” Speech to the American public, Washington, DC. August 6, 1945. Accessed May 16, 2014. https://www.trumanlibrary.org/publicpapers/index.php?pid=100&st=sun+dra ws&st1=.

Bibliography

315

“Truman Declares Atomic Bomb Will Hasten War’s End.” Globe and Mail, August 7, 1945. “The Uncertainty Principle.” Stanford Encyclopedia of Philosophy, modified, July 3, 2006. Accessed April 21, 2015. http://plato.stanford.edu/entries/qt -uncertainty/. Ungar, Sheldon. The Rise and Fall of Nuclearism: Faith and Fear as Determinants of the Arms Race. Philadelphia: Pennsylvania State University Press, 1992. “War’s Ending: Atomic Bomb and Soviet Entry Bring Jap Surrender Offer.” Life, August 20, 1945, 30–31. Waymack, W. W. “A Letter to Judith and Dickie.” National Education Association Journal 36 (1947): 214. Weart, Spencer R. “The Heyday of Myth and Cliché.” Bulletin of the Atomic Scientists 41 (7) (1985): 8–43. Weart, Spencer R. Nuclear Fear: A History of Images. Cambridge, MA: Harvard University Press, 1988. Weisskopf,Victor. The Joy of Insight. New York: Basic Books, 1991. Werner Heisenberg to Wolfgang Pauli, 8 June, 1926. In Scientific Correspondence with Bohr, Einstein, Heisenberg, ed. Alan Herman, Karl Von Meyan, and Victor F. Weisskopf. Berlin: Springer, 1985, 26–35. Werner Heisenberg to Wolfgang Pauli, 23 February, 1927. In Scientific Correspondence with Bohr, Einstein, Heisenberg, ed. Alan Herman, Karl Von Meyan, and Victor F. Weisskopf. Berlin: Springer, 1985, 26–35. Wertsche, J.V. “Nuclear Discourse.” Communication Research 14 (1987): 131–138. “What Does Anschauung Mean?” The Monist 2 (4) (1892). Accessed June 18, 2016. http://www.jstor.org/stable/27897002. What the Bleep Do We Know? 2004. Directed by Betsy Chasse, Mark Vicente, and William Arntz. United States: Fox Video. DVD. Whitworth, Michael H. Einstein’s Wake: Relativity, Metaphor and Modernist Literature. Oxford: Oxford University Press, 2001. Wigner, Eugene. “Remarks on the Mind-Body Problem.” In The Scientist Speculates, ed. I. J. Good. London: William Heinemann, Ltd., 1961; New York: Basic Books, Inc., 1962. Accessed March 14, 2017. http://www .infor mationphilosopher.com/solutions/scientists/wigner/Wigner _Remarks.pdf. Wigner, Eugene. Symmetries and Reflections: Scientific Essays. Cambridge, MA: MIT Press, 1970. Wigner, Eugene. “The Unreasonable Effectiveness of Mathematics in the Natural Sciences.” Communications on Pure and Applied Mathematics 13 (1) (1960): 1–14.

316

Bibliography

Wittner, Lawrence S. One World or None: A History of the World Nuclear Disarmament Movement Through 1953. Vol. 1. Redwood City: Stanford University Press, 1997. Woolf,Virginia. To the Lighthouse. Orlando: Harcourt and Brace, 1927. Wright, Will. Wild Knowledge: Science, Language, and Social Life in a Fragile Environment. Minneapolis: University of Minnesota Press, 1992. Yom, Sue Sun. “Biography and the Quantum Leap:Waves, Particles, and Light as a Theory of Writing the Human Life.” In Virginia Woolf: Texts and Contexts, ed. Beth Rigel Daugherty and Eileen Barrett, 145–150. New York: Pace University Press, 1996. Zinken, Jörg. “Introduction to Part Four: Metaphor, Metonymy and Blending.” In The Cognitive Linguistics Reader, ed. Vyvyan Evans, Benjamin K. Bergen, and Jörg Zinken, 263–267. London: Equinox Publishing, 2007. Zohar, Danah. “Forces of Reaction.” Sunday Times, February 6, 1994. Zohar, Danah, and Ian Marshall. Quantum Society: Mind, Physics, and a New Social Vision. London: Harper Perennial, 1995. Zukav, Gary. The Dancing Wu Li Masters: An Overview of the New Physics. New York: William Morrow and Company, 2001 [1979].

