[Grid] Biology in the Grid [Phillip Thurtle]

273 Pages • 94,683 Words • PDF • 12.5 MB
Uploaded at 2021-09-24 07:14

This document was submitted by our user and they confirm that they have the consent to share it. Assuming that you are writer or own the copyright of this document, report to us by using this DMCA report button.


Biology in the Grid

CARY WOLFE, SERIES EDITOR

46 Biology in the Grid: Graphic Design and the Envisioning of Life Phillip Thurtle



45 Neurotechnology and the End of Finitude Michael Haworth



44 Life: A Modern Invention Davide Tarizzo



43 Bioaesthetics: Making Sense of Life in Science and the Arts Carsten Strathausen



42 Creaturely Love: How Desire Makes Us More and Less Than Human Dominic Pettman



41 Matters of Care: Speculative Ethics in More Than Human Worlds Maria Puig de la Bellacasa



40 Of Sheep, Oranges, and Yeast: A Multispecies Impression Julian Yates



39 Fuel: A Speculative Dictionary Karen Pinkus



38 What Would Animals Say If We Asked the Right Questions? Vinciane Despret



37 Manifestly Haraway Donna J. Haraway



36 Neofinalism Raymond Ruyer



35 Inanimation: Theories of Inorganic Life David Wills



34 All Thoughts Are Equal: Laruelle and Nonhuman Philosophy John Ó Maoilearca



33 Necromedia Marcel O’Gorman



32 The Intellective Space: Thinking beyond Cognition Laurent Dubreuil



31 Laruelle: Against the Digital Alexander R. Galloway



30 The Universe of Things: On Speculative Realism Steven Shaviro



29 Neocybernetics and Narrative Bruce Clarke



28 Cinders Jacques Derrida



27 Hyperobjects: Philosophy and Ecology after the End of the World Timothy Morton



26 Humanesis: Sound and Technological Posthumanism David Cecchetto



25 Artist Animal Steve Baker



24 Without Offending Humans: A Critique of Animal Rights Élisabeth de Fontenay (continued on page 263)

Biology in the Grid Graphic Design and the Envisioning of Life posthumanities 42

posthumanities 43

posthumanities 44

Phillip Thurtle 45 posthumanities posthumanities 46

posthumanities 47

posthumanities 48

posthumanities 49

posthumanities 50

University of Minnesota Press

posthumanities Minneapolis51 London

posthumanities 52

posthumanities 53

A multimedia website accompanies the theoretical and historical analysis presented in Biology in the Grid: Graphic Design and the Envisioning of Life. At immaterialwings.org you will find vignettes from the history of biology, movie clips, images, and excerpts from speculative fiction. These resources can be interwoven to help you envision scientifically robust, imaginatively engaged, and visually rich stories about the role of bodies in the grid. Copyright 2018 by the Regents of the University of Minnesota All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior written permission of the publisher. Published by the University of Minnesota Press 111 Third Avenue South, Suite 290 Minneapolis, MN 55401-­2520 http://www.upress.umn.edu Printed in the United States of America on acid-­free paper The University of Minnesota is an equal-­opportunity educator and employer. 25 24 23 22 21 20 19 18

10 9 8 7 6 5 4 3 2 1

Library of Congress Cataloging-in-Publication Data Names: Thurtle, Phillip, author. Title: Biology in the grid : graphic design and the envisioning of life / Phillip Thurtle. Description: Minneapolis : University of Minnesota Press, [2018] | Series: Posthumanities ; 46 | Includes bibliographical references and index. | Identifiers: LCCN 2018008943 (print) | ISBN 978-1-5179-0276-6 (hc) | ISBN 978-1-5179-0277-3 (pb) Subjects: LCSH: Biology—History—19th century. | Biology—History—20th century. | Biology— Graphic methods—History. | Art and biology—History. Classification: LCC QH305 .T627 2018 (print) | DDC 570—dc23 LC record available at https://lccn.loc.gov/2018008943

For the warped, weird, and wandering

This page intentionally left blank

Contents Introduction: The Varieties of Gridded Experience

1

1. Life on the Line: Organic Form

25

2. Envisioning Grids

55

3. Warped Grids: Pests and the Problem of Order

91

4. Modulations: Envisioning Variations

129

5. Drawing Together: Composite Lives and Liquid Regulations

167

Epilogue: Toward the Nonsynthetic Care of the Molecular Self

211

Acknowledgments215 Notes219 Index249

This page intentionally left blank

Introduction The Varieties of Gridded Experience

When I trained as a molecular biologist in the 1980s, I spent most of my time characterizing biomolecules. I cloned genes to understand how that gene product functioned in cells; I isolated RNA to see how and when it was expressed; and I assayed under what conditions a specific protein might be expressed. One evening, while learning a new protocol for isolating proteins, I asked a postdoctoral fellow in the lab next to mine if he ever thought that the molecular biological sciences would move toward a science of “building things back up” after all this characterization. I was interested in seeing how all the bits of knowledge we gathered fit together to form a picture of how a cell, or an organism, might operate. Not in the piecemeal fashion we were currently glimpsing, a cellular function discovered in one part of the cell, a molecular signal glimpsed in another. I wanted to see how all these insights might be orchestrated together. If the protein expression problem I was working on could be compared to a single musical instrument, I wanted to hear the grandeur and complexity of the whole cellular symphony. The postdoctoral fellow replied that he didn’t think the type of science I was interested in would happen for a long time. As he saw it, there was still too much characterizing that had to be done. In his view, the molecular biological sciences of the twentieth century would remain as mostly detailed depictions of molecules, cells, and organisms. We just didn’t have enough information to develop a comprehensively synthetic (as opposed to an overtly analytic) molecular biology. Only after we had characterized the world sufficiently, would we be able to see how the parts fit together. To be a successful scientist, he suggested, I needed to dedicate myself to the intellectual world of a single amino acid on a single molecule. 1

2 · Introduction

Well, the era of “building things back up” has most certainly arrived. Many molecular biologists today embrace the use of computation, art, animation, engineering, systems thinking, and design to promote making as a new form of understanding. Disciplines as varied as synthetic biology, evolutionary and developmental biology, systems biology, bioengineering, bioinformatics, and even biodesign flourish because of a deeply held commitment that understanding how biomolecules interact to form organelles and cells is important for understanding how living things operate. If we frame this insight into terms of literary theory, one could claim that within the last few decades, biomolecular scientists increasingly adopted world building as a strategy for exploring the complex interactions that form living things.1 I’m not the first to notice this shift in biology. Just ask the scientists. Synthetic biologists have written on how “biology is technology” as cells can be used as “platforms” to fabricate economically important biomole­ cules, such as pharmaceutical drugs.2 Disciplines such as evolutionary and developmental biology temper the study of the selective pressure of environments with a renewed sense of the importance of the internal molecular and physiological constraints of developing organisms,3 and scientists have “pioneered” institutes dedicated to embracing “biological complexity” to fearlessly decipher “vast amounts of data” “to gain valuable insights and achieve breakthroughs across scientific disciplines.”4 Clearly something is afoot and people have noticed. What I find especially surprising is the speed with which this change appeared. For many commentators, the cause for the rapidity of this change is clear: the use of computers in biology.5 Specifically, these commentators point to the heavy use of computers in biological practices, with their ability to store and correlate large amounts of data, and the use of cyber­netic and biological metaphors for thinking about biological systems, giving researchers new conceptual tools to think about complex processes. Although I don’t think these histories of biology are wrong, and I will address these claims in more detail below, I think they are much too limited. An overemphasis on computation as a historical agent obscures two main points. The first is the problem of standardization. It is difficult to turn something as complex as a fruit fly’s body into the data and operable commands that a computer will recognize. There are a large number of historical studies that have labored to demonstrate the conceptual and technological innovations that needed to go into making a body calculable

Introduction · 3

in the first place.6 This history should not just be swept under the carpet. The second problem is the role of envisioning. I will offer a more concrete definition of “envisioning” below. For now, however, we can think about envisioning as a composite act that mixes imagination, visualization, and desire. To envision something in the biological sciences means having a vision for how something could occur under specific circumstances. Often envisioning requires tangible images, movies, animations, or diagrams to depict phenomena as well as the relations that birth them. Pointing to the adoption of computers as the prime historical agent diminishes the values, desires, and imaginative potentials that make science such an interesting field of study. My goal as a scholar of the cultural and conceptual basis of biology is to create stories that are historically informed, scientifically robust, and imaginatively captivating. Pointing to the increased rise of computing, while it is a superficially correct claim, too often tempts authors to condense these complex interactions into a single opaque technological box of agency. I have spent my career trying to understand these two moments in the history of biology, the standardization of human, animal, and plant bodies and the envisioning of how they work. My first book, The Emergence in Genetic Rationality, was a study of the role of standardization in defining biological relationships. At the time, it was one effort by many that were interested in the problem. This book, Biology in the Grid: Graphic Design and the Envisioning of Life, is a study of the role of envisioning in defining biological relationships. Understanding why so many biologists adopt strategies for world building all at once, however, demands we understand a bit more about twentieth-­century consumption practices and the values that they promote. This is key for understanding how computation and world building in biology evolve as conjoined visions for a world that can be envisioned (imagined and controlled) in all its complex interconnectedness.

Envisioning Consumers Early twentieth-­century industrialized economies had the capacity to produce a lot of stuff; what industries needed were mechanisms to convince consumers that they needed this stuff. Mid-­twentieth-­century business achieved these goals by changing their relationship to consumers.7 Mechanisms for relating to consumers, such as advertising and marketing, helped consumers visualize how products could change their lives, while

4 · Introduction

mechanisms for consumer feedback, such as focus groups, gave companies ideas about which products might be successfully marketed. Producing products no longer meant simply turning an assembly line on or off; it meant ensuring the object would sell once it was produced. The development of new ways for envisioning products was key for ensuring the regulation of goods. Companies began using mechanically, chemically, and electronically produced images to ensure that consumers desired what companies were making. Magazines swelled with advertisements for products, and newspapers promoted visually rich features, such as comic strips, to ensure that they sold even on slow news days. Publications and broadcasts splashed images across media, stores developed techniques to display goods in novel ways, and companies used novelties and gimmicks to ensure they maintained that special relationship between a brand and its consumer (do you remember the toy hidden in the cereal box?). These initiatives led to innovations in product design and distribution, and a world built on modern marketing ensued. Companies became especially adept at using visualization technologies to paint the consequences of a world transformed by consumption. This world was not simply focused on using an object’s utilitarian function to sell a product; it instead focused on the big picture, how these products could help a consumer attain a certain lifestyle. This was a world that you could participate in if you had the liquid income needed to purchase the advertised goods. Our ability to envision organisms in the biological sciences not only relies on some of these innovations, it remains tightly bound with the interplay of desire and spectacle in a consumerist economy.8

The Spectacle of the Grid One especially important innovation in visualization practices was the use of grids for graphic design.9 Grids allowed designers to break down complex forms into simpler elements and assemble them in novel configurations for maximum effect. Texts and images could be combined in novel ways, as publishers fought over the attention of readers. Grids also allowed distributers to circulate images across various media, allowing publishers to deliver scalable content to large audiences across diverse platforms (print, broadcast, and web). Overtly commercial enterprises such as advertising and the entertainment industry weren’t the only publications that used grids. Almost all publications began using grids as a standard-

Introduction · 5

ized form of layout, including scientific publications, where gridded layout design was especially welcomed for its ability to easily include more advertisements and illustrations. The advent of graphical user interfaces in personal computing programs ensured that even neophytes in graphic design could produce elegant yet standardized documents using grids. (I know, this neophyte used these programs to create some of the illustrations in the book.) The history is clear. During the twentieth century it became harder and harder to envision the world without using a grid. The sheer prevalence of grids revealed new ways for thinking about how organisms might build themselves. Scientists began seeing how bodies could be constructed as a series of parts or modules, much like a grid is composed of a series of panels.10 This not only offered scientists new ways of thinking about how bodies might be constructed, it also suggested new ways that bodies could be related. The key realization is that grids allowed scientists to think about how the autonomy of individual parts and the needs of the organism could work together. The key for this delicate balancing act is in how grids are organized. Each module is part of a larger array of modules, but it also has a degree of autonomy in defining its self-­organization. This allowed for the creation of designs based on assemblages of various types of modules as well as the use of single modules in multiple designs. Although it oversimplifies the complexity of how the scientists were thinking about the problem, it might be helpful to see how creating organisms is akin to assembling a structure with Legos. Legos use basic building blocks, similar enough that they can recombine with one another, but different enough that they can give one’s final creations different forms and functions when swapped with one another. Limbs can be switched (such as a fruit fly’s small stabilizers for full blown wings), heads modified (legs inserted where antenna should be), and (this is where the Legos example falters) biochemical pathways can be refashioned. Perhaps counter-­intuitively, theories of modularity provided biologists with a strategy for understanding how an incredible diversity of different forms could be created from seemingly homogenous building blocks. Similarity and difference, it seems, are much more closely related than we thought. These theories also reinvigorated old existential questions. If the materials of construction for a fly and a human are the same (proteins, nucleic acids, fats, and starches) and the process of construction is similar (with the same signals directing diverse, site specific structures), then how different from each other are we? The answer, it seemed, increasingly

6 · Introduction

appeared to be in how the actual modules were assembled, where and when they were placed to form an organism. Welcome to life in the grid, where a key source for our ability to understand new forms of biological complexity comes from one of the least likely sources, standardization wrought by desires for advertisements and entertainments. Grids (and as I will argue in the last chapter of the manuscript, layers for dynamic processes, as well) provide the scaffolding for envisioning how living things operate in the twenty-­first century. Analyzing our lives in the grid, then, requires a bifurcated research strategy. It is imperative to recognize the important political constraints and oppressive consequences of the use of grids in thinking about controlling life. As I will argue in depth in chapters 1 and 3, despite the homogenizing appearance of grids, not all lives in a grid are treated similarly. Uncovering these differences is politically important but also requires understanding how grids are ordered. This involves asking questions such as “What are the specific values that grids are intended to support?” Understanding life in the grid also demands the use of imagination. In this way we see how grids can be used to reorder lives to be less oppressive and more creative. Strangely, it is through a study of the most monotone and bureaucratic of terms, “regulation,” that we see how closely bound the impulse to control and the desire to imagine coexist through envisioning.

The Varieties of Regulated Experience The use of grids was just one aspect of a more important shift in the ways that products were regulated. Most people are used to thinking of regulation in terms of the application of rules or laws in social organization. For instance, laws limit how fast we can move by setting speed limits, what we can ingest by creating controlled substances, and how we interact with other bodies by criminalizing specific behaviors. A growing number of scholars from a remarkably diverse number of disciplines have argued that this is much too limited a way to think about how bodies are regulated in twenty-­first-­century society.11 Central to many of these claims is the work of the late Michel Foucault in thinking in terms of how lives, economies, and regulations interact to create regulatory practices supple in their application and responsive to the challenges of random events. Especially central to this new type of regulation is “the production of the collective interest through the play of desire.”12 Economic practices, legal norms, and

Introduction · 7

ideas about how bodies operate worked in tandem to create new forms of regulatory functions.13 This subtle form of regulation appeared more fluid in its application, more natural in its appearance, more insidious in its presence, and more expansive in its reach. Visual technologies of mid-­ twentieth-­century capital played into this change in how products were regulated. Regulation was no longer a matter of restricting which products were produced for consumption. As noted earlier, it also meant developing desire for the products in the first place. Regulation is an increasingly important concept in biology as well. Although biologists began figuring out how nucleic acids informed the expression of proteins, they couldn’t figure out why only certain proteins were produced at specific times and locations in organisms. This was an interesting conundrum as living things seemed to always be changing. How could the same materials that created a specific stage of an organism’s life create other stages as well? Insects, such as fruit flies, not only grow larger, they also dramatically change their forms as they grow. The untrained eye can find little similarity between the soft grub of a juvenile fly and the well-­articulated, chitin armored appearance of the adult. Even humans go through subtle chemical metamorphoses, as some lactose intolerant adults lose the ability to metabolize the sugar lactose, a metabolic skill they relied upon as an infant. Why did they lose this ability? Or, in more biologically precise terms, how could a limited nucleic acid code account for all the changes organisms exhibit during their lives? Clearly, something was happening to allow some parts of the code to be expressed at one time and not at another. This implied that genes not only needed to be turned on, they had to be turned on at the right time and in the right place. One point that Biology in the Grid argues is that understanding the complexity of genetic regulation not only involved changes in how scientists thought about genes, it required a change in how the concept of regulation was pragmatically deployed in society. Genetic regulation metamorphosed from being a simple directive (such as turning a gene “on” or “off ”) to a complex set of responses to localized cues (such as the response to chemicals in specific cellular environments) that occur in a specific order (where the expression of one chemical initiates an expression of another chemical). Genomes now appear less like a detailed set of instructions to build an organism and more like a set of potentially coordinated responses to localized changes. Take, for instance this claim, by developmental biologist Eric H. Davidson, that a genetic

8 · Introduction

regulatory network has two components. They are sensitive to cellular signals, meaning that signals have the capacity to “affect transcription of regulatory genes.” And they form networks, where “each regulatory gene has both multiple inputs . . . as well as multiple outputs . . . so each can be conceived as a node of the network.”14 This vision of genetic regulation as multicausal and conditionally responsive is much more complex than a single directive to turn a gene on or off. In this case, it is not just the code of the genes that makes a difference, it is how, when, and where the codes are translated and transcribed. This idea of genetic regulation was key for the emergence of world building in the biological sciences. Although this shift in the idea of regulation has registered greater impact in some fields of biology than in others, it has been surprisingly far reaching. Its influence on the field of evolutionary and developmental biology is especially profound, as the concept of genetic regulation has played a key role in thinking about how organisms develop. Genetics, too, has had to wrestle with how genes respond to cellular environments. The result has been the characterization of a long list of factors implicated in regulating genes. We now have specific nucleic acid sequences identified (such as initiation and termination sequences, promoters, enhancers, and microRNAs), specific proteins identified (such as transcription factors), and even modifications to the architecture of the nucleic acids (through methylation and histone interactions).15 Understanding the rich interplay between all these factors has led to productive debates in proteomics, structural biology, epigenetics, cell biology, bioengineering, and even synthetic biology. Grids not only helped structure how these interactions were thought to occur, in many cases they enabled the visualizations that allowed for thinking about interactions in the first place. Grids allowed for scientists to see how things could be put together in a way that wasn’t just propositional, it was additive and supple. This turn to world building has been fruitfully envisioned as a return to a biology of organic holism.16 The more I studied the development of this new way of thinking in biology, though, the more I found parallels with contemporary economic practices more compelling than parallels with older biological theories. Let me offer an extended example to illustrate what I mean. I hate to buy clothes. I really dislike spending an afternoon shopping. I find very little interest in trying things on, waiting in lines, and being jostled by others. Increasingly, I have relied on purchasing my clothes online. A series of innovations in online retail has made it pos-

Introduction · 9

sible for me to drastically reduce my visits to physical stores. For instance, I now have accounts at retail clothing sites that remember my body size and can even suggest small adjustments in size if a brand is known to run a little smaller or larger than other brands. Some sites even have virtual methods for trying on clothes, such as “virtual fitting rooms.” These can rely on mechanisms such as providing measurements made at home to create a virtual mannequin with the dimensions of your body or uploading an image so that this mannequin resembles you in more than your formal dimensions. In one sense the online experience of buying clothes is becoming more like the store experience, or even the much older, now elitist practice of going to a tailor.17 On the other hand, this process is clearly built from processes of standardization that made it possible to buy off-­the-­rack in the first place. Yet the process more thoroughly adjusts for how standards can be varied to build a body profile that accommodates even hard to fit parts of my body, such as my peculiarly short arms. Interestingly, innovations in how goods and information are regulated keep my buying habits of clothes and biologists’ conception of development from being a return to pre-­twentieth-­century emphasis on handcrafted organic holism or a simple extension of twentieth-­century standardization.18 When I purchase my shirt online, there is now a layer of informational practices added to a simple retail purchase. Information about my buying habits is stored, my body size and shape recorded, different products virtually compared, and new means of negotiating payment utilized. The same is true for envisioning the development of organisms. The movement of individual cells and molecules has been tracked, potential outcomes from molecular interactions are evaluated, cellular and physiological changes remembered, and complicated issues involving the physics of different scales of interaction are accounted for. Consequently, the seeming return to holism in biology is predicated upon the detailed regulatory apparatus of molecular development even as it confounds the reductionist assumptions of its practitioners. New forms of regulation have not only shaped my clothing buying habits, they have also informed how scientists think about the methods organisms use to shape themselves.