INDEX

Across the Frontiers (Heisenberg), 118, 296n9 “Actual Content of Quantum Theoretical Kinematics and Mechanics,The” (Heisenberg), 102 AEC, 237 Aesthetic preferences in quantum physics, 2, 13–15, 67–68, 108–123 Albright, Daniel, 134–135, 281n30 American exceptionalism, 249–251. See also Nuclear discourse Annalen der Physik (Schrödinger), 80 Anschaulichkeit. See Visualizability in quantum physics Anschauung. See Intuition in quantum physics Anthony, Robert, 203–205 “Anthropic Universe,The” (interview), 182 Anthropomorphism, 13, 22–24, 29, 49–54, 67, 198, 201, 232–240. See also Metaphors, quantum Appropriation of quantum concepts, 15–16, 19–25, 169, 258–261 “Are There Quantum Jumps? Part I” (Schrödinger), 58, 111–112 “Are There Quantum Jumps? Part II” (Schrödinger), 84 Asgari,Tayebeh, 31 Ashvaghosha, 174 Aspect, Alain, 265n1

Associated Press, 227 Athenaeum (journal), 127 “Atom: Death—or the Life Abundant, The” (Foulkes), 245 Atomic Adventures (textbook), 236 Atomic bomb, 2, 22–25, 211, 293nn51–52, 296n8. See also Nuclear discourse “Atomic Crusade and Its Social Implications,The” (Compton), 225 Atomic Energy Commission (AEC), 237 Atomics for the Millions (textbook), 236 Atomic Story,The (Campbell), 237 “Atomic Theory and Mechanics” (Bohr), 87 Atom in the History of Human Thought,The (Pullman), 49 Atomism, 3, 20–21, 187, 208 “Atoms and Human Knowledge” (Bohr), 63 “Atoms for Peace” (Eisenhower), 249–250 “At the Loom” (Duncan), 151–152 Aurobindo, Sri, 172 Barthe, Robert, 149 Barthelme, Donald, 143–144 Baruch, Bernard, 239, 246 Battle of Argonne, 231 Bazerman, Charles, 27 Becker,Ted, 194 Beer, Gillian, 127, 130

318 Bell, David, 183 Bell, J. S., 155, 183–185, 286n40 Beller, Mara, 71–72, 79–85 Bell’s nonlocality theorem, 183–185, 286n40 Benham, Renee, 134 Bergson, Henri, 129 Between the Acts (Woolf), 127 Bhagavad Gita, 225 “Bicephalous Writer:The Commingling of the Creative Writer and the Critic in a Single Body,The” (Carrion), 144 Bikini Atoll, 232, 247–248 Billiard ball metaphor, 37–38 Black Mountain poets, 141 Body Artist,The (DeLillo), 156 Body in the Mind:The Bodily Basis of Meaning, Imagination, and Reason,The (Johnson), 31 Bohm, David, 161–165, 171, 183, 191, 220, 265n5 Bohnenkamp,Thomas, 128 Bohr, Niels aesthetic preference and, 14–15, 74, 108–109, 113–117, 122 intuition and, 14, 72–73, 92–93, 96–99, 121–122 language and, 2–13, 27, 130–131, 136–140, 147, 172–173, 215, 229, 233, 243, 257–259 visualizability and, 76–91, 121–122 Booker, Keith, 149–151 Born, Max, 5–8, 13, 35–37, 56, 70–73, 79–83, 91, 101, 121–122, 195 Bothe,Walther, 103 Bothe-Geiger experiments, 103 Boyer, Paul, 242, 250 Bricmont, Jean, 158 British Association for the Advancement of Science, 212 Brown, Paul Tolliver, 131 Brown,Theodore, 54 Buddhism, 174, 178–179, 187, 191

Index Bureau of Standards, 223 Burke, Edmund, 218–219 Butterfly effect, 193 By the Bomb’s Early Light (Boyer), 242 Camilleri, Kristian, 46, 107 Campbell, Jeremy, 198 Campbell, John W., 237 “Can Quantum-Mechanical Description of Physical Reality Be Considered Complete?” (Bohr), 115 Capra, Fritjof, 20, 169–182, 191, 197–201, 208 Carnival Trilogy,The (Harris), 138–139 Carrion, Jorge, 144 Carroll, Lewis, 256 Carter, Steven, 151–154 Cassirer, Ernst, 113 Causality appropriation and, 177, 198–199, 289n84 classical physics and, 34–41, 46–47, 53 language and, 97, 107, 155, 241, 265n4, 266n12 literary criticism and, 17, 19, 127–128, 144, 154–155, 160 quantum physics and, 62–69, 265n4, 266n12 “Causality and Complementarity” (Bohr), 35, 90 Celebrity nuclearism, 24, 230–232, 293n52. See also Nuclear discourse Chaos theory, 193 Chevalley, Catherine, 113–115, 278n116 Chilton, Paul, 215, 220, 224 Chopra, Deepak, 199–201, 206, 289n84 Churchill,Winston, 222 Clark, Miriam, 130–131, 280n22 Classical physics aesthetic preference and, 15, 109–117, 121–122 appropriation and, 194–196 emergence of, 5–7