From Visualization to Envisioning Images play an especially important role in industrial economies for understanding how things fit together. Two points are especially important to

10 · Introduction

my argument, and I appeal to the work of media theorist and philosopher Vilém Flusser to help me explain them.19 The first point is phenomenological. Flusser grasped that readers engage with images differently than they do texts. He argued that reading lines of text requires readers to follow a structure “imposed upon us” as our eyes “follow the text of a line from left to right” and “jump from line to line from above to below it” and then “turn the pages from left to right.” Viewers of images on surfaces, such as a printed page, however, “may seize the totality of the picture at a glance, so to speak, and then proceed to analyze it.”20 Flusser is careful to acknowledge that this isn’t a difference in freedom, where our eyes are free to roam a page when looking at an image, but in the order of synthesis and analysis. Texts require analysis first, which can then lead to specific forms of synthetic meaning making, which Flusser terms “historical thought.” Images, however, provide a synthesized experience that invites specific types of analytics. “The one aims at getting somewhere, the other is already there, but may reveal how it got there.”21 The real difference involves the temporality of perception as each treats the viewer’s relationship to “past, present, and future” in different ways. Recognizing this phenomenological point is important for understanding how visual technologies are especially useful, but not necessarily required, for the types of open-ended regulations we discussed earlier. As most advertising executives know, images are especially useful for evoking desires. They present worlds where associations can be made and possible futures promised. As most lawyers know, images are very difficult for thinking in terms of laws.22 They infrequently provide clear prohibitions (the strong prohibition of a nonsmoking symbol is the exception) and often don’t suggest strong and clear causalities. Clearly, different forms of informing are important for suggesting different ways that things can be regulated. This is what makes images so useful for world building. They excel at providing a way to imagine how elements fit together to make a complex scene without necessarily saying how these elements specifically interact. This is one reason why many scientific models rely on visualization—­they allow a viewer to see how things might fit together. Flusser’s second important point about images is political economic. Flusser recognizes that images produced through technological means operate differently than traditional images. Consequently, they are implicated differently in meaning making. For instance, technical images are not direct representations of the world, they are mostly composed. The term Flusser uses, and I have appropriated, is “envisioned.” Image makers

Introduction · 11

(he calls them “envisioners”) draw “the concrete out of the abstract” by arranging bits of data into informative patterns. As anyone who has zoomed too closely into the pixels of a digital image can attest, electronic images are little more than collections of specific data points. What makes them meaningful is how they are arranged (or regulated). The representational veracity of these images, then, comes from the series of commands that organize pixels in specific locations in the image. This act of envisioning always takes place through the capabilities of a technical apparatus. “The technical image is an image produced by apparatuses.”23 This is what makes his point political economic, as all images are the products of the capability of the technological apparatus. In the universe of technical images, then, all images reflect the scientific statements that allowed for their production. “As apparatuses themselves are the products of applied scientific texts, in the case of technical images one is dealing with the indirect products of scientific texts.”24 When one envisions a world with technical images, one is always envisioning this world through the technical capacities of the society that created the imaging apparatus. This insight doesn’t necessarily suggest that images are now postrepre­ sentation (as other theorists have argued).25 It does suggest that the distinction between representation and imagination is no longer a very interesting epistemic criterion by which to judge scientific images: “The gesture of the envisioner is directed from a particle toward a surface that never can be achieved [because of its abstraction], whereas that of the traditional image maker is directed from the world of objects toward an actual surface.”26 Consequently, envisioners are much more interested in using data (visual data bound with the codes of the apparatus) to create informative configurations of images. “Envisioners press buttons to inform, in the strictest sense of that word, namely, to make something improbable out of possibilities. They press buttons to seduce the automatic apparatus into making something that is improbable within its program.” Traditional images inform by either reflecting (representation) or escaping (imagination) the normative visual qualities of a world. Technical images inform by creating unlikely images that reveal new relationships. Let’s analyze a specific image from evolutionary and developmental ­biology to anchor Flusser’s abstract insights with the particularities of biomolecular practice. Figure I.1 shows a drosophila embryo stained so that a viewer can visualize the expression of eve or even skipped pair rule genes in the upper part of the image.27 This regulatory protein is expressed as a series of discrete bands, in a modular fashion, arranged along the longitudinal axis

12 · Introduction

Figure I.1. A fruit fly embryo stained for the eve or even skipped pair rule gene bands (the seven prominent bands in the upper image and the seven less prominent bands in the lower image). The eve gene has seven different regulatory sequences that can be activated independently. Each of these regions is responsible for producing a discrete band. Each eve gene, then, acts as a module that is both dependent upon its placement in the whole organism as well as able to act independently. The fluorescent stipples dispersed across the bottom image help visualize individual nuclei of cells. The diffuse central band on the bottom image is the gene product, kruppel. This beautiful picture is also a testament to the sophistication of visualization techniques in biology that can combine the visual detail of confocal microscopy, the color of multiple forms of fluorescent staining, the specificity of monoclonal antibodies, and genetically engineered probes to identify single molecules. Photograph courtesy of David W. Knowles and Mark D. Biggin of the Berkeley Drosophila Transcription Network.

of the embryo. The bottom image combines the eve staining pattern with a stain that creates a stippled pattern used to visualize each of the cells in the embryo. The more diffuse single central band is a stain specific for the gene kruppel.28 Eve and kruppel are called “gap genes” in that if mutated, whole segments of the developing drosophila larva will be skipped during development. With eve, the even numbered segments of the larva are skipped during development, while kruppel deletes a massive segment from the posterior of the embryo (kruppel is German for “cripple”). During the decade

Introduction · 13

between 1990 and 2010, the pages of journals such as Nature, Science, and Cell were saturated with gorgeously detailed colored pictures tracking the expression of single molecules in (mostly) drosophila larvae. What makes these images a good example of Flusser’s concept of envisioning? First, these images are not directly represented in the same way that a traditional image is. They were produced using highly specific stains (using either monoclonal antibodies or specific nucleotide sequences) that effectively amplify the visual presence of molecules within a complex molecular soup. They could not be visualized without it. Second, the imaging apparatus, confocal microscopy, is different from conventional light microscopy. Confocal microscopy is known for its ability to provide complex images with especially high resolution. These images are created, however, by scanning the whole specimen as a series of closely focused two dimensional planes that are then reassembled into a grid of three dimensional images with stunning detail. Confocal microscopy is especially useful in the biological sciences as it gives researchers the ability to visualize planes of three dimensional objects with relatively little invasiveness. And finally, but certainly not exhaustively, the whole ability to identify and map these molecules in the first place comes from the creation of mutant flies who have lost the ability to express these genes in the first place. The ability to envision the normal banding of a developing fly is predicated upon the unlikely events of seeing mutant flies in the past. As Flusser argues, “Envisioners press buttons to inform, in the strictest sense of that word, namely, to make something improbable out of possibilities.”29 So, as you see, the technological coding that produces images of gene expression during development requires multiple levels of envisioning in a technical society. This point was impressively brought home to me when I ordered this image from the Berkeley Drosophila Transcription Network (BDTN). Specifically, I learned that no one fly is represented by this image. The BDTN created this composite image upon my demand by envisioning data from multiple images that they already had stored in their database. This image of eve expression in drosophila is what Flusser is referring to when he writes that envisioners draw “the concrete” from an “abstract” collection of particles or data points. It’s not that this image isn’t real, as it actually portrays how eve is expressed during drosophila development. It’s just that judging this image through the categories of real vs. unreal is less informative than with traditional images. Some of what Flusser argues will appear familiar to readers of media theory. Flusser’s recognition of the importance of the medium (or in

14 · Introduction

Flusser’s case, apparatus) in shaping messages can at times seem like Marshal McLuhan’s emphasis that the medium is the message. Flusser’s focus on the role of envisioning to create anomalous outcomes, however, shifts the focus of analysis away from the normative effects of a medium in meaning making.30 It’s not so much that “the medium is the message,” but that under specific circumstances envisioners can inform using media. His characterization of the abstract qualities of technical images can, at times, seem like Jean Baudrillard’s depiction of the “hyper-­real.” Flusser, though, emphasizes the production of concrete results from abstract processes, instead of allowing all meaning to implode through symbolic exchange.31 In a sense, Flusser stands Baudrillard on his head as he is interested in how the abstract elements of symbolic exchange still inform concrete processes. Also, in some circumstances, Flusser’s view of envisioning can seem a lot like Alexander Galloway’s conception of protocol, where envisioners and apparatuses are following protocols to create images.32 Yet, there is a major difference. Flusser thinks that use of technical images is industrial and not just informational. Consequently, the logic of the apparatus is also a chemical logic of association and not only an informational association through protocols. This is an important distinction as the experiences of industry might be chemically synthesized, but they are most certainly not informationally propositional in the way that protocols are. These distinctions are important for my argument as it helps me more deeply explore how bodies become industrially calculable before they became informationally computable. It also suggests why the industries of advertising and entertainment are especially useful for envisioning life in the twenty-­first century. Understanding how technical images are produced helps one to realize that bubbling below the slick sterile surface of the modern biological laboratory is an amusement show of special effects. Excess, illusion, and imagination fuel our hunger for knowledge about living organisms, shape the content of what it is we find, and drive the limits of what we think possible.

The Chapters Envisioned Flusser’s two insights about images, that images promote associative thought and that technical images reflect the codes of a technical society, allow me to argue several important claims in the book that follows. The first insight is that the associative power of images has been especially important for biology. Chapter 1, “Life on the Line: Organic Form,” explores this

Introduction · 15

insight through an analysis of the images of nineteenth-­century morpholo­ gist Ernst Haeckel by looking at how the aesthetics of his images relate to his theories of life. It begins by asking how one illustrates that something is alive. In addressing this question, the chapter then analyzes how in Haeckel’s work, a materialist idea of life as organic (meaning composed of carbon) is in tension with a formalist idea of life as organic (meaning the form of the organism regulates the assembly of the parts). The chapter traces these ideas through a formal analysis of Haeckel’s use of curved lines in his masterwork, Kunst-­Formen der Natur. It argues that Haeckel used an openly curved or wavy line to suggest vitality and a closed circular line to build architectural volumes for his forms. The chapter closes with a political analysis suggesting that an aesthetic analysis that privileges organic forms is insufficient to analyze the types of oppression that occur in twenty-­first-­century society. The second insight from Flusser developed by this book is that the codes of technical images reflect the codes and values of an industrial society. This insight structures much of the content of chapter 2, “Envisioning Grids.” This chapter documents the adoption of grids in the twentieth century as a publishing and display aesthetic: first in advertising and promotions and then in scientific print and electronic publications. The chapter suggests that grids promote two moments in design: a moment of partitioning a complex process into simpler elements and a moment of the reconstruction of these elements in a larger assemblage. This second moment is especially important when looking at the use of grids in the history of graphic design as it allows designers to build diverse displays from similar items. According to philosopher Vilém Flusser, this constructivist moment is due to the visual “magic” performed by images as they create meaning by demonstrating associations across surfaces. Flusser then contrasts this form of associative meaning making with the historical and sharply causal knowledge promoted by texts. The twentieth century especially, saw a proliferation of images, most of which were made using chemical and electronic technologies. Almost all of these images used the disciplinary logic of grids in their construction. According to Flusser, technologies now create images by assembling representations from quanta, bites, dots, or grains of silver. These quanta, when taken individually, are inherently absurd in that they have no intrinsic meaning. They only acquire meaning when they are assembled into specific patterns through the associative power of grids. What technological images represent then, is never just the content of the

16 · Introduction

image, but the political economic circumstances that produce the image and give it meaning. The chapter ends by showing how the development of printed scientific journals in the late twentieth century and the display of web-­based scientific journals adopted the regulative power of grids to bring coherence to the disparate facts, values, and sensory experiences produced by a technologized society. Science is thus seen as a constructivist practice in that it uses social codes to make meaning from inherently absurd collections of data. The idea of construction, though, has now changed from a conception of building up potential meanings from parts to a conception of using the codes of technical images to create visual constraints. Constructivism isn’t just a simple assembling of things but a path to creating meaning by limiting the inherently powerful act of association involved in all image making. The third insight from Flusser, that unlikely events are the most informative, drives much of the analysis of chapter 3, “Warped Grids: Pests and the Problem of Order.” The point behind this chapter is that grids are more than ideal conceptual constructs, they can be strangely responsive material constructs for ordering actual spaces. This chapter takes the historical phenomenology of grids developed in chapter 2 and applies it to a cultural analysis of how bodies, grids, and regulation were related in mid-­ twentieth-­century popular media. It begins with an analysis of a poster advertising 20th Century Fox’s 1958 motion picture The Fly to suggest how the seemingly unrelated spheres of the teleportation of goods and the elimination of pests are products of a desire to develop new strategies for regulating bodies. In the movie, the body parts of a housefly and a man are switched during a teleportation experiment when the teleportation device misregulates the reassembly of the two organisms. It then shows how this idea of mutation by misregulation was originally developed in biology in the work of biologist William Bateson, the father of cyberneticist Gregory Bateson. It does this by concentrating on how William Bateson’s focus on the role of variation in inheritance is an important milestone for thinking about the importance of regulation of bodies (most specifically through his studies of homeovariants). The chapter then moves to an analysis of Foucault’s later work on biopower and its relationship to the development of neoliberalism (from Security, Territory, and Population as well as The Birth of Biopolitics) that casts life and liberty as a problem of the regulation of the circulation of goods and bodies. The chapter ends with thoughts on how the regulative function of grids promotes some lives over other lives,

Introduction · 17

suggesting that the way a society circulates goods contributes not only to “the power to ‘make’ live and ‘let’ die,”33 as claimed by Foucault, but in the very terms of what life might be and how it structures itself. It makes this point, however, by demonstrating how in society and biology, grids are always material orders of spaces. Most complex spaces consist of many grids, or types of order, existing in relation with each other. So, although grids have defined the possibilities of life, grids can also interact, effectively warping each other, to create degrees of freedom for the pests they intend to control. This is an important insight as it broadens one’s political analysis away from the identification of conditions to seeing how conditions allow for specific types of futures. The fourth insight is that envisioning not only is a property of image surfaces, its use on surfaces helps define how we look at things; in this case, how organisms are regulated during development. Chapter 4, “Modulations: Envisioning Variations” begins by demonstrating how important the concept of “modules” is for the development of contemporary biology and for the discipline of evolutionary and developmental biology specifically. Modules, a single panel within a grid, are defined as having two key properties: they possess a degree of autonomy, meaning that they can act as individual agents with internal dynamics, and they are integrated with other modules, meaning that the functions of the entire grid depend upon the relationship of modules to each other. I locate William Bateson’s theories on variation as a key moment in biology for thinking in terms of modules. Through an analysis of Bateson’s illustrations and writing, the chapter suggests that Bateson’s ideas on segmental variation are an important precursor to the development of modularity in evolutionary and developmental biology. Bateson identified two types of variations: substantialist variations, based on how a part is put together; and meristic variations, based on how the parts are arranged. The chapter then turns to the work of Nobel Prize winning geneticist Edward B. Lewis to demonstrate how this concept of modularity works visually in Lewis’s illustrations, as well as conceptually, in Lewis’s arguments about how development occurs as a sequence of specific genetic events. According to Lewis, the expression of phenotypic traits, such as the wings of a fruit fly, depend on how molecular events are regulated at the time it was developed. Sean Carroll has elegantly described this conceptual variation as the “logic of making a series of initially similar modules and then making them different from one another.”34 This concept of development synthesizes a depiction of the developmental

18 · Introduction

sequence (as found in Haeckel) with the importance of variance (as found in Bateson). Most importantly, this logic of making modules and then varying them into new things allows for the development of an important insight in evolutionary developmental biology, that regulation can lead to variance despite the similarity of materials and processes. The final chapter of the book is an extension of Flusser’s theories on technical images into the role of grids and layers for envisioning, and controlling, over time. Chapter 5, “Drawing Together: Composite Lives and Liquid Regu­ lations,” begins where the last chapter left off, with the work of Ed Lewis. It does so, though, by returning to the question asked in the first chapter: how do you illustrate a living thing? Ed Lewis seems to have thought that he could best illustrate development by learning animation. Lewis used stop-­motion animation to put his ideas into motion so that he could show how mutations could be used to understand the steps of developmental sequences. This chapter reads Lewis’s animations through the lens of Thomas Lamarre’s theories of animation, where changes over time derive from the differential layering of images over each other, a technique in animation called “compositing.”35 I argue that scientists use animation to draw multiple types of evidence together in a single presentation and then order them alongside each other. This technique allowed animators to freely mix texts and diagrams with live action images, juxtapose different rates of change, and freely change scales between molecular, cellular, and organismic interactions. Biological explanations, I argue, are at their most robust when they draw on multiple sources to present compelling visions of life.

The Importance of Aesthetic Analysis Sprinkled throughout these chapters are several important insights about what it means to live composite lives within a grid. I will close this introduction by pointing to a few that I think are the most important. The first of these is the claim that how we order the world is a product of how we envision it. This insight carries with it aesthetic as well as epistemic consequences. Scholars of biology have long used the categories of “form” and “function” to explain why organisms are created the way that they are.36 An analysis of an organism’s functions often stresses how specific traits help an organism exploit an ecological niche. A simple example of this is how the development of the limb of a fish into a fin will help that fish swim. An analysis of an organism’s form, however, usually stresses how

Introduction · 19

the parts of an organism have been ordered to fit together in the whole organism. The presence of vestigial organs, such as the human tail bone, for instance, are often thought to be preserved because of similarities in development between humans and other primates even though humans have no tail. As I explore, mainly in chapters 1, 2, and 4, that while I see a resurgence in an emphasis on form in biology, we must not be too quick to assume it is a simple return to the holism of nineteenth-­century biology as some have assumed. I concur with Richard Burian when he claims, “The key to the integration of organisms is not dependent on a master plan, but on the coordination of quasi-­autonomous modules.”37 In fact, I would even go so far as to claim that the type of formal principles operative in biology today are similar to those described by philosopher Gilbert Simondon in his important L’individuation psychique et collective.38 Simondon rejects the idea of form as hylomorphic in that it presupposes what something will become. Instead, he posits the idea of “information” as process that occurs through the transductive ontogenesis of an individual.39 I believe, however (and this is very different from Simondon), that the history of twentieth-­century biology suggests that understanding how bodies are in-­ formed also demands understanding how modules interact through the relationships of grids.