Index intuition and, 14, 73, 93–102, 106–108, 121–122 language and, 29–42, 46–67, 257–258, 265n4 literary criticism and, 131, 135–136, 146–147 visualizability and, 75–88, 121–122, 226, 229 Clauser, John, 184 Clement, James, 231 Coale, Samuel, 156–158 Cockcroft, John, 212 Coexistent potentialities, 44 Cold War, 246 Coletta,W. John, 136–137 Commodification of quantum logic, 21, 202–209, 259 “Comparing Conceptions: Frost and Eddington, Heisenberg and Bohr” (Rotella), 144, 281n41 COMPATIBILITY OF IDEAS IS THE COMPATIBILITY OF SHAPES image schema, 61–65 Compton, Arthur, 221, 225–226 Conceptual Development of Quantum Physics, The (Jammer), 69 Conceptual Metaphor Theory, 11–12, 28–42, 49–54, 66, 146–147, 215–216 “Consciousness, Stream and Quanta in To the Lighthouse” (Clark), 130 “Contemporary Theory of Metaphor, The” (Lakoff), 54 Cooke, Morris L., 223 Cooper, Gary, 231 Copenhagen debates, 1926, 39–40, 78, 84, 87, 269n31 Copenhagen-Göttingen group, 7–8, 70, 81–87, 90, 93, 272n5 Copenhagen-Göttingen Interpretation, 7–8, 70, 81–87, 93, 272n5 Copenhagen Interpretation, 7, 265n5. See also Copenhagen-Göttingen Interpretation

319 “Copenhagen Interpretation of Quantum Theory,The” (Heisenberg), 265n5 Corpuscular model of light, 3–4 Cosmic oneness, 171, 220. See also Quantum consciousness Counter-Memorial Impulse in Twentieth-Century English Fiction,The (Henstra), 215 Cousins, Norman, 247–248 Creeley, Robert, 141–142 Creolization, 138–140 Cubism, 114 Cuffaro, Michael, 96 Cultural studies, 16 Dancing Wu Li Masters,The (Zukav), 169, 171, 175–176, 182, 186 Dawn over Zero (Laurence), 245 de Broglie, Louis, 4, 75, 80–83 Decision to Drop the Bomb,The (documentary), 225, 292n39 Deconstructionism, 12, 18, 129 de Gongóra, Louis, 112 DeLillo, Don, 127, 156–157 Dennis, John, 217–219 de Regt, Hank, 75–79, 94, 105–106 Derrida, Jacques, 145–147, 240–241 Descartes, René, 190–191 Determinism, 6, 10, 150–153, 192, 194, 258. See also Indeterminacy “Development of Concepts in the History of Quantum Theory” (Heisenberg), 44 DeWitt, Bryce, 8–9 Dieks, Dennis, 287n42 Dirac, Paul, 2, 72, 77, 81, 91, 105 Discontinuous theory of matter, 70 Discrete Series (Oppen), 137 “Discussion with Einstein on Epistemological Problems in Atomic Physics” (Bohr), 47–48 Dispenza, Joe, 198

320 Distance healing, 2, 20–21, 25, 169, 196–203, 207–208, 220, 259–260. See also Appropriation of quantum concepts Double-slit experiment, 3–5, 134, 176, 189–190, 204, 265n1 Duality. See Nuclear discourse; Subject/ object duality;Wave/particle duality Duncan, Robert, 151–152 DuPlessis, Rachel Blau, 138 Eastern mysticism, 2, 20, 170–180, 187, 202, 206–207, 219–220, 259–261, 265n4. See also Appropriation of quantum concepts Eddington, Arthur, 127, 130, 147, 256, 296n9 Edie, James, 27 Edlund, Rebekka, 138–139, 166 Einstein, Albert, 4, 9–10, 39, 62, 69–70, 77, 80, 87, 102–104, 115–118, 154, 183–184, 223–224, 266n12, 283n84, 286n40 Einstein-Podolsky-Rosen paper, 9–10, 115–116, 183 Eisenhower, Dwight D., 249–250 Enfoldment, 161–165 Enola Gay, 230–231, 293n51 Entanglement, 10, 16–19, 21, 154–160, 165, 182–183, 204, 214, 259, 286n40 Entrainment, 202 Entropy, 52, 233 EPR paradox, 9–10 Erhard,Werner, 182 “Eternal Apprentice,The” (article), 292n39 Ettinger, Ian, 130, 133, 154–158 Evans,Vyvyan, 28, 32 Everett, Hugh, 8, 149 Exclusion principle, 51 “Experimental Evidence for a Photon Anticorrelation Effect on a Beam Splitter: A New Light on Single-photon

Index Interferences” (Aspect, Grangier, Roger), 265n1 Familiarity-view of understanding, 105 Farrell,Thomas, 218–219 Fashionable Nonsense: Postmodern Intellectuals’ Abuse of Science (Bricmont), 158 Feminism, 129, 197 Fenollosa, Ernest, 147 Fermi, Enrico, 212 Feynman, Richard, 263 “Fluctuation Phenomena and Quantum Mechanics” (Heisenberg), 84 Flying Tigers (film), 231 “Forces of Reaction” (Zohar), 187–189 Forman, Paul, 69, 76, 101 “For Once,Then, Something” (Frost), 136 Foulkes,William, 245 Foundation metaphor, 61–65 Fowles, John, 150 Framework metaphor, 61–65 Freed, Fred, 292n39 Freire, Olival, Jr., 70, 266n5 French Lieutenant’s Woman,The (Fowles), 150, 152 Frequency model, 84 “Fresh Fields” (Heisenberg), 59 Freud, Sigmund, 129 Frost, Robert, 136–137, 144 Froula, Christine, 145–148, 166, 241, 282n65 Fundamental Fysiks Group, 180–186, 287n42 Galton, Francis, 78 Gamson,William, 226, 245 Gannon, Charles, 215 Geiger, Hans, 103 Gender literary criticism and, 16–17, 131–134, 280n22 nuclear discourse and, 234–236, 293n51

Index Gestalt approach, 128–129 Get-rich schemes, 2, 20–21, 25, 196–208, 259–260. See also Appropriation of quantum concepts Gilda, 232 Gilder, Louisa, 156–157 Giovannitti, Len, 292n39 Goldman, Bert, 205–206 Gordon, Richard, 202 Goswami, Amit, 198 Göttingen University, 69. See also Copenhagen-Göttingen Interpretation Grangier, Philippe, 265n1 Groves, Leslie, 237–239 Haldane, J. B. S., 263, 296n9 Hall, Henry S., 223–224 Harmony in aesthetic preferences, 14–15, 74, 108–109, 115–121 Harris, J. D., 143 Harris,Wilson, 138 Harrison, George L., 235 Hayles, Katherine, 128 Hayworth, Rita, 232 Heisenberg,Werner aesthetic preferences and, 14–15, 74, 108–109, 117–123 intuition and, 14, 72–74, 92–93, 99–108, 121–122 language and, 2–13, 27, 142, 145–147, 150, 172–174, 214–215, 243, 255, 257–258 visualizability and, 13, 71, 74–88, 122 Henry, Holly, 127 Henstra, Sarah, 215 Herbert, Nick, 138, 287n42 Hertz, Heinrich, 113 Hindu mythology, 178 Hiroshima, atomic bombing of, 24, 215, 218, 227–236, 242, 248–250 Holism, 161–165, 191, 196, 207 Hollywood, 231–232 Holographic model, 161–164

321 Hook, Glenn, 235 Hopeful Monsters (Mosley), 158–160, 165 Howard, Don, 272n5 How Hippies Saved Physics (Kaiser), 183 Humanists, 126 Humanities, 17, 125, 158 Hussey, Mark, 163–165 Huygens, Christian, 3 Hypodermic model, 127 Image schemas, 28–40, 66 Implicate order, 161–165, 171 Indeterminacy, 2, 85, 141–149, 153, 165–166, 257, 272n5. See also Determinism; Uncertainty Principle “Indeterminism in Physics” (Schrödinger), 36 Individualism. See Liberal individualism Indo-European languages, 12, 29, 41, 66, 262 Inquiries into Human Faculty and Its Development (Galton), 78 Institute for Theoretical Physics, 69 Interdisciplinarity, 16, 125 International Congress on Light Therapy, 116 “Introduction Survey” (Bohr), 31–32 Intuition in quantum physics, 2, 6, 13–14, 67–68, 72–74, 91–108, 121–123, 229 “Jabberwocky” (Carroll), 256 Jammer, Max, 69 Japan, atomic bombing of, 220–221, 224–231, 242, 247 Jeans, James, 127–128 Jobs, Steve, 184 Johnson, Mark, 28, 30–33, 38, 50, 233, 240 Jordan, Pascual, 5, 56–57, 70 Joyce, James, 144, 149–150 Kaiser, David, 183–186, 286n40 Kant, Immanuel, 92, 113, 278n116