Not Just Computing As I mentioned earlier, others have noticed the shift toward world building in biology but tend to emphasize the use of computation and algorithms as the historical catalyst. For instance, Alireza Iranbakhsh and Seyyed Hassan Seyyedrezaei argue that “The advent of computer and information technology thereafter has been making such conspicuously remarkable changes in every aspect of life that its importance cannot be over accentuated.”40 I understand the seductiveness of this view of history. Biologists today use computers to create databases and evaluate data, process images, keep records of samples, and model outcomes. Even seemingly mundane scientific tools, like water baths, are now computerized to lend them greater precision and programmability. Specifically, scholars have tended to identify the computer’s impact through two different mechanisms: the increasing number of uses with which computers are applied in biology41 or the use of computational and informational metaphors to

20 · Introduction

explain biological processes.42 I think there are compelling historical, sociological, and political reasons to complicate these narratives as they tend to underemphasize the amount of experimental and conceptual work that went into making bodies calculable in the first place and underestimate how firmly rooted scientific practice is in contemporary streams of capital. My first issue with this claim is historical. Biologists turned to computation relatively later than their colleagues in other disciplines, and even when they did adopt computers, it was initially only in a few applications with robust institutional and conceptual ties to other disciplines. As Robert Ledley wrote in 1959, most biologists and medical professionals using computers “are . . . relatively isolated research workers who are, with only few exceptions, people with extensive cross-­disciplinary backgrounds.”43 One of the key places that computers made their presence felt is in the role of molecular modeling, most notably with the studies of John Kendrew on myoglobin, where increasing the resolution of the X-­ray crystallographic data demanded more complex Fourier synthesis calculations.44 As Ledley points out, “Computers are being used to assist in the complicated and extensive computations that are frequently involved in obtaining information about the precise atom-­structure or the over-­all size and shape of crystallizable molecules, from x-­ray diffraction patterns.”45 As important as it is to recognize how computers were used for calculating model molecular structures, it is also important to recognize that researchers in other fields used different tools for helping them visualize complex problems. This consideration is especially important for understanding the work of geneticist Ed Lewis, whose experiments play a central role in the book that follows, as he turned to learning stop-­motion animation to help him explain his models for genetic contributions to development. Animating his model allowed Lewis to depict the complicated developmental changes of fruit flies while combining representational practices of illustration and live action photography at a time when computers could not. The second issue with this claim is sociological. Many recent sociologies of science have stressed the importance of work as an indicator of the effort that needs to go into making scientific claims.46 The focus on work as an analytic allows researchers to understand the important role played by an investigator’s tacit knowledge, the interaction of scientists with instruments and samples, and the institutional support that goes into making scientific claims. It also helps reveal how making items fit into grids requires constant policing. Much like a garden, scientific insights need to be

Introduction · 21

maintained and cultivated. Affective, clerical, and conceptual labor ensure that the world continues to fit into categories and can be “sorted out” when it needs to be.47 Focusing on science as at least partially sociological and conceptual work is especially revealing in thinking about the introduction of computers to biology as it encourages us to think about the experimental, conceptual, and institutional changes that were needed to make bodies computable in the first place. Standardized norms had to be defined, variances calculated, categories created, and causalities interpreted before the behavior of bodies could be reduced to standardized data sets. All this required work. This effort was the product of goals of biological thought that predate the introduction of computers to biology. We must also be careful when considering how concepts from cybernetics and information theory were used to think about biological processes. Again, I understand the attractiveness of this claim as biologists have assimilated many terms from information theory. Organisms are thought to carry “codes,” operate through “feedback,” and are composed of “networks” of interactions, while the hybrid field of bioinformatics is well established among university campuses and commercial ventures. This has led some scholars to think that biology has, in some form, always been an information science. This interaction between information theory and biology is what Eugene Thacker has referred to as “biomedia” or “an instance in which biological components and processes are informatically recontextualized for purposes that may be biological or non-­biological.”48 Thacker argues that thinking of bodies in terms of information theory promises to reconfigure “what a body can do” and drives promises of a postbiological future. “The ‘goal’ of biomedia is not simply the use of computer technology in the service of biology, but rather an emphasis on the ways in which an intersection between genetic and computer ‘codes’ can facilitate a qualitatively different notion of the biological body—­one that is technically articulated, and yet still fully ‘biological.’ ”49 Thacker is especially careful to point out that this isn’t the straightforward reductionist assumption that all biological processing can be digitized, but a more complex technoenthusiast enticement that the meeting of flesh and data might provide new forms of embodiment. Other thinkers, however, have not been as careful, and the result is to consider biomedia as a one-­way street, where computational processes usurped biological practices. Lily Kay’s careful history of the application of information theory concepts to problems of genetic coding is another well-­referenced source for

22 · Introduction

those wishing to make claims that computers have transformed biology. Close readers of Who Wrote the Book of Life: A History of the Genetic Code, however, realize that the story that Kay presents is not as one-­way as it is made out to be. Kay notes, for instance, that Norbert Weiner trained with the Harvard physiologist Walter Cannon (who wrote the book The Wisdom of the Body)50 and collaborator Arturo Rosenbleuth on the conductance of nervous tissue. It was Cannon’s work on homeostasis that Weiner credited as providing important insights on Weiner’s theories on feedback and oscillation.51 Also important is the recognition that to the great dismay of the physicists, it was the work of biochemists Marshall Nirenberg and Johann Matthaei that provided the first successful experiments to help discern the genetic code, and they were distinctly not part of the emerging informational theory “club.”52 It is also useful to remember that even François Jacob and Jacques Monod’s well-referenced experiments on the lac operon, often thought of as introducing the informational theoretical distinction between regulatory and structural genes, were designed to solve a very biochemical problem, enzyme induction, by monitoring a very biological process, bacterial sex.53 Even the important concept of negative control needed for the lac operon model, despite its many subsequent attributions to cybernetics, is historically easier to ascribe to the visit of “phenomenological” biochemist Arthur Pardee to the Pasteur Institute in 1957 as Pardee had been working on feedback inhibition of biochemical products.54 My goal in raising these seemingly detailed issues in the history of biology is to suggest that scientists in information theory, biochemistry, and genetics were working on similar problems, how standardized and calculable particles could regulate something as complex as the development of a body. I see this as a far reaching political economic problem and not a concept pristinely introduced by one discipline into another.

The Value of Warped and Composite Sensibilities This brings me to a final insight on the practice of envisioning life in the twenty-­first century. This is an insight that Flusser’s analysis implicitly suggests but doesn’t explicitly build upon. Many technical images are composites in two ways: they draw together modules into an ordered grid and then they layer grids upon each other to control differential rates of change. Therefore, technologies that draw elements together, like animation or illustration, are especially useful for biologists trying to understand how living

Introduction · 23

things are built. Yes, these technologies reduplicate inherently oppressive systems of labor and yes, these technologies are implicated in control. Too often, however, the constraints that provoke creative response are read through a purely restrictive politics of control.55 If unchecked, this practice can lead to a return to an antiquated politics of “wholes,” where the binary of control/not control overrides any analysis of the importance of understanding order. Not all oppressions are the same and a study of how things are ordered through envisioning practices is important to understand how deeply imagination, order, and oppression are routinely intertwined. The goal, I believe, is not to release one’s self from a grid, as all things exist in and through being included in multiple series of order. Rather, a good starting analysis would be to evaluate the “degrees of freedom” allowed to the elements differently situated within grids (as I do in chapter 3). This is why evaluating how lives are envisioned in the twenty-­first century is such an important project. Not all technical images are envisioned in the same way. Some images are created so that they obscure their composite nature, presenting the image as a seamless whole. This type of drawing together can obscure the differences that give an image its vitality and depth. As media theory recognizes, though, this is not the only way that images are constructed. Sometimes it is politically important to use techniques of open compositing, where relationships are held together without being subsumed to each other.56 And as we have learned from critical race studies, sometimes, it is politically important to draw conflicting orders together so that they warp each other, allowing for more degrees of freedom for all in the grid. The important invariant to all these approaches, however, is the imperative to keep envisioning (meaning using both the application of critique and imaginative experimentation) in order to understand the aesthetic and political consequences of biology in the grid.

This page intentionally left blank

1

Life on the Line Organic Form

Although philosophers and scientists have always had difficulty describing life, artists have illustrated it with abandon.1 From pictures in textbooks, sketches in natural history notebooks, or even a “life drawing” session at the community art center, drawing life just seems easier than explaining it. When we draw a picture of an organism, however, how do we show that it is a living thing? Are there special techniques that we use to bring an image on a page to life? If so, what are they and how do they operate? Do we condense an organism’s motions into a physical form that appears to unfold in time, like a Rodin sculpture (Figure 1.1)? Or do we emphasize movement across a diagonal picture plane, like in a William Blake etching (Figure 1.2)? Perhaps we choose moments where the curves and angles of the body are at their most extended, as Stan Lee and John Buscema proselytized in their book How to Draw Comics the Marvel Way (Figure 1.3).2 What qualities of living things do we need to emphasize to suggest that something is alive? And what do these drawings tell us about how we think bodies are organized or how we think they operate? Perhaps it is the relative ease with which we draw life that has made biology one of the most visual of sciences. In fact, it is hard to imagine twenty-­first-­century biological practice without the use of images. They are used in collecting data and specimens, working out one’s ideas, communicating results to other scientists, training young biologists, and disseminating scientific knowledge to audiences eager for new ways to see the world they inhabit. Because of this, the decisions one makes about how to illustrate living things hold specific consequences for how we think about life. These consequences could range from how one groups an organism with other organisms, how one demonstrates how organisms or species change over time, or how one organizes an organism according to how it is formed. It is not just that a picture conveniently condenses a textual 25

Figure 1.1. Figures 1.1–­1.3 show three different conceptions of living figures. In Auguste Rodin’s sculpture, Adam, notice how Rodin places Adam in an especially awkward pose that emphasizes the contraction of his body. This effect is sharpened through Adam’s pointing to the ground and the incongruous placement of his legs in relationship to his upper body. Rodin is most famously quoted discussing the relationship of sculpture to instantaneous photography with student Paul Gsell. In sculpture, Rodin thought, a gesture should be seen to unfurl. See Auguste Rodin, Rodin on Art and Artists (New York: Dover Publications, 1983), 33–­34. Image is courtesy of New York Metropolitan Museum of Art and is in the public domain.

Figure 1.2. The title page of William Blake’s, The Marriage of Heaven and Hell, copy D. An ascending diagonal movement suggests the energetics of life in many of Blake’s relief etchings. See, for instance, how souls emanating from the marriage of heaven and hell move across the diagonal, from left to right, from the fires of hell to the growing flora of the earth. Image courtesy of Library of Congress and is in the public domain, https://www.loc.gov/ item/scd-rbc.2004rosen1799D/.

28 · Life on the Line

Figure 1.3 Selection from Stan Lee and John Buscema, How to Draw the Marvel Way, reprint edition (New York: Touchstone, 1984). In this illustration from chapter 6, “The Name of the Game is –­Action!” Lee and Buscema argue that drawing the body at its most extended possible point lends the figure energy and a sense of action. Copyright 1978 by Stan Lee and John Buscema. Reprinted with the permission of Fireside, a division of Simon & Schuster, Inc. All rights reserved.

description (that a picture is worth a thousand words), but also that we use the tools of imaging, such as composition, form, color, and line, to picture what words have trouble articulating. So how have these tools helped us explore the ways that living things are related to one another, and how have we used them to comprehend the ways that organisms develop and evolve? Fortunately, the last forty years have seen a proliferation of analyses on the importance of images in science in general and biology in particular. Each of these publications enlarged a palette of approaches for thinking about how scientists and artists use images to depict living things. We now have excellent studies published by art historians,3 philosophers and theoretical sociologists,4 historians,5 and journalists and scholars of visual culture.6 We have studies of maps, atlases, models, guidebooks, films, bioart, and diagrams. We have seen the role of science in art, art in science,

Life on the Line · 29

aesthetics, and the power of images. You get the picture. Scholars today are armed with many approaches for thinking about the role of images in scientific thought. Informed by the goals of these projects, if not always the specific approaches, this chapter will focus on the role of illustration in one of the key concepts in biology: the idea of the organic. When asked about what comprises a living thing, scientists may resort to the retort that life is something that is “organic.” But this easy answer hides a deep philosophical ambiguity. On the one hand, the term “organic” refers to a property of form, as in the artistic sense of how an image is organized or composed. On the other hand, the term “organic” also refers to a material quality, as in the chemical sense of objects being composed of carbon. Although these two notions sit in tension for much of the history of biology,7 as we will see, they especially inflect the art and science of the nineteenth-­century morphologist Ernst Haeckel. As a scientist who contemplated a career in art and concentrated on forms in nature, Haeckel used his theories and images to explore this tension between material and formal properties in late nineteenth-­century biology. Haeckel is a popular figure in recent histories of biological science, and I intend to draw on these studies in my discussion below. My use of his work, however, is very different from that of most historians. I am not interested in evaluating his contributions to biology, adjudicating on his scientific ethics, or even elaborating on the relationship of his personal history to the time period in which he lived. For me, Haeckel is an exemplar who, through his personal interests and his popularity, reveals interesting nineteenth-­century assumptions for thinking about how forms are used to represent living things. In doing so, he proves to be an exceptional transitional figure in considering issues of form in biology. On one hand, Haeckel continues a tradition for thinking about biology as a science of form as related to aesthetic judgment. On the other hand, Haeckel’s efforts to compare developmental sequences across species allow him to develop certain forms of representational practices used in twentieth-­century developmental biology, such as the use of grids that we will study in chapter  2. Today these practices seem a bit at odds with each other as one evokes organic unity while the other industrial production. For Haeckel, they provided two different tools for contemplating how life productively channeled the tension between structure and change. This tension between structure and change, however, is as prevalent in

30 · Life on the Line

the history of ideas on aesthetics as it is in the history of ideas on biology. One reason for this similarity is that they share a homology of philosophical descent as inherited through the tradition of German idealism. The third critique of Immanuel Kant, The Critique of Judgment, for instance treats the appreciation of beauty and the appreciation of life as two related types of judgment. As is well known, this text was very influential for Johann Wolfgang von Goethe as he attempted to develop a science, a “delicate empiricism” or morphology, of the study of the dynamic nature of plant and animal forms.8 Although Goethe’s morphology was important for the development of biology it has also been an important touchstone for many thinkers wishing to rethink the role of forms in art. For example, writing in the 1930s, art historian Henri Focillon appealed to morphology in his attempt to understand some of the figural and formal complexities of art beyond a strictly narrow iconographic approach. As Focillon stressed in his influential text, The Life of Forms in Art, forms tended to follow their own logic, their own sense of rules for development, much like an organism. The most important method for studying forms in art then was not representational but developmental, not to its references to something outside itself but to the internal logic of how these forms changed when they related to one another. “Plastic forms are subjected to the principle of metamorphoses, by which they are perpetually renewed, as well as to the principle of styles, by which their relationship is, although by no means with any regularity or recurrence, first tested and then made fast and finally disrupted.”9As Jean Molino writes in his commentary on Focillon, “Nothing explains the genesis of forms, nothing, that is except forms themselves and their encounters with other forms.”10 There is a logic to how images are composed that can’t be reduced to representation. Even somebody as canonic as Wilhelm Worringer recognized the relationship between biology and aesthetics when he argued for the important role of abstraction in the development of art.11 Worringer appealed to the distinction between organic and inorganic matter to make his case: “Just as the urge to empathy as a pre-­assumption of aesthetic experience finds its gratifications in the beauty of the organic, so the urge to abstraction finds its beauty in the life-­denying inorganic, in the crystalline, or, in general terms, in all abstract law and necessity.”12 For Worringer, it was a sense of abstraction born from ornament, as opposed to the empathic copying of organic forms that provided the necessary impulse for the development of art.

Life on the Line · 31

I’ve come to view this tangled web between aesthetics and biology as a persistent and instructive mutual informing, as opposed to an anachronistic echo of a premodern conflation between art and science. When scholars focus on how science is a preeminent method for explaining the world (which it is), we’ve tended to cleave science from other forms of knowing. One of the main lessons of the entangled nature of biology and art then is that branches of both fields have been concerned with how forms change. As Molino insists, “[Focillon’s thought] does not make form a living organism so much as life itself a form.”13 When considering the relationship of life to forms, I think it is important to appeal to scientists, philosophers, artists, and historians of art. They tell us that biology and art have often started with similar assumptions about what it means to be alive, what it means to preserve coherence through change, and why certain shapes might be privileged over others as change occurs. Understanding how forms relate in biology and science is important for understanding what it meant to be alive at different times in history. Over the next few chapters, I will outline what I think is one of the most important changes in how organisms were conceived in the history of biology. During the eighteenth and nineteenth centuries, organisms were thought to be collections of parts that fit together to form a whole. In the twentieth century, these formal strategies gave way to an abstract logic of regulation where questions about the relationship of the parts to the whole were suspended in favor of questions about how parts related to each other. Forms were emptied of their volumes, and parts were reduced to steps in abstract sequences. What emerged as important was not the teleological relationship of the parts to the whole, but how regulation functions to ensure that parts are assembled in the correct sequence. This did not mean that biology lost its emphasis on forms or aesthetics; it does mean that the idea of form was fundamentally changed. As we will see, this change is important as each of these ways of conceiving of organisms brings with it a different form of politics. What this chapter and the next will argue is that Haeckel stood as a transitional figure in how life was conceived through images. Haeckel’s formal strategies, indeed much of his philosophy of development, depended on a holism of forms. Yet there are times when his emphasis on the role of materials and in his adoption of grids for comparing developmental sequences challenged this holism in troubling ways. For instance, Haeckel’s ability to demonstrate development through a series of images

32 · Life on the Line

marks his legacy for contemporary evolutionary and developmental biology. As Michael Richardson and Gerhard Keuck argued in an article entitled, “Haeckel’s ABC of Evolution and Development,” “Despite his obvious flaws, Haeckel can be seen as the father of a sequence-­based phylogenetic embryology.”14 This tension between form and sequence in Haeckel’s arguments and illustrations is especially intriguing for me. To understand this tension, I will investigate how it wends its way through the pages of several of his publications, but most especially his masterpiece of visual forms, Kunst-­Formen der Natur (Art Forms of Nature) and his popular science texts such as Natürliche Schöpfungsgeschichte (The History of Creation). As most biologists realize, the way that one displays a specimen also conveys a view of how the world is ordered. Cabinets of curiosities, dio­ ramas, encyclopedias and bestiaries, field guides, and atlases order conceptions of the world as they present the objects in the display. This requires biologists to make decisions about the qualities of the world that they want to stress and the qualities of the world that they want to ignore as they design the display. Recently, historians of biology have enthusiastically turned to studying the rich visual culture of biology. At the heart of these studies is the realization that the aesthetic and biological choices are informing each other in presenting conceptions of how the world is ordered. Questions such as “Should we group objects based on the location of where they were found? Or do we group them based on physical resemblances?” possess biological as well as aesthetic implications for how viewers understand and appreciate the relationship among organisms. A very good example of the role of the presentation of forms in the history of biology comes from studies in American paleontology and how dio­ramas of the evolution of the horse changed during the twentieth century at the American Museum of Natural History. The earlier diorama showed a single succession of horses, from the small mesohippus to the much larger modern equus. The exhibit now avoids any suggestion of a single progression by showing multiple forms in relationship to each other. In Agenda for Antiquity, Ronald Rainger argued that the early diorama relied on Henry Fairfield Osborn’s gradual and progressive ideas on evolution. This is just one important reminder that how one arranges the elements of the composition not only brings pleasure to the eye; it suggests how order might occur to the mind.15 It is especially hard to overstate the importance of lines as compositional devices in the history of biological and medical illustrations. The

Life on the Line · 33

line allows a surface to divide itself into greater complexity, creating two aligned spaces instead of the original single space. The addition of more lines to this surface can then build a sense of volume to suggest depth. Or the addition of a line to paper can bring an energy or rhythm to an illustration by giving it a sense of direction and movement. If creating an image is important for studying living things, then studying how scientists and artists use lines can tell us how they visualize what it means to be alive. In the rest of the chapter that follows we will look at how Haeckel uses curved lines in his illustrations to suggest important aesthetic and biological qualities of life in general.

The Artforms of Nature Haeckel’s Kunst-­Formen der Natur contains ten sets of ten illustrated plates published with his lithographer, Adolph Glitsch, between 1899 and 1904 (see Figure 1.4). The volume presents a visual argument for the importance of form in the study of nature. As such, the grouping of the lithographs, or the arrangement of the images, is part of the argument, where its “100 plates . . . serve as illustrations of morphological relations.”16 For Haeckel, the unifying principle behind these relationships was the forms of the organisms and the presentation of their arrangement. As a working scientist and an accomplished artist, Haeckel knew the constitutive relationship between visual display and the understanding of living things. An appreciation of the world required a reaching out, a thirst for knowledge, as well as a reflective admiration of its beauty: “The infinite variety of forms which we observe in the realm of organic life not only delight our senses with their beauty and diversity, but also excite our curiosity.”17 Morphology, or the scientific study of form, promised to bridge art and science. The visual study of life “provides inexhaustible material for the plastic arts, the scientific study of their relations, their structures, their origin and evolution, forms a special branch of study, the study of forms or morphology.”18 Beauty and knowledge are dual approaches to discerning relationships, one for receiving pleasure, the other for reflecting upon the world. Much like others in the German morphological tradition, such as Lorenz Oken and Carl Gustav Carus, Haeckel was deeply influenced by Johann Wolfgang von Goethe’s emphasis on form in the study of organisms: “Goethe here showed that in order to comprehend the whole of the phenomena, we must in the first place compare them, and, secondly,

Figure 1.4. The cover for Ernst Haeckel’s Kunst-­Formen der Natur, originally published in sets of ten lithographs between 1899 and 1904. Image courtesy of Wikimedia Commons.