322 Kazemi, Asghar, 192–193 Kepler, Johannes, 120 Kimmelman, Burt, 137, 141–143, 146, 148, 166 Kinch, Sean, 144, 158–160, 165–167 King Features Syndicate, 237–238 Kirkus Reviews, 192 Knowing particles, 53, 176, 201 Koan, 178–179 Kobylinski,Wayne, 142–143, 146, 148–151, 153 Kramers, Hans, 83 Kuhn,Thomas, 44–45, 81, 83, 102–103 Lakoff, George, 28–33, 37–38, 40, 50, 54, 233, 240 Language of classical physics, 3, 42–49, 65–67, 135–136. See also Classical physics Laurence,William, 218–219, 221–222, 225, 234, 243–247 Lawley, James, 28 Law of attraction, 289nn94–95 Lawrence, D. H., 135 Laws of motion, 39–40 Leane, Elizabeth, 216 Leo, Anthony, 227 Lewis, Flora, 194–195 Liberal individualism, 19–21, 170, 187, 190, 206 LIFE IS A JOURNEY image schema, 55–56 Life magazine, 221, 223, 227–229 Literary criticism through quantum concepts, 16–19, 125, 229, 257–258, 261 Lorentz, Hendrik, 91 Lyons, Jack, 202 MacArthur, Douglas, 231 Mackinnon, Edward, 75 MacLelland, Bruce, 289n95 Making Truth (Brown), 54

Index Manhattan Project, 2, 218, 221, 225, 237, 242 Many-Worlds Interpretation, 8, 149, 152–153, 205 Marshall, Ian, 189–190 Massless particle, 178–179 Mathematics of quantum physics, 56–57 Matrix mechanics, 5–6, 14, 40, 45, 51, 56–61, 68, 115, 123, 173, 265n4 “Matrix Theory before Schrödinger: Philosophy, Problems, Consequences” (Beller), 79 Maxwell, Jeremy Clerk, 47, 63 McCance, Dawne, 190–194 McNay, Michael, 235 “Media Discourse and Public Opinion on Nuclear Power: A Constructionist Approach” (Gamson, Modigliani), 245 Metaphors, quantum anthropic, 13, 29, 49–54, 67, 198, 201, 232–240 architecture, 13, 29, 61–65 journey, 13, 29, 54–61, 65, 243–244 limitations of, 28–46 of literary criticism, 17–19, 129, 258 misapplied, 259, 281n30 of mysticism, 175, 180, 188–191, 198–203 of nuclear discourse, 23–24, 215–217, 232–252, 262–263 of politics, 19–20, 170, 194–196, 258 Michelangelo, 219 Mikkonen, Kai, 29, 55, 243 Miller, Arthur I., 74–79, 82, 91–92 Modernism, 17–18, 126–136, 165 Modigliani, Andre, 226, 245 Monist,The (journal), 92 Moore,Walter J., 76–79 Mosley, Nicholas, 158–160 Muldoon, Paul, 142–143, 148–149, 165 Murdoch, Dugald R., 89 Murphy, James Gardner, 69 Mushroom clouds, 227–230, 245–246

Index Nadeau, Robert, 125–126, 128 Nagasaki, atomic bombing of, 24, 218, 227–228, 236, 242, 250 National Science Foundation, 223 “Natural Philosophy and Human Cultures” (Bohr), 32 New Age quantum philosophy, 20–21, 169, 219–220, 258 New Left, 197 Newsweek, 227 New Thought Spiritual Movement, 289n95 Newton, Isaac, 3–4, 19–21, 33–39, 44, 47, 63–67, 107, 119–121, 192 Newton’s laws of motion, 37–40, 64 New York Times, 242, 244 No-cloning theorem, 183, 287n42 Nominal language, 42–43, 66, 85, 147, 255, 257, 262. See also Noun-based language;Thingness Nonlocality, 2, 10, 18, 154–160, 182–186, 198–199, 266n12 Norris, Christopher, 193 Noun-based language, 12, 29, 41, 66, 147, 255, 257, 262. See also Nominal language;Thingness Nuclear arms race, 195, 246 Nuclear celebrity, 24, 230–232, 293n52 Nuclear discourse, 2, 22, 211, 262–263, 293nn51–52 Nuclear Fear: A History of Images (Weart), 293n52 Nuclearism, 22–24, 211 Nuclear proliferation, 246–249 OBJECT schema, 29, 36–38, 66 Observer effect, 8–9, 16–25, 148–154, 171–173, 177, 187–188, 193–194, 197, 214, 258–260, 266n6. See also Superposition;Wavefunction collapse Observer/participant interaction, 9, 20, 193, 204 O’Donnell, Patrick, 156