Life on the Line · 35

search for a simple type, a simple fundamental form, of which all other forms are only infinite variations.”19 Goethe sought to identify how the diversity of living things might be connected together through relationships of form. Ultimately, he identified two fundamental forms, the leaf for plants and the skull for animals. It would miscast the subtleness of Goethe’s reasoning, however, to think of these forms as fixed structures, like cement building blocks.20 For Goethe, forms were always passing into and out of each other, hence the title of the treatise, Versuch die Metamorphose der Pflanzen zu erklären, or The Metamorphosis of Plants. For Goethe, form was always manifesting itself in new ways. It was this potential for change that brought forms into harmony with one another under an archetype: Here we would obviously need a general term to describe this organ that metamorphosed into such a variety of forms, a term descriptive of the standard against which to compare the various manifestations of its form. For the present, however, we must be satisfied with learning to relate these manifestations both forward and backward. Thus we can say that a stamen is a contracted petal or, with equal justification, that a petal is a stamen in a state of expansion; that a sepal is a contracted stem leaf with a certain degree of refinement, or that a stem leaf is a sepal expanded by an influx of cruder juices.21 No singular form was adequate. Life comprised a series of relationships, as forms morphed into one another. The illustrator’s use of lines tended to suggest how sequences of forms related to one another. Haeckel especially loved curved lines and he used them in at least three ways in his illustrations. We see all three of these uses of curved lines in the illustration for the cover of Kunst-­Formen der Natur. First is a series of open curves that tend to snake around the surface of the page, adding vitality and movement. On the far right of the illustration, we have Haeckel’s treasured Discomedusa in its fully tentacled glory. Its body is set at a diagonal, reinforcing the thrust of the layout of the title of the book but with the tentacles forming ringlets22 as they snake promiscuously around the side and the bottom of the illustration. The overall effect is to suggest the movement and vitality of the organism. Haeckel also uses lines as compositional elements used to connect two disparate elements with each

36 · Life on the Line

other. In the bottom right of the illustration, for instance, the tentacles of the Discomedusa reach toward the coral on the lower left of the page. The effect is to draw together the two forms and lend a sense of unity to the page. They also give the Discomedusa a sense of movement and direction, as the tentacles suggest that the jellyfish is moving away from the coral, despite the unity of the composition. When an open line folds back onto itself, however, the form builds volume and creates an enclosure. This is Haeckel’s third use of lines and can be seen in the drawing of the coral on the lower left of the page. At times these enclosures remain as soft folds; other times, the folds stiffen into corners and they make rich geometric forms. Although these enclosures tend to sacrifice vitality because they lack movement, they gain an ability to build complexity of design by adding architectural volume. The small spheroid coral at the bottom left, for instance, folds back onto itself suggesting soft membranes beginning to calcify. The invaginations lend the very compact and self-­contained vesicle a sense of detailed complexity. At their most developed these folds present a sense of a unique world, or perhaps even a cosmos, within a single organism. In the drawing of the Siphonophorae, on the upper left side of the illustration, Haeckel balances vitality with complexity by building small forms from closed lines and nestling these forms in tendrils of open lines, like a flower or a seedpod. The overall effect is that the organism appears more vegetative than its mineralized companion, the coral. Haeckel doesn’t just use lines to depict organisms in this illustration; he uses them to make an argument about the role of form for understanding living things. Consequently, this analysis of how Haeckel uses forms to suggest different dynamics of living organisms needs to be augmented with an analysis of how Haeckel composes the whole image. For it was through the overall composition of the page that Haeckel demonstrated how a use of form could provide unity for life despite its diversity. He mostly accomplishes this through two strategies. The first is to compose the illustration so that the disparate forms create a single circle of life around the title of the book. The title provides an ascending diagonal that complements how the tentacles of the medusa reach to the coral in the opposite direction. The eye then is drawn from the coral to the Siphonophorae and then across the siphons at the top to complete the circle. The second mechanism is a bit harder to grasp as it involves understanding how forms echo across developmental and evolutionary time scales. The intent of this strategy was to suggest how an appreciation of

Life on the Line · 37

forms reinforced Haeckel’s belief in the biogenetic law, where ontogeny, the development of a single organism, was thought to recapitulate phylogeny, the development of the species. This point is perhaps best understood by comparing the cover of Kunst-­Formen der Natur to a 2008 National Science Foundation artist’s rendition of the life cycle of the jellyfish shown in Figure 1.5. The forms that the jellyfish passes through during development are highly suggestive of the forms presented on the cover of Kunst-­Formen der Natur. The coral on the bottom left of Figure 1.4 resembles the enfolded larval form of the stage 2 jellyfish life cycle presented in Figure 1.5, while the Siphonophorae on the upper left of Figure 1.4, resembles the polyps of stage 4 in Figure 1.5. These two resemblances are combined in the unity of the adult form of the stage 6 medusa seen in Figure 1.5 and on the lower right in Figure 1.4. Now the overall effect of the cover illustration is to demonstrate how an understanding of forms can help one see how Haeckel’s view of evolution provides a sense of the unity of life despite its complexity. Haeckel’s most forthright discussion of the unity of existence behind all things comes in his elaboration of his belief in the philosophy of monism. Haeckel upheld monism as an alternative to vitalism, and what he characterized as the “dualism” of vitalist thought. Vitalism asserted that the universe was composed of two different substances: an inorganic substance that made up the world of things and a vital substance that made up the world of living beings. In Haeckel’s view, all matter held both “mental power and corporeal substance,” which he deemed inseparable.23 It is

Figure 1.5. An illustration of the “Reproductive Cycle of the Jellyfish” produced by the National Science Foundation. The image charts the life of a jellyfish from gamete, through larva, polyp, and mature organism. Image credit: Zina Deretsky, National Science Foundation.

38 · Life on the Line

important to emphasize that when Haeckel rejected vitalism he wasn’t rejecting metaphysics outright, rather, he was rejecting a dualistic metaphysics in favor of a monistic metaphysics. Haeckel thought vitalism was flawed in that it split matter into two different types, vital and mechanistic matter. Haeckel thought that all matter incorporated both properties. At the end of his career, Haeckel even appealed to liquid crystals as a substance that best exemplified the monist construction of the universe.24 Although Haeckel refused the doctrine of vitalism, he did enjoy describing how the special chemical properties of carbon helped to create the curves and swirls of living things, properties that he referred to as his “Carbon Theory.” These sections are especially interesting as they provide a link between substance and form in Haeckel’s monism. For Haeckel, carbon was special because of the way that it turned molecular forces such as attraction and repulsion into concrete forms that could develop, change, and reproduce. The foundation of his theory was that carbon could accomplish this because it could bind to many different elements. Most importantly for Haeckel, carbon liked to bind with itself as this allowed carbon to concatenate into long chains of molecules, the types of molecules that form living processes and that shaped the lines in his illustrations. He noted that carbon also liked to bind with other elements, such as hydrogen, oxygen, nitrogen, phosphorous, and sulfur, enabling carbon to create highly specialized molecules, such as fats, carbohydrates, proteins, and nucleic acids: [Carbon] is that element which, by its peculiar tendency to form complicated combinations with the other elements, produces the greatest variety of chemical compounds, and among them the forms and living substance of animal and vegetable bodies. Carbon is especially distinguished by the fact that it can unite with the other elements in infinitely manifold relations of number and weight. By the combination of carbon with three other elements, with oxygen, hydrogen, and nitrogen (to which generally sulphur, and frequently, also, phosphorus is added), there arise those exceedingly important compounds which we have become acquainted with as the first and most indispensable substratum of all vital phenomena, the albuminous combinations, or albuminous bodies (protean matter). Of these, again, the most important are the plasson-­body or plasma combinations (karyoplasm and protoplasm).25

Life on the Line · 39

What especially pleases me about Haeckel’s generalization is his claim that carbon’s binding properties not only allowed for life, they marked life with some of its most important phenomenal qualities: its complexity, ephemerality, and its capacity to change. Since carbon could easily bind to things, it created quasi-­stable molecules that could easily dissipate over time. As Haeckel wrote: The peculiar chemico-­physical properties, and especially the semi-­ fluid state of aggregation, and the easy decomposability of the exceedingly composite albuminous combinations of carbon, are the mechanical causes of those peculiar phenomena of motion which distinguish organisms from anorgana, and which in a narrow sense are usually called “life.”26 Haeckel thought that carbon marked how life felt, its slickness and gooeyness, as well as its temporal durations, its ability to balance structure and change. These in turn gave rise to many surprising formal properties of the world, such as the coherence found in a globule of oil, the gelatinous resili­ ence of animal flesh, the sturdy granularity of starch, or the long feathery lines of nucleic acids in precipitate. In Haeckel’s biology, matter supplied specific potentials that informed how an organism existed. These potentials were then expressed as specific formal qualities important in how an organism changed over time. We still recognize the associative properties of carbon, but we have displaced the immediacy of the phenomenological description of its effects with more abstract stories about electron sharing. Few of us now think about how the delicate swirls of a drop of egg white dispersing in water are due to carbon’s ability to bind with itself and then with the hydrogen and oxygen in the water. Few of us can make the leap from the stories we are taught about the looseness of electrons on carbon’s outer shell to the forms and rhythms of our mortal coil. This is a shame, as many of the processes of life and death come from carbon’s promiscuity. For when carbon shares its electrons, it creates bonds. Under some conditions, these bonds can lead to many other bonds and large associations emerge. Under other conditions, the same ability to make bonds can lead to the dissolution of associations, as bonds are broken in a desire to make new and different bonds. Carbon’s singular capacity for philos and its ability to use attraction and repulsion to build complex but fleeting structures mark the flesh of all living

40 · Life on the Line

things. These stories about carbon shape the formal qualities of the lines in Haeckel’s illustrations as a depiction of the forces of life. On the one hand, it reaches out to other things, the open curling lines; on the other hand, it reaches back on itself, the closed curling lines. In Haeckel’s cosmology, these centripetal and centrifugal tendencies expressed themselves in two major forces that shaped living things: the force toward diversification, which supported a division of labor leading to more complex associations, and a force toward perfection, which resulted in the coordination of the diversified elements. Another law of progress, which is quite independent of differentiation, nay, even appears to a certain extent opposed to it, is the law of centralization. In general the whole organism is the more perfect the more it is organized as a unit, the more the parts are subordinate to the whole, and the more the functions and their organs are centralized. Thus, for example, the system of blood-­vessels is most perfect where a centralized heart exists.27 Thus, the whole process of evolution and development is a process of increasing perfection toward an ur-­type, or archetypal form. Haeckel recognized the importance of variation in evolution, yet that variation was only biologically productive if it was directed by the centralized needs of the organism. It is here where the circle of life closes in on itself for Haeckel, and where life’s promiscuity is tempered by a sense of direction. It is also the aspect of Haeckel’s philosophy that seems the most indebted to a legacy of Kantian thought. For it is the needs of the whole organism that eventually subordinates the functions of the individual parts. The way that Haeckel describes the interactions of the forces toward diversification and perfection, based on the dialectical relationship between the whole and the parts, distills the relationship between morphology and Kant’s teleological judgment into a singular statement of emerging form. This is a topic that others have written about in depth and with intelligence. For my purposes, I just want to indicate how Kant’s Critique of Judgment can inform a broad conception of the role of aesthetics in biological knowledge as well as see how it suggests strategies for thinking about how organisms are shaped. Some historians have even indicated the third Critique as a source for tendencies within German culture to appeal to holism in general.28

Life on the Line · 41

Most philosophers read Kant’s third Critique as Kant described it in its first introduction, as using the concept of “judgment” to provide a conceptual bridge between the idea of causal determinism presented in the first Critique and the development of moral reason provided in the second Critique.29 In doing so, Kant worked to complete a philosophical system that described how understanding is a product of general laws that structure human experience, where scientific understanding, the development of morality, an appreciation of the beautiful, and the workings of nature are mutually consistent. By the time he wrote the first Critique, or the Critique of Pure Reason, Kant had come to believe that all human experience was a product of the concepts, or forms, that structure experience. Consequently, knowledge was not just a product of an experience of the world but included the conditions for reason and how our understanding constructs experience. In his second Critique, or the Critique of Practical Reason, Kant turns to understanding how the same types of a priori understandings inform a moral philosophy, or how the world ought to be. This allows for a sense of ends that can guide conduct. In his third Critique, Kant attempts to bridge these two positions by investigating what he means by judgment, the faculty “for thinking the particular as contained under the universal.”30 The book is divided into two main parts: an investigation into the role of aesthetic judgment, and an investigation into the role of explanations in nature, or teleological judgment. Key to the importance of the third Critique is the role of “reflective judgments” which Kant sets against “determinative judgments.” Determinative judgments are important as they allow us to ascertain whether something falls under a general rule.31 A reflective judgment, on the other hand, is a judgment based on feeling and allows one to see the possibility of a general rule by reflecting on something specific: The power of judgment can be regarded either as a mere faculty for reflecting on a given representation, in accordance with a certain principle, for the sake of a concept that is thereby made possible, or as a faculty for determining an underlying concept through a given empirical representation. In the first case it is the reflecting, in the second case the determining power of judgment.32 Kant then proceeds to develop the distinction between teleological judgments and judgments of taste as different types of reflective judgments.

42 · Life on the Line

In teleological judgments, as discussed in more depth below, the ultimate purpose is considered. In judgments of taste, however, one can reflect without seeing a specific purpose or an end. Judgments of taste are interesting for Kant in that they identify universal principles from observations. They do this through their purposiveness, which puts into harmony the faculties of free imagination and understanding. Purposiveness, however, needs to be seen in distinction to purposeful. Purposiveness allows one to see how things may be possible without giving them a specific purpose or end, allowing for an understanding of the properties in which a purpose could be identified without seeing the purpose “as the cause of the object.”33 In other words, one can reflect on the composition of an organism or the beauty of an object without assigning it an ultimate purpose. Purposiveness can thus exist without an end, insofar as we do not place the causes of this form in a will, but can still make the explanation of its possibility conceivable to ourselves only by deriving it from a will. Now we do not always necessarily need to have insight through reason (concerning its possibility) into what we observe. Thus we can at least observe a purposiveness concerning form, even without basing it in an end (as the matter of the nexus finalis), and notice it in objects, although in no other way than by reflection.34 Thus, we sense that an object may have a goal of its own, although we may not perceive what that is, something that Kant calls “purposiveness without an end.”35 Reflection is a means to recognize purposiveness as it coordinates the cognitive faculties with the purposiveness of the object’s form. This allows for disinterested aesthetic contemplation where, according to Robert Wicks, “the intelligibility of the object’s form resonates with our cognitive faculties and generates a feeling of approval—­a feeling that is independent of sensory gratification and which concerns merely the quality of the object’s configuration.”36 The reflective judgment of beauty, for instance, helps in ascertaining the purposive organization of an object without reducing it to a singular functional outcome. “Beauty is the form of the purposiveness of an object, insofar as it is perceived in it without representation of an end.”37 Beauty for Kant is then implicated in formal principles of organization as they generate reflective judgments that engage the interplay of understanding and imagination.38 Teleological judgment, the judgment used for understanding living be-

Life on the Line · 43

ings, is different from aesthetic judgment in that an end is suggested. Thus, reflection is no longer just a reflection on form without a determining concept; teleological judgment consists of the concept of the object as well. This makes the object both the cause and effect of itself. For Kant then, teleological judgment sets up a special relationship between the parts of an organism and the regulative capacity that brings these parts together into a functioning whole. Just as the mechanism of nature, according to the preceding section, is not by itself sufficient for conceiving of the possibility of an organized being, but must (at least given the constitution of our cognitive faculty) be subordinated to an intentionally acting cause, the mere teleological ground of such a being is equally inadequate for considering and judging it as a product of nature unless the mechanism of the latter is associated with the former, as if it were the tool of an intentionally acting cause to whose ends nature is subordinated, even in its mechanical laws.39 Kant, then, remains important for our understanding of Haeckel as he provides a sense for how forms could lend themselves to aesthetic appreciation as well as an ability to discern the purposiveness of living things. The important point is that the relationship of mechanism to organization relied on the object as the organizing principle. Haeckel, however, was critical of Kant’s conception of teleological judgment as he thought it reintroduced an Aristotelian reliance on final causes.40 The consequence is that it reintroduced a duality in Kant between judgment and reasoning and thus between physics and biology, and mechanism and vitalism. In the whole domain of Biology, on the other hand—­in Botany, Zoology, and Anthropology—­mechanism is not considered [by Kant to be] sufficient to explain to us all their phenomena; but we are supposed to be able to comprehend them [according to Kant] only by an assumption of a final cause acting for a definite purpose (causa finalis).41 Although it is this pointed criticism that would lead to controversy with the neo-­Kantians of his day,42 it is important to note that Haeckel didn’t necessarily disagree with Kant’s overall analytical scheme, his work on

44 · Life on the Line

purposiveness and judgment. In some sense, it wasn’t that Kant was wrong, he just didn’t anticipate how important theories of evolution were for elevating judgments to the status of reason. Haeckel believed that Darwin fulfilled Kant’s wish for the “Newton of a blade of grass” by providing a theory of descent that offered a unifying explanation of all biological processes through the mechanical principles of reasoning: “Now, however, this impossible Newton has really appeared seventy years later in Darwin, whose Theory of Selection has actually solved the problem, the solution of which Kant had considered absolutely inconceivable!”43 Historians and philosophers have discussed at length exactly how important teleological thought was for biology.44 Philosophers have also weighed in on whether natural selection provides an epistemological foundation like that found in the mechanist reasoning of the physical sciences. Even with the controversy, Haeckel’s work remains striking for his thoughts on how material properties inflect formal manifestation of these properties. Or, as Robert Brain has recently observed, how Haeckel’s “protoplasmania” could be built into the forms needed for a functioning organism.45 This form was built from gestures, the curved line, but it also had a shape, if not exactly a teleology, to close in on itself to build wholes from parts. It was at this point that lines turned into forms with volume and complexity of structure. This is an important observation on the role of form in biological processes. As we saw in some of Haeckel’s popular lectures, forms were thought to possess a specific relationship of the parts to the whole. Forms and totalities were thought to be in harmony, and this harmony sublimated the vitality of life to increasing the perfection of the organism. Thus, the Kunst-­Formen der Natur is as much a disquisition on the formal properties of life as it is a treatise on the relationships of organisms. The formal properties of the illustrations, especially the relationship of shape to line, are the arguments that help define the relationships. The positioning of the elements within each illustration is part of Haeckel’s larger project, inherited from Goethe, of demonstrating how form is the central unifying tendency in all living things. This view allows one to see the forms that resonate through the whole collection of prints. In the plates of the amoeboid protists, the Phaeodarea, presented in Figure 1.6a, Haeckel seems to revel in life’s capacity to organize into elaborate forms as he labors over the architectural complexity of the small organisms. In other prints, such as that of the faces of the Chiroptera, shown in Figure 1.6b, Haeckel takes great joy in tracing the

Life on the Line · 45

lines, vibrations, ruffles, and curves that suggest a vitality, perhaps even a baroque strangeness, in life. Yet, even as he clearly loves the frills found on the faces of bats, it is important to note which of these two tendencies still gets the upper hand, and the Chiroptera are a wonderful example of this. As the art historian Henri Focillon reminds us, “Whereas an image implies the representation of an object, and a sign signifies an object, a form signifies only itself.”46 In a world whose order is predicated upon forms, the logic of the order of this world comes from how the forms relate to one another through space and time, where they are placed, and how they unfold. For a morphologist like Haeckel, an organism always subdues its carbon-­philia to the dictates of the whole. Yet there is also a danger to read too much into single illustrations, especially as Haeckel thought of form as an abstract unfolding in four dimensions. Consequently, each illustration is only a single manifestation of a myriad of possible manifestations of a dynamic form. The challenge was to use compositional tools to satisfactorily express this dynamic relationship through the confines of a printed page. a

b

Figure 1.6. Plate 1,”Phaeodarea,” in Figure 1.6a and Plate 67, “Chiroptera,” in Figure 1.6b from Haeckel’s Kunst-­ Formen der Natur. Notice the use of enclosed lines to build architectural complexity in 1.6a and the elaboration of an ornamented frill suggesting a baroque vitality to life in 1.6b. Image courtesy of Wikimedia Commons.