323 Olson, Charles, 141 “On the Perceptual Content of Quantum Theoretical Kinematics and Mechanics” (Heisenberg), 5 “On the Quantum Mechanics of Collisions” (Born), 82 “On the Relation Between the Quantum Mechanics of Heisenberg, Born, and Jordan, and that of Schrödinger” (Schrödinger), 93 “On the Relationship of the Heisenberg-Born-Jordan Quantum Mechanics to Mine” (Schrödinger), 80 Operation Crossroads, 247–248 Oppen, George, 137–138, 141–142 Oppenheimer, J. Robert, 174, 223–225, 292n39 Optics (Newton), 3 Orlando (Woolf), 133–134 Pagel, Heinz, 136 Panofsky, Erwin, 113 Papin, Liliane, 12, 27–29, 41, 66 Paradise (Barthelme), 143–144 “Paradise or Doomsday?” (Laurence), 246 Paris Review, 143 Participatory Anthropic Principle, 182 Participatory democracy. See Participatory politics Participatory democratic theory, 194 Participatory politics, 9, 19–21, 188–189, 194–196, 207, 260 Particulate theory of atomic matter, 6, 51–53, 81, 85–86 PATH schema, 29, 38–40 Pauli,Wolfgang, 2, 6, 13, 37, 41–43, 51, 67, 70–81, 85–87, 92, 105–106, 122, 171 Peat, F. David, 161 Philosopher’s stone, 243, 246 Philosophical Inquiry into the Origins of Our Ideas of the Sublime and the Beautiful, A (Burke), 218

324 Philosophy of Science (journal), 272n5 Photoelectric effect, 4 Photons, 4–5 Physical Principles of the Quantum Theory, The (Schrödinger), 105, 265n2 Physics. See Classical physics “Physics, Buddhism, and Postmodernism” (McCance), 190 Physics and Beyond: Encounters and Conversations (Heisenberg), 59, 84, 269n31 Physics and Philosophy:The Revolution of Modern Science (Heisenberg), 36, 41, 85, 102, 107, 258, 265n5 “Physics as Art:The German Tradition and the Symbolic Turn in Philosophy, History of Art and Natural Science in the 1920s” (Chevalley), 113 Physics-Consciousness Research Group, 182 Pitts, Mary Ellen, 162–164 “Place to Step Further: Jack Spicer’s Quantum Poetics, A” (Carter), 151 Planck, Max, 4, 62 Plato, 15, 74, 118, 120, 122 Podolsky, Boris, 9, 115–116, 154 Poetic form, 16, 136–140, 256–257 Poincaré, Henri, 77 Popper, Karl, 185 Possible Worlds and Other Papers (Haldane), 263, 296n9 Postcolonial studies, 16, 139, 244 Postmodernism, 12, 16–18, 129, 141–148, 156, 165, 191–193, 261 Post-New Age mysticism, 170, 197–208, 258–260 Post-New Age quantum practitioners, 196–207 Poststructuralism, 12, 16, 125, 144–145 Pound, Ezra, 135 Power,Tyrone, 231 Principle of Complementarity aesthetics and, 15, 114–116, 122

Index appropriation of, 20–21, 49, 195, 258–259 duality and, 244 emergence of, 5–7, 32, 35, 48, 57, 63–67, 265n4 intuition and, 92, 96–99 literary criticism and, 16, 130, 136–142, 166, 281n41 visualizability and, 89–91, 109 Probability distributions, 35, 41–44, 53, 66, 226 Probability theory, 5, 7, 38, 45, 56, 71, 81–84, 173–175, 255–256 Probability waves, 7, 35, 42–44, 82–84, 173–174, 195, 287n51 Prometheus, scientist as, 221–222 Prosperity Through Thought Force (MacLelland), 290n95 Pulaczewska, Hanna, 43 Pullman, Bernard, 49 Putnam, James A., 201–202, 206 Pythagoras, 74, 118 Quantum agent, 21, 37–38, 208, 272n5 Quantum consciousness, 9, 21, 161–164, 169–188, 196–200, 207–208, 220, 258 Quantum field theory, 128, 173 Quantum flux, 157, 284n91 Quantum get-rich schemes, 2, 20–21, 24, 203–208, 259–260. See also Appropriation of quantum concepts Quantum grammar, 40–46. See also Quantum language Quantum healing, 2, 20–21, 25, 169, 196–203, 207–208, 220, 259–260. See also Appropriation of quantum concepts Quantum interpretation, 1–15, 56–68, 70–71, 75 Quantum jumping, 8, 13, 38–39, 44, 54–55, 67, 81, 84–87, 103, 111–112, 169, 205, 259 Quantum language appropriation of, 19–22, 169, 259