46 · Life on the Line

Haeckel was especially fond of two strategies for suggesting the dynam­ ics of life through changes in form. The first strategy was one already covered: the use of the open curved line. As noted, the curls and swirls of the  lines suggest movement across the page or over time by representing the movement of an organism. The lines make up organic structures, like  the tentacles for the medusa in Figure 1.4, but they also operate as motion lines, like those we see in comic books and graphic novels, where lines appearing behind an object suggest the direction and velocity of the object’s movement. Thereby lines help translate the kinetic movement of the organism to movement in direction over time. This is interesting as the strategies pair how curves are used to add ornamentation with their use to suggest activity. This pairing of ornament and activity has helped to turn these prints into icons of a beaux-­arts style that has made them popular with audiences unfamiliar with Haeckel’s role as a popularizer of Darwin. Perhaps the word “ornament” is not quite precise, though, as present-­day viewers are too quick to think in terms that remove ornaments from functional purposes. Ornaments tend to be anything that is added afterward as a type of decoration or embellishment to the real stuff, the function of the form. To look at these figures in this light, however, is to see them with eyes that have already been informed by the modernist visual world of squares, grids, models, and functions that we will cover in the next chapter. For Haeckel, these embellishments were part and parcel of the chemistry of life. They were constitutive to form and function of the organism and not a superfluous addition. If there is a conception of ornamentation in these prints, it isn’t in their florid craftsmanship, but in the necessity of having to use something as ephemeral as an individual to indicate a more fundamental process. As Olaf Breidbach observes: “Accordingly, the features of the individual examples were only the external form of a natural development which was revealed in the total­ity, but which in the individual example manifested only variations of this shaping process.”47 Again, Focillon observes that “it is still perfectly true that an ornamental style takes shape and exists as such only by virtue of the development of an internal logic. . . .”48 An ornament is an expression of a logic of a specific style and not an embellishment added after the structure has been built. In some illustrations, following the logic of ornamentation allows for one to follow a logic of form. The second technique that Haeckel used to suggest the development of form is the use of the developmental series, where time itself becomes an

Life on the Line · 47

ordering principle in the display of an object unfolding. Haeckel created some of the most infamous developmental series printed in the history of biology. Nick Hopwood published a detailed investigation into the construction and use of Haeckel’s developmental series, so I refer the reader to that work if interested in more detail. In what I cover below, I want to indicate how Haeckel’s work constructing developmental series fits into the arguments that I have been discussing.49 In one of the most notorious examples, Plates 6 and 7 of the Anthropoge­ nie, Haeckel presents a gridded illustration demonstrating Von Baer’s law of specialization of embryological structures through development (see Figure 1.7). These plates are intended to be viewed as a combination of two different types of series. The first is the developmental series (viewed top to bottom) specific to each of the organisms. These illustrations operate like frames from a film, allowing the viewer to see a specific type of embryo develop over time: fish, salamander, tortoise, and so forth. Yet Haeckel composes the illustration to allude to another type of series as well. This other series is viewed from left to right, and it depicts a resemblance of the early embryos in the top row. When viewed in combination with the vertical series, the arrangement suggests how embryos become more highly specialized as they develop, as formal differences between organisms are expressed in the later embryos. Castigated as frauds in how they were constructed and yet frequently reproduced, these images seem to have exposed a nerve in how biology creates its arguments. The charges of fraud appeared shortly after the Anthropogenie was first printed and have more recently been renewed. The charge, that Haeckel fraudulently misrepresented embryos of different species to appear more similar than they actually are, continues to fuel debate. Nick Hopwood and Robert Richards have each spent much time in evaluating the charges of fraud, and I refer you to these authors for a summary of the debate.50 Haeckel himself replied to his critics in the preface to the third edition of the Anthropogenie, and I think his reply goes a long way toward revealing how confusing the visual codes were in these illustrations, even at the time: Many naturalists have especially blamed the diagrammatic figures given in the Anthropogeny. Certain technical embryologists have brought most severe accusations against me on this account, and have advised me to substitute a larger number of elaborated

48 · Life on the Line

figures, as accurate as possible. I, however, consider that diagrams are much more instructive than such figures, especially in popular scientific works. For each simple diagrammatic figure gives only those essential form-­features which it is intended to explain, and omits all those unessential details which in finished, exact figures, generally rather disturb and confuse than instruct and explain. . . . All diagrammatic figures are “inaccurate.”51 Nick Hopwood’s recent archival work is especially informative as it helps reveal the type of work that goes into making a pictorial series intended to compare the developmental sequences from different organisms at different stages. Embryology organized its objects by making developmental series. Specimens, often difficult to obtain at desired stages, were collected and framed as embryos; some had previously been interpreted in very different terms—­for example, as children to come or as waste

Figure 1.7. Plates 6 and 7 of the Anthropogenie, Haeckel’s illustration of von Baer’s law of specialization. The images are arranged to depict increasing variety between organisms as development ensues.

Life on the Line · 49

material. . . . The resulting pictures and models were arranged in developmental order, normal representatives selected, and the series prepared for publication or display.52 So, although these grids purport to show the sequential development of single specimens, as if they were a series of snapshots or a time-­lapse sequence of the changes experienced by a single pluripotent embryo, this is not how the nitty-­gritty of embryology works. Observing a sample most often meant stopping development to represent it. Consequently, even the most seamless, or better yet, especially the most seamless of developmental series are often the compendium of numerous samples, each manipulated to increase the resemblance between stages of development. At first glance, Haeckel’s illustrations of developmental series seem to emanate from a very different aesthetic sensibility than the art nouveau designs presented in Kunst-­Formen der Natur. Whereas the illustrations from Kunst-­Formen subsumed scientific arguments to aesthetic principles, his developmental series used aesthetic technique to make explicit scientific arguments. Still, there are similarities. Both types of prints counter­ balance the intricate detail of the organism with an appreciation of its form. And both prints use the composition of the prints, especially the overall arrangements of the shapes within the image, to suggest how ordering forms can help understand the relationship of living things. So, although one set of prints was intended for a scientifically inclined audience and the other for an artistically inclined audience, they operate visually in some of the same ways. In both cases, details provide a sense of veracity by giving the sample volume and weight, and the scientific arguments rested on how the forms were presented.

The Racial Politics of Forms The political implications of this are brought closer to home when one looks at branches closer to Homo sapiens in Haeckel’s trees of life. Haeckel’s influence on Social Darwinism has also been written about extensively. Some scholars point to Haeckel’s participation in the Monist League as an enabling condition for the rise of racial hygiene as it equipped a form of scientific materialism;53 others point to the racial context of the time and situate Haeckel as a much more benign figure in the history of scientific racism.54 It is true that Haeckel did not prescribe an explicitly racist

50 · Life on the Line

ideology or publicly lend support to eugenic practices, and the popularity of his writings ensured that they were used for a variety of political ends. However, his tendency to subsume an evolutionary pressure for organisms to vary to an evolutionary pressure to unify variety in the name of progress contributed to solidifying racial prejudices into scientific racism. In The Emergence of Genetic Rationality I went into great detail about how narrative structures around evolution facilitated thinking of race, class, and gender in recapitulationist terms. In the short discussion that follows, I would like to gesture to how a similar type of narrative logic emerges in Haeckel’s illustrations on racial categories, thus subsuming his views on the importance of diversity in life to an overall logic of progression to perfection, the ur-­type as we saw previously. This focus on the political implications of forms will be developed more thoroughly in chapters 2 and 3. I hope to quickly demonstrate below how formal relationships possess a politics. Take for instance this often-­reproduced illustration from the Anthropogenie (Figure 1.8). In it we see the same emphasis on organic unity in composition as we saw in many of Haeckel’s other compositions (such as the circular arrangement found in the illustration of the cover of Kunst-­ Formen der Natur), but the focus of the illustration has shifted to primates. Although this is not a strictly recapitulationist depiction in that it doesn’t necessarily encourage a specific linear reading of the relationship of the primates, the drawing does involve a grouping of like forms. In the illustration, the African human is presented as more like other apes rather than other humans. The arrangement of the lithograph is comparable to the lithographs used in Der Kunst-­formen: a highly stylized tableau incorporating the naturalistic elements of place, the tree for instance, and an overall arrangement intended to highlight the organic unity of the composition. The same elements appear in this print as we saw in earlier prints. The open curved lines of the limbs of the tree and the limbs of the primates suggest a potential for movement and variety, but this potential is ultimately subsumed to the overall unity of the composition where the circular lines of the overall composition, especially as expressed in how the primate limbs are positioned, tend to point to each other. The politics of the tableaux come from what is grouped together and what is left out. By placing an African in the tree with the other higher primates, Haeckel indicates that humans need to be included in the species tree of primates. By not placing white-­skinned humans within the composition, Haeckel eschews a politics of inclusion for a politics of white exceptionalism. This

Life on the Line · 51

Figure 1.8. Plate 11 of the Anthropogenie, from the 1874 edition. A tableau illustrating an organic composition of primates. This type of formal organization emphasizes the unity of the elements portrayed and heightens differences with elements not included in the group, such as European Caucasians.

allowed Haeckel to make the argument that humans were related in form to other animals without directly confronting European sensibilities with the implications of that argument. I think it is important to read these images alongside Anne Harrington’s work tracing the rise of holism in German culture in the early twentieth century. In Reenchanted Science: Holism in German Culture from Wilhelm II to Hitler, Harrington suggests that Haeckel, Darwin, and even the more mechanistically inclined Wilhelm Roux “were . . . haunted by an image of a fractured Germany and were motivated by a desire to discover conditions under which some sort of synthesis and integration could be imagined and lived.”55 Harrington is also careful to construct her history to demonstrate the many consequences of holism in German thought and wisely warns that there is a danger of constructing simplistic causalities, since

52 · Life on the Line

organicism as a school of thought can provide a “fund of metaphors.”56 Still, I think it is important to point out that conceptions of how bodies are organized and how they relate to one another have political consequences. Scientific movements that too strongly seek to identify a purity of forms for life will tend to err toward limiting the overall diversity of types. In other words, we need to ask if there is something in the very formal properties of holism as it imaged racial thinking. This is not to suggest that all holism is racist or that other ways of thinking in terms of the formal relationship of living things are not racist (a point we will argue against in chapter 3); rather, it is important to recognize that all ways of picturing suggest a compositional relationship between living things and thus imply a politics of race. When pictures are created so that the formal constraints of the whole rein in the unruly tendencies of the parts, then the politics of race tend to manifest itself in specific ways. In this case racism is especially expressed through what is included and what is excluded. Compare this illustration with the frontispiece from the original 1868 edition of Natürliche Schöpfungsgeschichte (see Figure 1.9). In this illustration Haeckel uses a grid to arrange the racial relationship of humans as three different series. The grid adds an extra element of control that suggests both inclusion as well as ranking within a series. The racial politics of the grid still relies on placement; it depends, for instance on where in the grid each racial type is presented. The grid, however, allows for a specific type of politics where inclusion exists but inequality arises in other types of relationships. In this case, for instance, higher primates are considered as related by family, but the grid provides a way to limit that relationship by ordering it. The grid can be read as a degeneration of type from the upper left panel to the lower right panel. The racism presented in this illustration is subtly different from the one in Figure 1.8. It is not so much a racism predicated on inclusion versus exclusion but on a hierarchy of placement, what I will be referring to in later chapters as a logic of regulation. We will find that when grids eschew simple hierarchies, they allow for a more multivariable, or intersectional, view that discerns how all forms of participation are not equal. Organic wholes tend to lose that refinement through an emphasis on what is and what isn’t a part of an organized whole. The expression of racism through regulation is especially important to remember in a twenty-­first-­century racial environment where combatting racism through inclusion shouldn’t be confused with combatting racism through a prolonged investigation into how interactions are regulated to reify racial categories.

Life on the Line · 53

Figure 1.9. The frontispiece from the original 1868 edition of Natürliche Schöpfungsgeschichte. In this illustration Haeckel uses a grid to argue for a linear hierarchy of evolutionary development among primates. As we will argue later, he subsumes the relative autonomy of the grid’s modules to support a racist narrative hierarchy about the development of the races.

The rest of this manuscript will build on the observation that despite the seeming continuity of the aims of morphology and evolutionary and developmental biology, the way that organisms are conceived in each of these traditions is profoundly different. Whereas morphology appealed to the importance of forms to order biological relationships, evolutionary and developmental biology emptied forms of their volumes to construct a more abstract set of relationships among organisms. These were series composed by regulating specific elements of a relationship. This is a very different type of formalism from the Kantian-inspired organic formalism of Kunst-­Formen der Natur. The architectural historian Sanford Kwinter called this a “true formalism”: “What I call true formalism refers to any method that diagrams the proliferation of fundamental resonances and

54 · Life on the Line

demonstrates how these accumulate into figures of order and shape.”57 This type of formalism eschews a teleological emphasis on the overall organism to replace it with an emergent logic of resonances interacting with one another. This new type of formalism brings new ways to conceive of how organisms relate to their self over time (through development) as well as new ways to think about how organisms relate to one another (through evolution). It also brings an appreciation for how regulations have shaped hierarchies of power in the twentieth and twenty-­first century. And finally, this reliance on form through regulation would further entwine the relationship of biology to highly mechanized, even computerized, views of life. This way of conceiving of forms conveys a way of making living forms computable.58 The following chapters will use the lessons from aesthetics presented here to understand the scientific and political implications of formalisms built on regulations. Finally, this chapter operates not only as a point of departure but as a caution. There may not be a type of aesthetics that is, in and of itself, more equitable than others. In fact, there are many ways to draw life and living things, and each of these ways highlights a specific set of presuppositions on what it means to be alive. There are means of formulating one’s arrangements of living things, however, that are less oppressive and perhaps even more creative. The answer is not to stop thinking in terms of biology; it is to keep interrogating how one’s biology operates in tandem with one’s aesthetics and politics of life.

2

Envisioning Grids

The grid, simply a set of lines that cross each other to form rectangles, has become so fundamental to the way we think, imagine, plan, discover, and organize that it is hard to imagine life without it. Architecture, industrial production, advertisements, and scientific images all use grids to construct things. In this chapter, I will suggest that grids not only pervade contemporary life, they provide a way to understand how consumer experiences informed popular and scientific discourses as well. So, although Haeckel used grids in his illustrations of developmental sequences, it isn’t until the development of industrialized mass communication of the twentieth century that we have a fully exploited aesthetics of grids in culture. Understanding this change, however, requires us to be more exact in describing how users interact with grids as well as seeing how grids were utilized in an economy based on selling goods through mass distribution. In the book, Haeckel’s Embryos: Images, Evolution, and Fraud, Nick Hopwood skillfully traces the development of Haeckel’s embryological images from the display of a sequence of images showing the development of a single organism to the use of an “imperial grid” that compared the developmental sequences of multiple organisms (see for instance my description of Figure 1.7 in chapter 1). “The biggest innovation was the invention of the grid,” writes Hopwood. What makes grids such a powerful visualization tool is that they are “an aid to comparative seeing of pairs, tableaus, series, arrays, and overlays which helped eyes detect similarities and differences, and so distinguish patterns.”1 Even though Hopwood recognizes Haeckel’s use of the grid as innovative, Hopwood also saw Haeckel’s use of grids as “embodying problematic assumptions” about embryology.2 Presenting data as a series of comparable images requires providing an image for the square of the grid being compared. The implication of Hopwood’s observation is that some of the controversy generated by the gridded illustrations was due to the inflexibility of the grid as a tool for comparing changes in forms across species. A succinct but overly simplified version of 55

56 · Envisioning Grids

this argument might be that a comparison of the complexities of development across species can’t be reduced to a grid. Life is too complex, perhaps too organically connected, to have its changes in forms quickly compared as panels on a grid. I think Hopwood’s observation on Haeckel’s novel and problematic use of grids is insightful. If the way that we illustrate life has scientific, aesthetic, and political consequences, then it is important to try to understand how the most pervasive aesthetic for illustrating life in the twentieth century, the use of grids and modules, contributes to how organisms, species, populations, and ecosystems are imagined. This chapter explores why grids are useful for graphic designers and then seeks to understand how the use of grids changes ideas about regulation. Although I will point to how grids are used in the biological sciences, it isn’t until chapter 4 that I investigate how a modular sense of design permeates the way bodies were conceived at the end of the twentieth century. The balance of the chapter, however, will move between media theoretical accounts on the role of images in the twentieth century and a historical account on the use of grids in visual design. I chose this narrative tactic to accomplish three main goals: (1) to investigate why grids became such an important part of industrial visual culture, (2) to show that the ways that grids are deployed reflects industrial qualities of precision and interchangeability but also allows for the crea­ tion of complex forms from homogenous parts, and (3) to demonstrate how scientific communications use grids in their page layout and illustration practices. If there is a visual legacy to a biological science of modularity it draws more heavily from a twentieth century aesthetic of abstract grids than from a nineteenth century aesthetic of wholes.

Using Grids to Picture Development It is difficult to overemphasize how much analytical work grids perform in the illustrations of life during the early twenty-­first century. As we will see in this chapter, grids define the layout of a page in a scientific journal, delimit functionalities for a scientific website, and can even train the eye of a viewer to look at an illustration in a specific way. Understanding how images work in biology in the twenty-­first century must, at some point, wrestle with all the analytical work that grids perform for twenty-­first century thought. Let’s begin our analysis of the epistemological work of grids by comparing two grids used by Nobel Prize–­winning developmental biolo­

Envisioning Grids · 57

gist, Christiane Nüsslein-­Volhard. In Figure 2.1a, Nüsslein-­Volhard uses a grid to present a narrative. This is a familiar account for developmental biologists as it tells the story of embryonic development of the zebrafish. Presented at her 1995 Nobel Prize lecture, the grid presents a series of panels that trace the development of the larva from a four-­cell phase to a larva formed enough to possess the capacity to swim. This grid tells the story of how a relatively undifferentiated structure, such as a fertilized egg, can increase in complexity to look like something that we would recognize as an animal. Familiar organs, such as a backbone, eyes, gills, and a tail, emerge and give the organism a recognizable form and the capacity to swim. It is important to point out, however, that this story is not the streamlined narrative of a single heroic zebrafish, but a reconstruction of what the development of this zebrafish should look like as constructed from samples taken from multiple embryos. As Nüsslein-­ Volhard notes, of “the embryos were dissected out of the chorion for photography.” Consequently, we are comparing different zebrafish embryos stopped at successive developmental periods in such a way as to tell a specific story about how development occurs. We have no indication on what criteria the embryos were chosen. Were some discarded? Were some manipulated in a way to provide continuity? Figure 2.1 is a very good example of how development biologists use grids to tell a story over time. The way that we as viewers are encouraged to read this narrative is similar to how we might read the panels in a comic book, from left to right and from top to bottom. In Figure 2.1b, I’ve overlaid a design grid on this illustration to help readers identify different structures in a gridded layout. Readers will note that grids are constructed from columns (see part a) and rows (see part d) placed next to each other. These allow for greater control in the placement of vertical and horizontal elements within the illustration. Readers will also note that the intersection of columns and rows creates specific “modules” (which I will also call panels, see part b) providing the primary division of space for the grid. All rows and columns are made up of modules. Although some modules are relatively autonomous from other modules (again, see how part b provides an illustration that could stand alone in other contexts). Some modules blend into other modules to form a “specific visual field” comprised of many parts of the grid (see part d where three modules are used to construct an illustration of a five-­day old zebra­ fish). As I will cover in more detail in chapter 4, it is the semiautonomy of

Figure 2.1. Figure 2.1a is “Embryonic Development of the Zebrafish” from the Nobel lecture“The Identification of Genes Controlling Development in Flies and Fishes” by Christiane Nüsslein-­Volhard, 1995. This illustration documents the development of a zebrafish embryo from one hour after fertilization to five days. The uniformity of the grid suggests that we are witnessing the development of a single zebrafish over time, when in fact, this illustration was most likely created as a composite of many different embryos. From Nobel Lectures, Physiology or Medicine 1991–­1995, Nils Ringertz, ed. (Singapore: World Scientific Publishing Co., 1997). Figure 2.1b is a grid overlay on “Embryonic Development of the Zebrafish” demonstrating different parts of design grids (figure adapted from Timothy Samara, Making and Breaking the Grid, 25). Grid part a is called a “column” and it is used to provide vertical alignment to the illustration. Columns can be equal or varied in widths, depending on how the designer wants to arrange the spaces in the grid. Grid part b is known as a “module” or a “panel” and defines the individual unit of space for the grid. Modules can be aligned to form the columns and rows of the grid. Part c is known as a “row.” Rows are composed of horizontal “flow lines” that help organize the horizontal elements of a grid. Part d is known as a “spatial zone.” Spatial zones are groups of modules organized to create a specific visual field. In this case, the modules are grouped to form a row in the grid illustrating a single “swimming larva” of a zebrafish.

a

b

Envisioning Grids · 59

the modules in the grid that allows for them to be useful elements of control. Different modules can be ordered in different ways to provide many types of meaningful associations. This is illustrated by Figure 2.2, which presents an illustration from the paper that helped Nüsslein-­Volhard and her coauthor, Eric Wieschaus, secure the Nobel Prize. Here grids are used for a very different purpose. In this illustration from the 1980 paper, “Mutations Affecting Segment Number and Polarity in Drosophila,” Nüsslein-­Volhard and Wieschaus arrange a series of mutant drosophila larva in a grid so that readers can easily compare the forms and structures of the different mutants. The illustration is comprised of eight modules, where each module compares a specific deletion mutant with a wild-­type fly larva. The comparison emphasizes how a specific mutation of a regulatory molecule drastically alters the emerging form of the fly. As opposed to Figure 2.1, which intends to present a narrative of development with a single outcome—­normal development, the overall effect of Figure 2.2 is to compare a variety of outcomes. Effectively, the illustration sets up the defining question for the textual exegesis

Figure 2.2. Figure 1 from Christiane Nüsslein-­Volhard and Eric Wieschaus, “Mutations Affecting Segment Number and Polarity in Drosophila,” Nature 287, no. 5785 (1980): 796. Notice Nüsslein-­Volhard and Wieschaus’s use of the grid to show diversity amongst the drosophila larval mutants. Image courtesy of Springer Nature Publishing.