Index intuition and, 73, 100 limitations of, 10–13, 27, 73, 86, 98–100, 106–108, 113–117, 255–258, 269n44 in literary criticism, 16–19, 125 Quantum mechanical theory, 80 “Quantum Mechanics of Politics,The” (Lewis), 194 Quantum mysticism, 20, 23, 169 Quantum packets, 4 “Quantum Physics and Philosophy: Causality and Complementarity” (Bohr), 89 “Quantum Physics/Postmodern Metaphysics:The Nature of Jacques Derrida” (Froula), 145 Quantum politics, 20, 24, 170, 186–196, 208, 260–261 Quantum Politics (Becker), 194 “Quantum Postulate and the Recent Developments of Atomic Theory,The” (Bohr), 53, 56, 85, 89, 115 Quantum self-help, 170, 196–207 Quantum Society: Mind, Physics, and a New Social Vision (Marshall, Zohar), 189, 192 Quantumstuff, 138–139 “Quantum-theoretical Reinterpretation of Kinematic and Mechanical Relations” (Heisenberg), 76 Quantum Theory and the Flight from Realism: Philosophical Responses to Quantum Mechanics (Norris), 193 Quantum touch, 169, 202 Quantum Touch (website), 202–203 Rauscher, Elizabeth, 180 Reading from the Book of Nature, Metaphysics, and the Novel (Nadeau), 125–126 “Relativity and Quantum Theory in Virginia Woolf ’s The Waves” (Ettinger), 130 Religious language in nuclear discourse, 217–226, 262–263

325 “Remarks on the Mind-Body Problem” (Wigner), 181 “Representation of Nature in Contemporary Physics,The” (Heisenberg), 60 Resonance, 202 “Resonance, Life-Force, and the Principles of Quantum-Touch” (Gordon), 202 Restivo, Sal, 175 Richards, I. A., 215–216 “Robert Duncan and Erwin Schrödinger: Esthetics of the Wavefunction” (Carter), 151 Roger, Gérard, 265n1 Rosen, Nathan, 9, 115–116, 154 Rotella, Guy, 144, 281n41 Rowland, Antony, 234 Rutherford, Ernest, 212 Saffo, Paul, 221 Sarfatti, Jack, 182 Saturday Evening Post,The, 127 Schrödinger, Erwin. See also Wave mechanics aesthetic preference and, 14–15, 74, 108–113 intuition and, 14, 72–73, 92–96, 105–108, 121–122 language and, 2–13, 27, 147, 155, 229, 233, 243, 257 mysticism and, 171 visualizability and, 13, 71–72, 75–88, 121–122, 226 Schrödinger’s cat, 8, 151–152, 266n7, 286n30 “Science, Art, and Play” (Schrödinger), 57 Science, Order, and Creativity (Bohm, Peat), 161 Science,Theory, and Man (Schrödinger), 50, 52, 109 “Science and Exile” (Freire), 265n5

326 Scientific Explanation and Atomic Physics (MacKinnon), 75 Second wave feminism, 197 Sendak, Maurice, 236 Sense perception, 10–11, 23, 90 Sergeant York (film), 231 Shakespeare,William, 256 Ships with Wings (film), 231 “Significance of Wave Mechanics,The” (Schrödinger), 83 “Simplicity and Purposefulness in the Arts and Crafts” (Schrödinger), 109–110 Simplicity in aesthetic preferences, 14–15, 74, 109–110, 117–121 Slater, John, 83 Smith, Jason C., 150, 152–153, 283n82 Smith,Victoria, 134 Social Text (journal), 158 Sokal, Alan, 158, 261 Solomon, James, 145, 148 Sommerfeld, Arnold, 101 SOURCE-PATH-GOAL schema, 38–39, 67 Space. See Spatial orientations;Temporal/ spatial relations Spacetime, 35, 47, 53, 62–64, 67, 80–87, 97, 116, 199, 265n4. See also Spacetime continuum;Temporal/spatial relations Spacetime continuum, 62, 70, 93, 213, 265n4. See also Spacetime;Temporal/ spatial relations “Spacetime Visualisation and the Intelligibility of Physical Theories” (de Regt), 75, 105 Spatial orientations, 11, 32–36. See also Temporal/spatial relations Spatz,Tooey, 229 Stapp, Henry, 182 Stenger,Victor, 172–173, 188 Stinson, Henry L., 235 Strehle, Susan, 143–144 Subatomic particles, 4, 6

Index Subjectivity appropriation of quantum phenomena and, 21, 182, 187, 192, 197–199, 208, 259 literary criticism and, 24, 27, 130, 145, 148 perceived experience and, 55–56, 117, 172, 272n5 Subject/object duality appropriation and, 173, 187, 190–191 complementarity and, 7, 35, 90, 114 language and, 40, 61 literary criticism and, 16–18, 125, 131, 148, 282n65 visualizability and, 100 Superposition, 2, 8, 17–18, 22–23, 25, 115, 148–154, 205. See also Observer effect; Wavefunction collapse Surveillance society, 197 Swing, Raymond, 245, 248 Symmetries and Reflections (Wigner), 181 Szilárd, Leó, 212 Tamre, David H., 136–137 Tao of Physics,The (Capra), 169, 171, 186 Tempest,The (Shakespeare), 256–257 Temporal/spatial relations. See also Spacetime; Spacetime continuum appropriation of, 199–200 intuition and, 72, 95–97, 100 language and, 28, 35, 40–44, 47, 66–67, 106–107, 113, 116, 213, 241, 243 literary criticism and, 17, 139, 159, 161 metaphors and, 53–55, 62 visualizability and, 70–71, 75, 80–87, 91, 226 Tenor, 215–217 Thingness, 12, 33–37, 41–43, 54, 66, 85, 147, 226, 241, 255, 257, 262. See also Nominal language THINKING IS BUILDING/ FORMING/SHAPING image schema, 61–65