60 · Envisioning Grids

of the paper: how can such a varied number of outcomes derive from a singular process? Nüsslein-­Volhard and Wieschaus then answer this question by showing how this pattern of deletion mutants can be explained by envisioning the developing drosophila embryo as a grid itself, where the embryo is formed “by the transition from a single field into a repeated pattern of homologous smaller subfields.”3 A more in-­depth investigation into how organisms also began to be conceived as grids will be the subject of chapter 4. The important point now is that although grids are used in both illustrations, the grids do very different types of analytical work. In the first case, the grid is used to tell a narrative of normal development. In the second case, the grid is used to present a series of different outcomes in the hopes of identifying a unifying principle from variations that don’t fit normal development. Now let’s look at a third grid and see how it functions to make an argument for the field of evolutionary and developmental biology (evo devo). Figure 2.3 is an illustration from Lewis Held Jr.’s book How the Snake Lost Its Legs entitled “How Bilaterians Use Hox Genes to Specify Area Codes along Their Anterior-­Posterior (A-­P) Axis.” The grid in this illustration is harder to discern as this grid is an assemblage of grids laid on top of one another. This is a type of grid that designers call a “hierarchical grid,” where the grids “are based more on an intuitive placement of alignments customized to the various proportions of the elements, rather than on regu­lar repeated intervals.”4 Held uses the hierarchical grid’s flexibility in module placement to present an especially complex and sweeping vision of evolution and development. Some of the comparisons that allow him to do this are already familiar, such as using grids as narrative, where an organism is being compared to itself over time, and using grids as a display of variance, where samples are compared with one another. At least two other types of comparisons are important for Held as well: a comparison between existential scales (the scale of molecules and the scale of organisms) and the comparison between representational types (an abstract schematic versus a picture of an organism). Perhaps most importantly, the fact that Held presents them in one illustration allows him to show how an understanding of development and evolution depends on understanding how all these comparisons inform one another.5 The hinge that makes all Held’s comparisons possible is that scientists found that the single class of genes, Hox genes, differentially regulated the expression of other genes in a remarkable number of organisms.6 These

Envisioning Grids · 61

Figure 2.3. An illustration of Hox organization from How the Snake Lost Its Legs. Lewis I. Held Jr., How the Snake Lost Its Legs: Curious Tales from the Frontier of Evo-­Devo (Cambridge, UK: Cambridge University Press, 2014), 2.

genes that regulate other genes have secured an especially important place in the discipline of evolutionary and developmental biology. Whereas genes have long been understood to create proteins for cell function, it was never understood why some genes produced proteins in some tissues at a specific time whereas other genes produced proteins at other times. If cells from both tissues had the same genes, then how was all this genetic expression coordinated? Understanding how genes were regulated, as opposed to just what they coded, proved key for understanding the genesis of animal

62 · Envisioning Grids

forms. The ubiquity of Hox gene expression helped them become one of the most highly studied systems for thinking genetic regulation in the late twentieth century. Held’s illustration is an attempt to synthesize this in­ formation into a coherent view of how evolutionary and developmental biology operates in tandem. The vertical axis of the illustration is divided into three zones, where each zone depicts a specific organism (see Figure 2.4a): at the top is a human (labeled as “a” on the left side of the diagram), at the bottom is a fruit fly (“c”), and an urbilateran is in the middle (“b”). Held’s “urbilateran” is a purely speculative creature, conjured so that he can think about the evolutionary precursors for flies and humans. The horizontal axis of the illustration has three components. The first component presents the nar­ rative of typical development from embryo to adult organism (see Figure 2.4b for an example of how this works for humans). In this case, however, Held has placed a diagram demonstrating how gradients of regulatory molecules help regulate the formation of each segment of the organisms. The second part of the horizontal axis compares the sequential arrange­ ment of Hox genes with the overall form of the adult organism (see Fig­ ure 2.4c for an example of how this works for humans). Held’s decision to use a hierarchical grid for the design of the illustration allows him to pro­ vide an overall view of how development and evolution can work together, even while it preserves some crucial differences in how this relationship may play out in different organisms. Trying to work out the precise use of grids in such a complex illustra­ tion can be a bit overwhelming. Therefore, I’ve created Figure 2.4, which overlays specific grids on the illustration to help clarify how Held makes specific comparisons. Each panel of Figure 2.4 presents one of the four common comparisons biologists often use grids for as mentioned above: 1. The use of grids to make evolutionary comparisons between organisms. In Figure 2.4a I’ve highlighted the three organisms that Held intends to compare in his figure. Two of these organisms are well known, the fruit fly and the human, while the third is entirely conjectural, the urbilateran, as there is no specific fos­ sil evidence for its existence. The tripartite division of this grid, then, creates a third panel that links the other two panels into an evolutionary argument. This comparison is highlighted by

a

b

c

d

Figure 2.4. A demonstration of how Held uses an open-­ended, or hierarchical, grid system to make the argument for the importance of Hox genes in the genesis of animal form. Figure 2.4a shows how Held compares three types of organisms: humans, fruit flies, and the speculative urbilateran. Figure 2.4b shows how the same organism is arranged similarly despite its different appearances over its life course. Figure 2.4c compares how the arrangement of genes on a chromosome can affect the physiological arrangement of segments in the adult organism. Figure 2.4d compares a schematic rendering of animal segmentation with a realistic drawing of the adult form of the animal. An illustration of Hox organization from How the Snake Lost Its Legs. Lewis I. Held Jr., How the Snake Lost Its Legs: Curious Tales from the Frontier of Evo-Devo (Cambridge, UK: Cambridge University Press, 2014), 2.

64 · Envisioning Grids

Held’s placement of arrows with the label “evolution” emanating from the urbilateran to the other organisms. 2. The use of grids to illustrate the passage of a single organism over time. This use of grids in evolutionary and developmental biologi­ cal arguments emphasizes the change of a single organism over time (such as we saw in Figure 2.1 earlier). Figure 2.4b compares the development of somites, or body segments, in the human embryo to the spinal column of an adult. This comparison em­ phasizes how the placement of segments is conserved from fetus to adult. (I could have just as easily emphasized the segmentation of the fruit fly in this comparison.) The important point is that the comparison suggests a conservation of segmentation in hu­ mans and flies despite the formal changes in the organism as they mature into adults. This comparison could not be made with the urbilateran as no details of its development are known. 3. The use of grids to illustrate changes in scale. The grid in Figure 2.4c compares the organization of adult human bodies with the organization of human Hox genes. In doing so it urges viewers to connect the genetic structure of the Hox genes with the de­ velopment of specific organs during development across vastly different existential scales. (Again, I’ve only compared the genes and the segments of the adult human, but it could be extended to the fruit fly as well.) 4. The use of grids to illustrate changes in representational practices. In Figure 2.4d I compare a representational drawing of a fly to a highly abstract drawing of segment placement. The represen­ tational drawing, in this case a drawing intended to look like an organism, is included to help the viewer locate each of the segments on a picture of a fly while the abstract drawing is used to help the viewer see how the segments relate to one another. This type of comparison between abstract schematic and repre­ sentational illustration is used frequently in late twentieth-­and early twenty-­first-­century biology. Figure 2.4c uses this com­ parative technique as well. Although there are even more comparisons urged by Held in this single diagram, I think the reader now has a sense for how useful an overall grid format can be for creating biological arguments. The fecundity of the grid,

Envisioning Grids · 65

in this case, comes from its open-­ended arrangement that allows for multiple comparative analyses. The overall effect is to supply a complex vision of how the seemingly disparate organisms of flies and humans are evolutionarily, developmentally, and genetically related without being identical. At the heart of this vision is the claim that Hox genes help organize these two different forms of bodies, flies and humans, in a similar but not identical way. It is not that humans and flies both have wings, for instance, but that humans and flies both have appendages in similar areas of the body. Grids may be useful visual comparative devices but they certainly aren’t epistemologically or ontologically neutral. As I will argue more explicitly in chapter 4, the use of grids for showing how organisms come into form changes how scientists view organisms. By allowing for easy comparisons between different parts of organisms, grids establish new ways of thinking about how organisms may be related. The use of grids in Held’s illustration, for instance, allows viewers to see how human and insect segmentation may be more alike than originally supposed. Even just the act of elevating comparative cases encourages finding less visible types of commonalities between organisms than a simple comparison of adult forms. Organic form in the twenty-­first century appears as a consequence and not a cause of the fluid relationship among genetic, molecular, and organismic processes. The use of grids in illustrations is not culturally or politically neutral either. As a brief history of the use of grids in illustration and publishing will show, grids make it especially easy for publishers to set aside spaces for advertising and illustrations. They also make it easier to scale particular illustrations to fit a specific space on the page. Both developments have allowed the advertising industry to flourish by providing a means to plug advertisements into the layout of the page without disrupting it. The grid’s flexibility as a design tool encourages us to think differently about “control” in the twenty-­first century. We usually think about grids as deterministic elements of control, where you are put in your place and you stay there. But fully appreciating the usefulness of grids demands that we hold an appreciation for their ability to control things with an appreciation of how they help construct new types of relationships. Take for instance, the open-­ended grid used by Held in Figure 2.3. By layering grids over one another, Held creates a potential for numerous types of associations. No longer are we forced to follow the single narrative logic of development as presented in Figure 2.1, we are now encouraged to see how a narrative logic can work by incorporating a diversity of evolutionary outcomes, such

66 · Envisioning Grids

as the final forms of a fly, human, or urbilateran. The control of a complex series of grids is not always the simple assurance for creating a specific logic outcome for its viewers, it can now also mean displaying how a field of potential outcomes can be procedurally related. It will become clear as the book progresses, that this new open-­ended idea of control is not necessarily more liberating.7 In fact, multiple elements of visual control are much harder to locate and consciously resist than the straightforward control of a single outcome. As I will treat in more depth in chapter 3, just as these open-­ended types of grids change what it means to be in control, they also change what it might mean to be liberated from control. Liberation no longer means leaving the grid by breaking the chains of its constraints. It now means using one set of constraints against another to create new types of relationships and even new types of possibilities. Living one’s life in the grid often means using grids to go elsewhere, as opposed to escaping all notions of gridded life through a return to organic holism.8 Developing these arguments, however, depends on a more thorough investigation into what it means to engage with a surface as a viewer and how grids exploit that experience. Developing this argument will involve interweaving two important threads of analysis, a phenomenology of surfaces as developed by media theorist Vilém Flusser and a brief history of the use of grids in publishing with special attention paid to scientific publishing. Although this transdisciplinary analysis begins in the present chapter, the full force of this view of grids won’t develop until the final chapters of the book, as new layers of analysis are added. The idea is not to suggest a single antidote or corrective for the ubiquity of grids, but to more thoroughly explore the complexity and paradoxes of lives in relation to grids. For now, however, let’s begin where all published images begin, through an exploration of the strange phenomenological dynamics of images on surfaces.

The Magic of the Surface: Vilém Flusser One of the most important challenges for understanding how forms perform epistemological and political work is to identify how different forms engage viewers in different ways.9 There will be two aspects of my analyses of surfaces, and I will develop both by working through the ideas of media theorist, Vilém Flusser. I will begin by interrogating how images and texts

Envisioning Grids · 67

inform viewers differently. I will then historicize this analysis by looking at how political economic forces changed how surfaces were produced and what this means for understanding how they communicate information to viewers. In his short but fecund essay entitled “Line and Surface,” Vilém Flusser argues that lines of texts engage viewers differently than images. When we read texts, argues Flusser, we train the eye to move in a specific way. Reading English, for instance, requires that “we follow the text of a line from left to right” and then when finished with a line, we must “jump from line to line from above to below” to finish the train of thought. When we finish with a page, we must turn the pages in a specific way, “from left to right” to understand how the author builds her argument.10 Making sense of images, however, requires looking at a page differently. According to Flusser, looking at an image requires we first “seize the totality of the picture at a glance, so to speak, and then proceed to analyze.”11 When looking at an image, a viewer must first apprehend what may be happening in the image and only then can one begin to understand how the compositional elements of the image trains one’s eyes across the page. This would be like having to comprehend all the events on a page of text all at once, before deciding how best to read the page. This seemingly small difference in how we apprehend texts and images makes a profound difference in how we treat them: “This gives us the following difference between reading written lines and pictures: we follow the written text if we want to get at its message, but in pictures we may get the message first, and then try to decompose it.”12 This singular difference in how the eye engages with the content of a page can lead to very different ways of thinking about how readers use publications to help understand the world around them. The most important epistemological difference for Flusser is that lines of texts, in contrast to images, provide a more clearly linear logic. In texts for instance, narratives are created by subjects operating on objects by using verbs. Meanings from texts, then, tend to emphasize causal sequences removed from other associations. Meanings from images, however, participate more in a magical type of meaning making that directly depends on the role of repetition of the image, the context it is presented in, and the details of presentation. This space and time peculiar to the image is none other than the world of magic, a world in which everything is repeated and in

68 · Envisioning Grids

which everything participates in a significant context. Such a world is structurally different from that of the linear world of history in which nothing is repeated and in which everything has causes and will have consequences.13 A picture of an explosion for instance, presents the spectacle of the explosion all at once. The story of an explosion offers a timeline for how the explosion occurred and even provides a platform for thinking about its consequences. For Flusser, writing in 1973, analyzing the dynamics of the interactions of images was becoming especially important for a world increasingly dominated by images printed on surfaces of paper or illuminated on screens. For many, the term “magic” seems antithetical to any type of understanding. In a textually based society, magic appears as an obfuscation of clear and distinct thought, which is often held as a hallmark of scientific reasoning at its best. Flusser posited, however, that the magic of images didn’t deny logic, as images still suggested meaningful associations; rather, images were used to create meanings in ways that were difficult for texts. “Facts are represented more fully by imaginal thought, more clearly by conceptual thought. The messages of imaginal media are richer, and the messages of conceptual [or textual] media are sharper.”14 Surfaces provide a plethora of spatial and relational information at a glance—­an ecology of meaning making, if you will. Authors can use texts on a surface to suggest direct causation and pictures to demonstrate associations. When images and texts are used together, as they often are in the biological sciences, one can appeal to the precision of causal thinking, as well as to provide a mechanism for suggesting how individual causes might fit together in a larger picture. Flusser has been deservedly criticized for creating too strong of a distinction between the properties of lines and surfaces.15 As I will argue a bit later, it is possible to restrict the associative power of images to create more linear depictions of causation. Still, I think the distinction is useful, even if overwrought in practice. One just needs to engage in the image analysis we developed with the work of Ernst Haeckel in chapter 1 to see how important images have been for developing and elaborating on an argument. Images can help one create displays that easily demonstrate how objects relate to one another, whereas texts are especially useful for making specific claims.

Envisioning Grids · 69

The Grid as a Design Strategy What happens when surfaces include grids? Grids allow for controlling how images and texts are arranged on a page. For this reason, printing in general but especially printing with a mechanical printing press has a long history of working with grids.16 Grids are especially useful in designing the layout of a page as it allows for the printer to break a large surface into many smaller surfaces. This allows an artist or designer to concentrate on a smaller area of space to focus on the compositional details of that specific area. A technique for reproducing complex images, for instance, involves overlaying a grid over the complex image and copying each square of that grid to a corresponding square in a grid on a new surface. But grids aren’t just used to transfer images with precision, they also break up space on surfaces in interesting ways. For instance, the use of grids allowed printers to incorporate text with images on a single printed surface. Blocks of type were interspersed with blocks of woodcut images when printing pages. The development of inexpensive publications designed to appeal to mass audiences in the nineteenth century intensified this use of grids to enhance viewers’ experiences by combining texts, illustrations, advertisements, announcements, and amusements in easily printable formats. A grid’s ability to break a surface down into smaller areas was only part of its usefulness. Designers quickly found that grids could also provide a scaffolding for building complexity back up from simple elements. As Armin Vit and Bryony Gomez-­Palacio wrote, “The immediate reaction to the notion of the grid might be to feel constrained, limited, and bound to a boring set of modular columns and horizontal axes. Luckily, nothing could be farther from the truth. . . . The grid is, at its best, an infrastructure upon which to build both complex and austere layouts that enable hierarchy and accessibility through flexibility and consistency.”17 It was this constructivist function of the grid, its potential to enable complex structures, that made them especially important conceptual tools for the display and advertising of manufactured goods. Understanding how surfaces can be used for developing associative meanings is one of the most important lessons for understanding how something as seemingly rigid as a grid can enable a constructivist logic of image making.18 This was one of the conceptual leaps taken by modernist experiments in printing during the mid-­twentieth century. One of the most important of these was Josef Müller-­Brockmann’s 1968 text, Grid Systems in Graphic

70 · Envisioning Grids

Design. Müller-­Brockmann claims that the grid as a “controlling principle” in design was first used in Switzerland in the late 1940s. “This new trend was characterized by a disposition of text and illustrations conceived on strict principles, by uniformity in the layout of pages, and by an objective attitude in the presentation of the subject.”19 By the early 1970s, Massimo Vignelli would be able to exclaim: “All the work I do is based on grids. I can’t design anything without a grid. I am so accustomed to using a grid that I use it for everything, even stationery. The grid provides the tool for quick solutions.”20 Yes, grids provided elements of control for designers, but in doing so, they also allowed for new types of associative meanings.

The Display of Goods This was why the growing twentieth-­century marketing and advertising industry found creative new ways to use grids for product display. Historians such as James Beniger and Susan Strasser have described how the over­production of goods at the turn of the twentieth century—­that is, the capacity and tendency of manufacturers to produce more goods than they could find markets for—­led to new relationships between producers and consumers.21 Beniger, for instance, demonstrates how new informational techniques based on grids, such as standardized forms and advertising, ensured the smooth distribution of goods to larger markets. Strasser, on the other hand, shows how the need to find new markets for products led to “new pacts” with consumers, such as the invention of S&H Green Stamps (a consumer rewards program) and coupons that consumers could clip from one product and redeem for other products. Through devices such as S&H Green Stamps, which could be easily packaged with consumer goods, manufacturers promoted brand loyalty by rewarding customers with the ability to trade the stamps they collected for additional consumable goods. Developing these new relationships often involved creating increasingly novel ways of appealing to customers. This often involved creating new ways to display goods to attract customers and engaging customers in new ways of thinking about how products could be used in their lives. The problem, however, is that manufacturing produces large quantities of identical items. As commentators on industrialization have identified, achieving economies of scale meant selling large amounts of products.22 This often involved establishing new types of arrangements with customers through advertising, product promotion, and visual display. The late nineteenth and early twentieth centuries saw these new arrangements

Envisioning Grids · 71

proliferate as companies experimented with different ways of packaging items to increase the consumption of the goods they were producing. The associative logic of images was central to building engaging landscapes of industrially produced goods. Perhaps the most straightforward example for how grids were used to promote products is the simplest: the stacking of the actual goods for sale into arresting configurations. Take for instance this photo from Jenkins’ Groceteria in Calgary, Alberta, taken around 1945 (Figure 2.5). Grocers proudly stand behind a display of different Kraft products. The display grabs consumers’ attention by changing the scale of a collection of a few single products into a noticeable presence, approximately human size in height and difficult to ignore. Stacking identical containers of mustard, cheese, mayonnaise, and other food items gave a larger scale to previously unnoticeable items.

Figure 2.5. A display made from Kraft products in Jenkins’ Groceteria, in 1945. Notice how large associations are constructed from similar units. This allows for the products to attract the attention of the consumer. Photograph courtesy of the Glenbow Archives, PA-­2453-­232.

72 · Envisioning Grids

This type of stacking, which builds on the associative aspects of imaging, allowed marketers and designers to create distinct and affecting displays from a limited diversity of building blocks. According to Flusser, the magic of images derives from a logic of repetition and ritual. It was the circumstances of the stacking—­how the products were used as modules in a larger assemblage, as opposed to the materials of their construction—­ that lent these displays their visual power. Although these displays were not grids in the strictest sense of the word, their method of construction and intended purpose help illuminate a predominant logic of consumer driven capitalism, the desire to build more complex worlds from uniform manufactured elements.

Comics and the Affective Power of Grids It was in comics, advertising’s even tackier twin, that many of these techniques were applied to developing coherent visual strategies for consistently creating meaning from the partitioned panels of grids. Comic books were developed from the same materials and needs that led to the development of product displays and the gridded surfaces of graphic design and advertising. Comic books emerged as one of these new pacts with consumers to increase product distribution, for the books and their immediate predecessors, the newspaper comic strip, were initially designed to increase market size for other consumer products. During the late nineteenth century, comic strips were added to newspapers to create a readership independent of the actual content of the news: one couldn’t always expect daily events to be newsworthy, but one could expect a comic strip character to persist in her or his exploits. Comics, then, guaranteed the sale of the newspaper on slow news days that might not have attracted readers otherwise. As comic strip writer and artist Milton Caniff describes the development of newspaper strips, “I think, first of all, we were a circulation device.”23 Perhaps anti-­intuitively, comic strips emerged as an important tool for understanding the complex regulating mechanisms for industrial societies, as they helped establish customers for a product. Comic books emerged several decades later, in the 1930s, and played essentially the same role of the earlier strips since they began as a form of repackaging for other consumer goods. Not only were early comic books composed of the same material from earlier journalistic comic strips (thus repackaging the content of newspapers), they were frequently marketed as promotional items

Envisioning Grids · 73

(thus providing packaging for other items). Funnies on Parade #1, which comics historian Mike Benton has described as the “first ‘comic book’ as we know it,”24 was created in order to increase the sale of Procter & Gamble soap products, and it was initially available only to customers who sent in coupons clipped from these Procter & Gamble products.25 Max C. Gaines, the editor of the Funnies on Parade comic book, began his career as a salesman and apparently understood comic books primarily, if not exclusively, as a means for increasing the sales of other consumer items. He thus approached other manufacturers, such as “Milk-­O-­Malt, Canada Dry, Wheatena and Kinney Shoes—­to sponsor publication of another giveaway, or premium comic book.”26 Also, there is an interesting relationship between illustrators who drew ads and comic book artists, as many artists moved between comics and advertising. Neal Adams, Joe Simon, Lou Fine, and Jack Davis, for instance, also pursued careers in the better paying and more upscale advertising illustration trade. They could do this because illustrations for mass-­produced printings required similar skills. The primary artist for Golden Age Captain Marvel, C. C. Beck, especially understood how comics brought together a whole set of skills important for a society increasingly reliant on the use of images: It’s the same form I’ve used when I’ve done things for photographers. It’s the same as a storyboard before they produce a movie. It’s in all forms of advertising. A guy sits there and lays out the ads in sequence so they don’t contradict each other. It’s months before they ever appear. It’s only the word “comic” connected with it that turns people against it.27 Beck’s insights are especially profound. Comics as a product seemed a limited development in mass entertainment; comics as process, however, were key to industrialized entertainment, in that they elucidate how items, moments, and scenes can fit into temporal or spatial assemblages. The danger here, of course, is to think that comics and advertising were the same thing. But this is exactly the point of the growing ubiquity of grids. Saying that comics and advertising both used grids does not mean that they used them for the same purpose. This is, in fact, why grids were so useful. They could be adopted for multiple uses, even within a single industry, such as the comics industry.

74 · Envisioning Grids

Though the success of some of the early comic strips encouraged producers to understand them as consumer items that were marketable apart from other manufactured products, the narrative content of comic books nevertheless continued—­and continues—­to serve as a vector for other consumer goods. The relatively cheap cover price for comic books produced between the 1940s and the 1970s was possible only because of the extensive advertising each book contained: in a typical comic book of roughly thirty-­four pages, more than eleven of those pages—­including the inside of the front cover, and both the inside and outside of the back cover—­contained advertisements. The content of such advertising was remarkably constant for the first fifty years of comic books: cheap gimmick items, such as pepper gum and X-­ray glasses; self-­help manuals that promised a more muscular body or job skills; relatively low-­cost collectible items, such as coins and stamps; and so on (see Figure 2.6). All of these products served no special purpose, per se, and testify to how “A WHOLE TREASURE CHEST OF FUN” was created from the inherently absurd collection of novelty items. Comics provided their own innovations to the associative logic of the grid that we saw in display and advertising, and they began to develop visual techniques for leading the reader’s eyes across the page in a specific way. In doing so, comic books not only formalized a series of strategies for harnessing the associative power of images, they also used the repetitive logic of industrialization to create whole new mythical cosmologies while supplying the clues for how to navigate these new universes in the arrangement of forms on a surface of girds. Art Spiegelman and Chip Kidd’s analysis of Jack Cole’s Plastic Man shows how this logic plays out in the panels of a comic book page (Figure 2.7). Specifically, they point out how Cole used the flexibility of Plastic Man’s body to lead the eye of the viewer through each panel and across the page. The arrangement of the panels on the page thus adds a dimension to the reading experience that cannot be reduced to the dynamics in each panel. The associative logic of a page composed of panels is so great that artists and authors often need to develop strong visual elements to help direct the reader’s eyes across the page in a specific way. According to Art Spiegelman, “The art ricochets like a racquet ball slammed full force in a closet. Your eye, however, is guided as if it were a skillfully controlled pinball, often by Plastic Man himself acting as a compositional device. His distended body is an arrow pointing out the sights as it hurtles through time.”28

Figure 2.6. Advertisement page from The Fantastic Four #94 (1970). The panel structure allows a rough division between different kinds of novelties: most of the items in the leftmost panels promise new abilities (for example, throwing one’s voice or seeing through clothing); most of the items in the central panels advertise techniques and tools by means of which others can be deceived without their knowledge (such as the secret book safe or secret spy scope); while most of the items in the right panels promise an ability to amaze or surprise others (for example, the joy buzzer and monster-­sized monsters).

a

Figure 2.7. An illustration of a page of panels from the Plastic Man comic (2.7a) with an accompanying visual analysis by Chip Kidd and Art Spiegelman (2.7b). Notice how the stretchiness of Plastic Man’s body acts as a device to lead the eyes of the reader through the proper panel order. Art Spiegelman and Chip Kidd, Jack Cole and Plastic Man: Forms Stretched to their Limits! (New York: DC Comics, 2001), 38. Plastic ManTM and copyright DC Comics. All rights reserved.

b

78 · Envisioning Grids

The effect is that the borders found on the page of a Plastic Man comic aren’t just hard and fast divisions that divide space, they are a tool for constituting relationships through a combination of their internal composition and their arrangement with one another. In fact, this relationship of possible associations on the page is so rich that it is probably best to view the use of compositional elements as giving clues to readers to read a page in a specific way. Surfaces are so fecund that composition is a necessary means for limiting associations even as it promotes them. The narrative relationship, or the specific sequence of panels, is not only created from surfaces, it also must be regulated in a specific way. The incredible fecundity of possible meanings on a page inspired Marshal McLuhan to describe advertising and comics as a type of game. “The comic strip and the ad, then, both belong to the world of games, to the world of models and extensions of situations elsewhere.”29 This gaming logic comes from the power of images to provide multiple associations and thus suggest the potential for different outcomes. This type of modeling is very useful for thinking in terms of places we could inhabit, as opposed to finding out more about the places that we currently inhabit. Another spin of the dial, a new card, or a role of the dice is all that is needed to intervene and create a different type of future, a future elsewhere. This gaming logic of comics directly plays on the associative potentials of space, where the differential between a desired outcome in relationship to a potential set of outcomes is what is being gamed. Ads and comics weren’t intended to send a specific message to a reader, or even to tell the reader that this is how the world is. Rather, ads and comics were intended to display a model for how things come into being by imaginatively engaging readers so that they can see new possibilities. It is important to note that these outcomes are already limited in that they are contained within the possibilities of relationships of the images on the page, and much more importantly, they are already a part of the technical apparatus that created them. A reader engaging in the gaming logic of ads and comics needs to realize that they are gaming within the limits of a specific political economy, despite the enticements that the game itself offers. Jackson Lears appeals to the gaming logic of comics and advertising when he suggests that advertising is similar to “animism”: “the desire to endow objects with symbolic, perhaps even spiritual, significance.”30 In his book Fables of Abundance, Lears suggests advertising is only one of many manifestations of the hope for abundance that have marked much of western society, from the carnival to the trade show. Advertising always has

Envisioning Grids · 79

to deal with the new and excessive; it needs to persuade you to purchase something that you might not have, something that may be novel to the consumer but really is only one of a set of possible outcomes in a game of chance programmed by the outcomes of a consumer society.

The Surfaces of Science Although the advertising and entertainment industry helped develop the rules of engagement for the role of images for envisioning possible outcomes, their effects aren’t just limited to the funny pages. Scientific publications have increasingly appealed to a gridded logic of production to simultaneously make their publications more businesslike as well as more varied. In an article on the history of the visual strategies of the journal Nature, Martin Kemp describes how the masthead of the publication moved from the rustic calligraphy of the late nineteenth century, through the simple and streamlined modernism of the two-­color image in 1958, to the “brisk, up-­to-­date and snappy”31 computer-­generated image of today. Consequently, Kemp articulates how scientific design practice often reflects concerns found in other parts of culture as even though one may be reading a scientific journal, the reader is still participating in the dominant aesthetic values of the day. What is especially interesting about Kemp’s analysis is his recognition for how the dominant aesthetic mid-­twentieth-­ century values embrace the modular aesthetics of the grid itself—­its ability to create complexity from homogenous units: The graphic vehicle is correspondingly clean and business-­like, favouring sanserif typefaces for headers, titles and captions. Its natural visual habitat is the top of a metal desk, beside a computer, or on a modular shelving unit in a modern laboratory.32 Nature, the publication, not only delivered information to its readers, it participated in the development of a society built on the modular principles of the grid. The organizing capabilities of grids allowed scientific publications to provide content specific to the study of science while they opened new types of spaces for advertising. According to the editor of Science at the time, Daele Wolfle, this is what drove the 1955 major makeover of Science that increased the trim size of the publication to a more standardized size (thus increasing its surface area) as well as moving from the bookish

80 · Envisioning Grids

two columns per page to a more magazine and newslike three columns per page: The principal reason for the change of page size is the expectation of greater income—­income that can be used for the larger and stronger, and hence more expensive, editorial staff that can make Science into a more useful journal. Like many another periodical, Science depends on advertising revenue for a fair portion of its budget. For a wide range of advertisers, Science is a good advertising medium. But many of these advertisers were handicapped, and sometimes frightened away, by our nonstandard page size, for many already designed advertisements had to be remade to fit a Science page. Advertisers, advertising space salesmen, printers, and publishers were unanimous in advising the change and in predicting a larger advertising revenue if we made it possible for companies that advertise to scientists to use in Science the standard-­size plates that can also be used in other journals.33 Making these changes not only increased the types of presentation strategies by offering new ways to construct a grid across the page, they also helped science fit the standard grids of other publications, making it easier for advertisers to adapt their messages for publication. This allowed advertisers to maintain the integrity of the design of the ad, essentially giving them a means to plug the ad into a variety of different publications no matter the discipline they catered to. As long as they used grids to order their page, that is. Some scientific societies hired visual designers to help them design coherent visual strategies to raise their institutional profile. “We were asked by the New York Botanical Garden to develop a corporate identity,” boasts Massimo Vignelli in his Grids: Their Meaning and Use for Federal Designers. For this task Vignelli turned to his cherished tool for construction, the grid, to design the layout of the page for “very varied” products demanded by the botanical garden. He used them to design standardized forms for the internal management of the garden as well as stationery, posters, newsletters, and booklets. We can see the different layout possibilities in these applications. Everything is much more disciplined. There is space at the top for

Envisioning Grids · 81

running titles, for headlines, and for everything else. The grid helps to position things. If these things had to be laid out without a grid, it would be really difficult to know where to position them. This kind of tool provides a frame of reference, a reference point.34 The ability of grids to simultaneously order and create allowed Vignelli to produce a unified sense of design across platforms demanding different forms of use. With the increase in the use of software programs for designing and layout, grids have become even more ubiquitous but not necessarily easier to spot. Although the illustration might not look like a grid, the chances are high that grids were used in their construction. As graphic artist and creative director of the Broad Institute of the Massachusetts Institute of Technology and Harvard, Bang Wong wrote in the column on graphic design for the publication Nature Methods entitled “Layout”: Using a grid to aid layout . . . can dramatically streamline the design process by taking the guesswork out of sizing and placing content. Try creating a set of strategically placed guides in Microsoft PowerPoint or Adobe Illustrator before you work. Grids help to anchor content and create stability within a design. They also build consistency between slides that allows the audience to anticipate where content will appear.35 The illustration that Wong provides is telling as it recalls the lessons we learned from Chip Kidd and Art Spiegelman’s analysis of Plastic Man on how to use forms to direct one’s eyes across gridded landscapes (see Figure 2.8). As Wong elegantly states, “Layout is more than adhering to lines of a grid system: it is the process of planning out exactly the journey we want the eyes to travel across the arrangements.” This is the work of those involved in having readers envision how to read the book of nature as well as a superhero comic. For visual designers, it is all in the layout of panels on a surface as this is where, according to Wong, layout happens and “layout underlies everything we do when we communicate visually.” This form of communication is a delicate dance of control and construction, of dream and representation, of fact and absurdity. Understanding the strange and conflicting choreography of this dance demands a bit more exploration in how grids combine restriction and creation in the spatial regulation of society.

82 · Envisioning Grids

Figure 2.8. Bang Wong’s depiction of the use of grids in the layout of scientific publications for Nature Methods. The first section denotes examples of grids for “presentation slides.” Section b shows how to arrange elements according to how they will be viewed. Section c urges designers to surround points with white spaces in the grid. Bang Wong, “Points of View: Layout,” Nature Methods 8, no. 10 (2011): 783. Image courtesy of Springer Nature Publishing.

In the grocery aisle, on the printed page, and in scientific publications, grids provide a mechanism to move between a specific focus on products to a general view of the form of the display. And grids did this before there was a specific theory on the role of regulation in building complex forms from repeated elements. Key to these displays were finding ways to limit the associative logic of images by giving them a specific sense of direction, a type of overall order, or form. For most images in the twentieth century this involved developing protocols that ensured that the modules of the grids, their bits, grains, and primary units, were aligned in the correct way.

Envisioning Grids · 83

The Surfaces of the Technical Image Although all images operate through associative logic, Flusser does recognize that mechanically and electronically rendered images fit differently into political economies than hand-­wrought surfaces. This, in turn, changes how they function in creating meaning and how they relate to textual understanding. The key difference is that technical images are produced through apparatuses, and because of this, they are already a product of a scientific culture that turns things into information. Each bit of an electronic image is a product of a command or code. Even the chemical process of photography, for Flusser, is part of a world of informational particles, as the image is rendered unto a surface of grains of silver nitrate reducing it to patterns of information based on the responsiveness of the silver nitrate. “The technical image is an image produced by apparatuses. As apparatuses themselves are the products of applied scientific texts, in the case of technical images one is dealing with the indirect products of scientific texts.”36 Consequently images produced through technical apparatuses introduce a new level of abstraction where the major referent is to a technical culture as opposed to the material world. This means that when we read technical images of the world it is always through their own immanent possibilities of production, or the very special way they have of turning the world into information. A clear example for how a technical image works is the process of half-­ tone reproduction. In half-­tone reproductions, images are printed on surfaces by breaking them down into small dots or specks of ink. The image, then, emerges from the changes in the proportional density of the dots, where densely placed dots create the effect of a dark surface in the image, and less dense dots allow for lighter tones. It is this process of breaking images into particles that Flusser has in mind when he refers to how technical images are products of abstractions. By creating massed patterns of dots, one could vary the darkness of the ink, allowing for the nuanced inclusion of shadows and fields of colors. The use of photomechanical reproduction, much like the use of grids, helped open up the flexibility of the design on the page as it allowed “the staff to manipulate the sizes of the illustration at will,”37 in that it made illustrations easily sizable to the restraints of the grids of published surfaces. The directions for assembling the modules of the grid into visual fields depended on developing means for controlling how the modules were placed. For Flusser, this protocol is never far removed from the political economy

84 · Envisioning Grids

that created technical images in the first place. Technical images are valuable because they offer their creators access to the ability to regulate information. Although specific tools are used to accomplish this, such as a camera or a computer, it is the access to the codes that are built into these apparatuses that are the most important. “One can see from the softest of the apparatus, e.g. political apparatus, what is characteristic of the whole of post-­industrial society: It is not those who own the hard object who have something of value at their disposal but those who control its soft program. The soft symbol, not the hard object, is valuable: a revaluation of all values.”38 It is the soft symbols, or the symbols that operate like programs instead of things, that technical images serve. Soft symbols are the codes and rules that create the patterns that help make sense of abstract bits of information. We need these patterns to make sense of the bits of information, and making sense with these patterns entails constructing images with them. Using soft symbols, however, also supports the values of a society that turns things into bits of information. Constructing an image from a surface requires a set of programs, a protocol for linking the particles into coherent images that can be organized.39 Imaging no longer means revealing fundamental relationships of form, it also means regulating particles to create meaningful patterns. This universe of technical images possessed its own quantum effects of wave/­ particle duality as patterns emerged from a sea of potential images, only to disappear into new configurations: The photographic universe is made up of such little pieces, made up of quanta, and is calculable (calculus = little piece or “particle”)—­an atomized, democratic universe, a jigsaw puzzle. The quantum-­like structure of the photographic universe is not surprising, since it has arisen out of the act of photography, whose quantum-­like character has already been discussed. Yet an examination of the photographic universe allows us to see the deeper reason for the grainy character of all aspects of photography. It reveals, for example, that the atomized, punctuated structure is characteristic of all things relating to apparatus, and that even those camera functions that appear to slide (e.g. film and television pictures) are actually based on punctuated structures. In the world of apparatus, all “waves” are made up of grains, and all “processes” are made up of punctuated situations.40

Envisioning Grids · 85

The universe of technical images is propagated through particles, where waves of images gain meaning by the constant and incessant referral to one another, and the potential for the magical association inherent in all images is securely harnessed through the coded values of industrialized society. Finding meaning in this universe of photographic images is not just an added value of perception, or even an act of revelation, but a constitutive and regulative function of society. It really matters how you put these bits of data back together again. So, although control is built into how we imagine a world, this control is not a determination. The amazing implication of Flusser’s argument is that there is a bit of absurdity, or nonsense, in all forms of technical images despite their very real and very concrete existence. Not only does absurdity come from the inability of a single particle to provide meaning, it also comes from the associative logic of imaginal thought. Technical images create real relationships from abstractions and illusions and thus also cause havoc with logical categories used for textual arguments. The abstract particle universe from which we are emerging has shown us that anything that is not illusory is not anything. This is why we must abandon such categories as true–­false, real–­artificial, or real–­apparent in favor of such categories as concrete–­abstract. The power to envision is the power of drawing the concrete out of the abstract.41 In the age of technical images all types of representation require some element of fabrication. Representation is no longer just a property of a mimetic revealing of the world, it is a fundamentally constructive element that allows the concrete to emerge from a sea of potentials. Perhaps returning to an epistemology of hard categorical distinctions, of clear and distinct thought, is not the best way to approach knowledge-­ making through technical images. Maybe we need to also develop an “associative” epistemology that is flexible enough to see how things come into being, but that also recognizes the conditions that helped produce it. Take my discussion of the grid, for instance. The point has been to identify how the constraints of grids can produce new types of associations. I have attempted to do this, however, by recognizing the role of new forms of capital that enable, even privilege, this type of flexibility. So, we may not be able to easily demarcate depictions (for example, a photograph of a house) from

86 · Envisioning Grids

models (for example, a computer-­aided design of a house) based on the categories of truth and fiction, as both are conjured from particles of information. This absurdity, however, does not make either “untrue” or “unreal.” It just means that truth emerges by carefully weighing how these associations are produced, how they fit together, and how they fit with other sensory experiences. This insight is especially important as it suggests which technical images carry the most value as epistemic objects. Truth is no longer found simply in creating images of how the world is, since this depends on the normative value of representational veracity. Truth is increasingly found in improbable, strange, and even bizarre patterns and the challenge to envision the circumstances under which these patterns could be created. If we look at the Figure I.1 in the Introduction, for instance, of the expression of the gap pair rule gene in a drosophila embryo, it is the highly improbable nature of being able to represent molecular and organismic scales simultaneously that makes this picture informative. The implication is that scientific images do not guarantee certainty about the world nor are they absolutely removed from the world that creates them. Rather, science creates new fantasies while it creates certainty. No longer are fantasy and truth thought to be in an antagonistic relationship with each other, where the establishment of one abolishes the other. Truth and fiction now proceed hand in hand as new scientific findings create new questions, new fears, and new fantastic futures envisioned by imaginal thought. In this sense, science and technology will existentially contract, even as they will un­doubtedly expand exponentially as methods. They will be absorbed by new fields of interest. We will no longer be below science and technology (in “superstition”); rather science and technology will grow beneath us. From now on, superstition will be in images that will grow over us. This is how science and technology will change. They will be subordinate to the computation of images.42 There now is a seed of the spectacular in even the most mundane or depictive illustration. Yet, even this idea of the spectacular has changed as these spectacles of vision don’t point away from the values of a scientific society but back to it. Each technical image supports the values that produced it, such as the

Envisioning Grids · 87

desire to quantify the world by breaking it into particles and then reorganizing these particles to make abstractions concrete. This is as true for the technical images used in science as it is for the technical images used in the entertainment industry. When we render a molecule observable, we conjure a fantasy that reflects its own conditions of production even as it helps one envision a world. Whether this fantasy is true or not is an important but not primary point; the main value of this fantasy comes from its ability to envision abstractions as concrete processes in the process of actualization. As Flusser summed up: “The production of technical images occurs in a field of possibili­ties: in and of themselves, the particles are nothing but possibilities from which something accidentally emerges. ‘Possibility’ is, in other words, the stuff of the universe and the consciousness that is emerg­ing.”43 Although the specific implications for biology will be drawn out more fully in the next three chapters, it is important to note that Flusser’s argument makes it increasingly difficult to disassociate the use of a grid as a presentation tool and the use of grids as a systemic element for understanding all forms of organization. In the universe of technical images, grids don’t just facilitate the representation of things such as organisms as existing outside of grids; they help bring forth a way of looking at the world that ultimately sees things as composed of grids with calculable particles. This is a truth and a fiction of the universe of technical images. They point to the world and to the circumstances of their production at the same time. The significance isn’t so much how I escape the all-­pervasive grid, but how can I use the grid to find the weirdness in the fabric of all reality?