Index Thirty Seconds Over Tokyo (film), 231 Threshold concepts, 263 Thye, Edward, 223 Tibbets, Paul, 230–231, 293nn51–52 Tiller,William, 198 TIME AS SPACE image schema, 54 Time magazine, 227, 292n39 Times,The, 127 Tompkins, Penny, 28 To the Lighthouse (Woolf), 131–133, 163 Tracy, Spencer, 231 Trinity test, 217–218, 221–222, 235, 247 Truman, Harry, 220–222, 228, 248–249 Turner, Mark, 28 Turning Point: Science, Society, and the Rising Culture,The (McCance), 191 “Ultimate Edition of Quantum Jumping” (CD collection), 205 Ulysses (Joyce), 149–150 Uncertainty Principle appropriation of, 21, 173–174, 192–195, 258–260, 289n84 causality and, 38, 53 emergence of, 5–6, 34–35, 56–57, 104–105, 265nn2–3, 283n82 language and, 42, 45, 55, 67, 85–86, 106, 116, 226, 266n6, 272n5, 284n91 literary criticism and, 17–18, 140–151, 154–157, 166, 282n65 misuse of, 192–195 visualizability and, 13, 71, 81 Underworld (DeLillo), 127, 156 “Undulatory Theory of the Mechanics of Atoms and Molecules, An” (Schrödinger), 70 Ungar, Sheldon, 223 United Nations Atomic Energy Commission, 246 “Unreasonable Effectiveness of Mathematics in the Natural Sciences, The” (Wigner), 180 UP-DOWN schema, 29, 32–33

327 Van Kirk,Theodore, 230–231 Vehicle, 215–217 Verb-based language, 12, 41–43, 66 Virginia Woolf:The Common Ground (Beer), 127 Visualizability in quantum physics, 2, 6, 13, 67, 71–72, 121–123, 226, 229 “Visualization Lost and Regained:The Genesis of the Quantum Theory in the Period 1913–1927” (Miller), 74, 92 von Helmholtz, Hermann, 113 von Humboldt,Wilhelm, 113 von Neumann, Johann, 2 Walton, Ernest, 212 Wavefunction collapse, 8, 18, 148–154, 181–182, 205, 266n6, 283n84. See also Observer effect; Superposition Wavefunction equation, 5–6, 8, 70–72, 83–87, 195, 295n4 Wave mechanics, 2–5, 45, 56–59, 67, 70, 75, 79–95, 101, 112, 115, 265n4 Wave packets, 35, 287n51 Wave/particle duality appropriation of, 173, 188–189, 196–198, 259–260 language and, 41–43, 47, 77, 229 literary criticism and, 16–17, 127–136, 139–140, 151, 157, 166, 280n22, 281n30, 282n65 theory of, 1, 4–8, 35, 38 Waves,The (Woolf), 127, 133, 154–156, 158 “Wave Theory and the Rise of Modernism” (Beer), 130 Wavicle, 188–189 Waymack,W.W., 250 Wayne, John, 231 Weart, Spencer, 293n52 Web philosophy of quantum physics, 171–172 Weissmann, George, 180 Western logos, 125, 146, 214, 282n65

328 “What Does Anschauung Mean?” (article), 92 “What Is an Elementary Particle?” (Schrödinger), 36 What the Bleep Do We Know? (film), 198 Wheeler, John, 180–182, 186 “Who Invented the ‘Copenhagen Interpretation’? A Study in Mythology” (Howard), 272n5 Wigner, Eugene, 180–182, 186, 286n30 “Women, Fire, and Dangerous Things” (Lakoff), 37 Women’s studies, 16 Woolf, Leonard, 127–128 Woolf,Virginia, 16–17, 127–134, 141, 163–164 Wooters,William, 287n42 Wozniak, Steve, 184 Wu Li, 175–176, 180 Yank in the RAF, A (film), 231 Years,The (Woolf), 127 Yom, Sue Sun, 127, 129 Young,Thomas, 3–4 Zohar, Danah, 187–192, 287n51 Zukav, Gary, 20, 169, 171, 175–182, 197, 199, 201, 208 Zurek,Wojciech, 287n42

Index