Science Online Google the phrase “science journal” and click on the first entry. This should bring you to www.sciencemag.org, the website for the American Association for the Advancement of Science’s journal. Most people know it simply as Science. When I performed this exercise on June 2, 2014, I found a grid containing advertisements, menus for navigation, the masthead for the publication, access to tools for social networking, a space to log in to your account, a search box, an announcement banner, and a rotating menu of headline stories, where the description for the story was accompanied by a vivid illustration or photograph. When I scrolled down I was greeted with a table of contents containing photographs, audioclips (a podcast),

88 · Envisioning Grids

links to videos, and links to stories containing textual and visual elements. All of these elements are packaged as modules on a grid (see Figure 2.9). At first glance, these grids seem different from the panels of a comic book in that a content management system has been used to order a variety of multimedia, turning each panel in the grid into a window for a new type of experience. However, all technical images already order content in specific ways, and they all have the phenomenological power of panels for opening windows onto a world. As we saw, a comic also maintains a balance between the specificity of a single panel on a page and the general meaning of how panels are associated across the page. The ability to click on a panel to open a new window or webpage should be seen as an extension, or deepening manifestation, of the logic of technical images, where the specificity of looking at a panel can at times seem like the opening up to another world.44 As Flusser notes: This apparently non-­symbolic, objective character of technical images leads whoever looks at them to see them not as images but as windows. Observers thus do not believe them as they do their own eyes. Consequently they do not criticize them as images, but as ways of looking at the world (to the extent that they criticize them at all).45 This point by Flusser is especially interesting in that it suggests that the phenomenological qualities of looking at a specific panel in a comic, or the ability to have a window on the world, can, in some cases, be laden with an epistemology where representations tend to deny their own constructed nature.46 On the Science website, each panel opens unto its own media content, possessing its own narrative, and making specific knowledge claims. The grid allows for the bundling of these potential experiences into a treasure chest of new ways of scientifically understanding the world. Novel experiences are promised because clicking on a specific panel opens a new window on a world. When I was employed to teach a class in visual communications at the University of Washington in the late 1990s, I began to understand how important advertising and comics were for understanding the design sensibility of the newly developed World Wide Web. The first year I taught the class, there was no specific manual on webpage design to help students gain an understanding of how to best organize space on a webpage. Other

Envisioning Grids · 89

Figure 2.9. A screen capture of the website for Science magazine. Notice how grids are used to present a combination of media (including words, movies, and images) and a combination of different stories and advertisements through a unified design format. www.sciencemag.org.

instructors, however, told me that they had used Scott McCloud’s comic book about comic books, Understanding Comics, as their primary design text. McCloud’s book was remarkable as it helped students think about the relationship between text and image and how panels can be used within grids to separate content. As Erik Loyer, a designer for the interactive online humanities publication Vectors, recently wrote, “Since its publication in 1993, interactive media designers have turned to Scott McCloud’s Understanding Comics as a kind of bible of visual communication in the digital realm.”47 Andy Hertzfeld, cocreator of the Macintosh computer, also paid his tribute to McCloud’s book when he acknowledged its importance for creating graphic user interfaces: “One of the most insightful books about designing graphic user interfaces ever written, even though it never discusses the subject directly.”48 The usefulness of this book is not a coincidence. During the twentieth century, panels and grids became highly codified ways for designers to navigate the tensions between a well-­ordered present and the promise of a novel future. The imperatives of technical images, as they were developed

90 · Envisioning Grids

in the advertising and comic book industries, helped to shape most forms of display at the turn of the twenty-­first century. It isn’t so much that we must read our news or our science with the imperatives of mass entertainment, although this is most certainly the case. But industrialization also helped to make images so easily codified that they could easily be adapted to computing. The grid helped to make the world computable. Two important lessons emerge from this analysis that we will build on in later chapters. The first lesson is phenomenological. There is a power to images that defies the linear logic of arguments and offers an aesthetic of how we conceive of science and its content. As we saw in the first chapter, this power was extremely useful in thinking about how different life forms develop and relate to one another. In this chapter, we have seen this power increasingly channeled through the confines of surfaces, grids, and panels. The second point is political-­economic. Technical images harness the power of vision as a form of regulation that constructs experiences from the fragments of industrial and post-­industrial processes. One shouldn’t fall in the trap, though, of thinking of all regulation as inherently injurious. As we saw, a narrative in a comic book is also a way of regulating the attention of the reader, just as an advertisement is a form of regulating desires in conjunction with distribution. Both elements help to give meaning to an inherently absurd universe of technical images. In the next three chapters, we will look at how deeply this type of regulatory logic marked new ways of conceiving how organisms developed and how they relate to one another. In the next chapter, we ask how pests, or those organisms that evade grids, help us understand life in the grid. In chapter 4, we ask what happens when the power of association and regulation found in the visual design of grids is thought to drive the development of organisms. And in chapter 5, we will ask what happens when animation and computer modeling put grids into motion. Each chapter offers insights on the bizarre and very real experience of finding life in the grid.

3

Warped Grids Pests and the Problem of Order

Let’s begin again at the surface, or rather, at a specific surface: a publicity poster for Kurt Neumann’s horror science fiction classic, The Fly (Figure 3.1). Released in 1958 by 20th Century Fox, the movie was produced on a budget of $400,000 and shot in only eighteen days. It was an immediate success grossing $34,000 on its opening day.1 This poster was one of many created by 20th Century Fox to attract viewers to theaters by creating a “buzz.”2 Based on an award-­winning story of the same title published in Playboy the year before, The Fly is a movie of a teleportation experiment gone horribly wrong.3 A scientist working for a Montreal-­based technology firm experiments with the use of nuclear power to teleport goods and people. The moment that he tries to teleport himself, a fly slips into the chamber with the scientist. The teleportation device confuses the integrity of the organisms and switches the head and the right front limb of the fly with those of the man. What results are a man-­sized man/fly (the body of a human with the limb and head of a fly) and a fly-­sized fly/man (the body of a fly with a human front limb and head). Produced at a time when electronic communications helped fuel a remarkable growth in consumer spending, the movie provocatively toys with the relationship of consumption, the role of regulation in a complex society, and the acceptable limits of the mutability of life. In doing so, it offers an opportunity to demonstrate how thinking in terms of grids adds new insights into how life was regulated in the twentieth century. Take, for instance, the dominant image in the movie poster: a woman, the scientist’s wife, Helene Delambre, screaming in terror as she stands behind a window screen. Written across and above her forehead, and in a style that suggests a thought bubble from a comic book, are the words: “Once it was human—­even as you and I! The Fly” (Figure 3.1). All the 91

92 · Warped Grids

Figure 3.1. Theatrical release poster for The Fly (1958).

words in this thought bubble are printed with a font that highlights Helene’s terror, while the words “The Fly” are even further emphasized by being printed in red ink. Standing on the window screen, almost on top of the woman’s head, is the source of the woman’s terror: a bipedal creature, wearing a lab coat and possessing the head and limb of a fly. The designer’s use of the gridded window screen in the advertisement is

Warped Grids · 93

genius. Screens are often used to control populations of organisms. We use window screens to keep flies out of houses, and we use fly swatters made from screens to kill flies once we find them where they don’t belong. As the mid-­twentieth-­century scholar of the dangers of flies, Luther S. West, writes: “Screening, as a general protection against mosquitoes, flies, and other pests, is so generally practiced in temperate zones . . . that we often fail to realize how generally lacking is such protection in many regions of the world.”4 The screen then is a wonderful example of the use of material grids as regulative mechanisms for controlling living things. They help confine organisms to specific territories, and they offer corrective elimination if that organism strays from its proper place. Flies aren’t the only organisms regulated by grids and screens. As Gilles Deleuze and Félix Guattari argued in A Thousand Plateaus, striated space is space that is overcoded with the demands of the state and belongs to the purview of royal science, the science of governing people.5 Grids introduce an element of hierarchical control with the demands of fitting singular elements into a larger pattern. A grid, like striated space, “intertwines fixed and variable elements, produces an order and succession of distinct forms.”6 This type of restriction can evoke fear when it is excessive. In very much the spirit of Deleuze and Guattari’s royal science, the fly is loose, a nomad perhaps, and it is the woman who is imprisoned behind the grid’s surface. Does the terror come from the vision of the monstrous fly/human chimera or from the realization of how attempts to control life end up controlling the controllers as well? This is one of the strange realizations of this new type of regulation built on aligning affects with the distribution of goods. Although these types of regulations are seemingly natural to some populations, as they are built on their attractions and repulsions, regulations cease to appear natural when they don’t fit well into the desires of the targeted population. As explored throughout this book, though, the relationship between grids and living things is more complicated than it first appears. Grids aren’t just used to partition spaces. As we read in our analysis of printed grids, grids can be most effectively utilized in consumer-­based economies when they provide the constitutive constraints for building things back up again. Comics, for instance, are an art form that exploits the use of grids to generate new types of associations. The same is true for the use of grids for regulating living things. First of all, despite the prevalence of the use of screens to contain organisms, not all organisms are effectively controlled

94 · Warped Grids

by screens. Humans have often thought of these less controllable types of lives as “pests” because they resist human efforts to regulate them. Controlling these hardy types of organisms is a never-­ending battle. This important lesson suggests that variance is much more robust than control. One cannot just tame the world and walk away. Second of all, many nonhuman organisms also use grids to collect food or manipulate their environment to their liking. Spiders weave webs and all cellular membranes use gridlike strategies to selectively filter components from their environment. And, finally, as we will touch upon in this chapter but more firmly develop in the chapter that follows, thinking of living things as composed of a grid has given scientists new insights into how organisms are organized, how they are related, and how they change over time. When the regulatory power of grids meets the procreative power of bodies, constructivism turns into generativity. Some of this generativity is channeled into socially useful directions. Some of this generativity might even change, or warp, the grid. Given the ambiguity of the potential interactions of living things with something as basic as a grid, it is instructive to think about how recent debates on the role of regulation in lives has been limited. Although recent discussions on the prevalence of biopower have helped some in political theory to think in broader terms about the relationship of power to living things, it often does so with an impoverished idea on the role of regulation in biology.7 Even one of the most sensitive analysts of biopower, Michel Foucault, refers to only two specific biological properties in his investigation of how the governance of subjects is informed by the dynamics of living things. He uses the term “population” when thinking about how groups of people are governed, and he uses the term “homeostasis” when thinking about how biopower remains responsive to aleatory events. More recent interlocutors, such as Giorgio Agamben and Roberto Esposito, have focused mainly on the idea of population and have tended to treat biopower as purely selective. This truncated view of life may fit the dialectical model of political theory that both theorists favor, but it is very limited for thinking about the manifold ways that life politically manifests itself. To more fully understand the ways that power animates lives, we must first gain greater insight into how theories of power have informed what it means to be alive in the first place. Becoming a better philosopher of the relationship of life to power demands for us to become more engaged biologists, not to define political subjects more biologically, but to better appreciate the material complexity of all forms of subjectivity. Each bio-

Warped Grids · 95

logical concept supports a politics, provides a justification for oppression, and offers an opportunity to breed new types of living things. This is exactly why looking at how organisms develop in addition to how they evolve provides such an interesting perspective on how life regulates. Many aspects of biology are interesting beyond the concept of selection. For example, even in evolutionary theory, selectionist accounts of evolution need to be balanced with ideas on how variance is generated. Evolutionary and developmental biology recognizes, for instance, that evolving organisms are also developing organisms and developing organisms create variation. It is time that the complexity of thinking about how life is biosocially regulated reflects this greater repertoire of concepts of what organisms actually do. This chapter uses the advertising poster of the 1958 movie, The Fly, to enlarge observations on the importance of grids for regulating forms of life as biological and political subjects. Life exists as a form of pestilence to political thinking based solely on rules and negotiated rights in that life continually challenges such political thinking, breaks it, and even incorporates it in warped and monstrous forms. It is not that pests are disordered, it’s just that they embody a suggestion for a different type of order, one that uses grids in new ways to find new potentials. Understanding life as a form of political pestilence is politically important for understanding the potentials behind the constrained options open to all forms of lives trapped in the grid.

Regulating Consumers A quick look into the circumstances of the release of the 1958 film, The Fly, helps illuminate the deeply classed, gendered, and racialized dimensions of the image in the poster presented above.8 The post–World War II years saw an unprecedented expansion of consumption in the United States. New types of relationships between humans and objects helped drive this expansion. Communication and entertainment technologies especially helped reconfigure the spaces that many people inhabited. Some of these technologies were revitalized forms of already familiar technologies of mass culture, such as movies, and some of these technologies were more novel, such as the transistor radio and the television. Historian Lizabeth Cohen has called this time period the development of a “consumers’ republic,” which she describes as “a strategy that emerged after the Second World War for reconstructing the nation’s economy and

96 · Warped Grids

reaffirming its democratic values through promoting the expansion of mass consumption.”9 Central to this new republic was the partitioning of mass culture into segmented populations and spaces. In the suburbs, new zoning laws were introduced, neighborhood associations created, or restriction clauses strengthened, and the racist and classist dispositions of realtors and buyers played into identifying restricted populations for homeownership.10 The influence of social science research on marketing also helped to sharpen “narrowcasting,” or the creation of content and goods for specific target groups.11 All of these developments ensured the circulation of people and goods by segmenting populations and targeting them. The scientist’s wife’s terror, then, forces us to ask the important question: exactly whose life will be made better with the further expansion of this type of consumerism? These types of reforms go beyond the politics of forms that we saw with organicism in chapter 1. In these cases, individuals are no longer just excluded; they are ranked, grouped, placed, privileged, condemned, and exalted. By looking at how the movie The Fly portrays the horror of misregulation, we can begin to trace the fears of misregulating vexatious late twentieth-­century bodies.

Circulation and Desire In a consumer-­based economy, companies profit by ensuring that goods are delivered to where consumer demand is highest. This led to the increased importance of research in business to create products to fulfill new types of consumer demands as well as the development of new strategies for product marketing and distribution to ensure that the products sold once they were manufactured. These changes in the functioning of industry provide the preconditions for the horrors generated in The Fly. In the movie, the brothers André and François Delambre own a machine shop in Montreal. One brother, François, runs the business, the other brother, André, is a research scientist developing new products. In a sense, they’ve literally bred the dictates of mid-­century capital into their family structure as one sibling innovates while the other sibling capitalizes on the innovations. André is busy working in the basement laboratory of his suburban home as he attempts to create a teleporter that he calls an integrator/­ disintegrator. Working as a commercial research scientist hasn’t dulled

Warped Grids · 97

André’s keen humanitarian interests. Teleportation, André enthuses to his wife, will alleviate human suffering: I can transport matter, anything, at the speed of light perfectly. . . . Think what it will mean. Food, anything, even humans, will go through one of these devices. No need for cars or railways, or airplanes, even spaceships. We will just set up matter transmitter/ receiving stations throughout the world and later the universe. There will never be a need of famine. Surpluses can be sent instantaneously at almost no cost anywhere. Humanity need never want nor fear again.12 The hyperbole of André’s speech reveals his industrial-­sized ambitions. Selling large amounts of goods requires the ability to transport these goods to where they are needed most. The integrator/disintegrator promises to do just that, and with little overhead costs. Although posed in the futuristic terms of atomic power in the movie, the question of distribution is as old as capital itself: how do we ensure that the right amount of product is delivered where it can accrue the greatest amount of profit? André’s plan to end human suffering, then, is to offer a form of instantaneous regulation where desire can be immediately satisfied through distribution. From this perspective, the goals of the scientist in the movie are aligned with the goals of the advertisers who designed the poster for the movie. They are both focused on how to align distribution with desire to earn profits. In one sense the movie poster has an even more difficult job than the integrator/disintegrator. The transporter must move goods to those who already desire them; the poster, on the other hand, must create desire in a potential audience unfamiliar with the final product. This problem was especially acute for movies, where advertising helped determine the early returns for the film. As Olen Earnest argues, “The film market is characterized by quick entry and quick exit and the majority of the box office gross must be made early—­for there may well be no later.”13 Ads for movies, then, were created in order to stimulate audience demand at the times and places that the movies opened. In some ways, this could be considered an even greater trick than delivering goods to where they were already desired. Yet we had already created numerous mechanisms for harnessing the role of desire in the regulation of goods, people, services, and things.

98 · Warped Grids

The lesson once again comes from investigating the role of grids in society. Instead of looking at the role of grids as a graphic design device, as we did in chapter 2, we will focus more on the role of grids in triaging goods and bodies in a consumer society.

The Complex Ontology of Grids: The Constructivist Moment of Discipline In a famous passage from Discipline and Punish, Michel Foucault describes the gridded design of the printing workspace at the Oberkampf manufactory at Jouy. The use of a grid better allowed the shop foreman to discipline the activities of each worker by coordinating the distribution of their bodies. By walking up and down the central aisle of the workshop, it was possible to carry out a supervision that was both general and individual: to observe the worker’s presence and application, and the quality of his work; to compare workers with one another, to classify them according to skill and speed; to follow the successive stages of the production process. All these serializations formed a permanent grid: confusion was eliminated; that is to say, production was divided up and the labour process was articulated, on the one hand, according to its stages or elementary operations, and on the other hand, according to the individuals, the particular bodies, that carried it out. . . . At the emergence of large-­scale industry, one finds, beneath the division of the production process, the individual­izing fragmentation of labour power; the distributions of the disciplinary space often assured both.14 The supervisor’s ability to freely move between the stations in the grid allowed him to appraise the speed and quality of each worker’s labor as well as evaluate how each worker’s labor fits into the flow of work overall. This depiction of discipline involved two especially important moments: a moment of breaking down, what Foucault called “partitioning,” and a moment of reassembly or “usefulness.” Foucault indicated that the structure of the grid is important for social control as it was used to promote both moments. In partitioning, a worker’s tasks are broken down, or atomized, to reduce complexity. Grids allow one to do this easily as they separate one partitioned space from another. Things that happen in that

Warped Grids · 99

space can be thoroughly examined and changed if necessary. In usefulness, the supervisor reassembles the previously atomized tasks to construct something different from before, something more efficient and better organized. Grids are useful, then, in that they package together an analytic moment of partitioning with a constructivist moment of usefulness. This is an especially important insight as it suggests how grids not only help fragment motions; they help fabricate them as well. Most criticisms of industrialization overemphasize the act of partitioning at the expense of coordination. When we think of how control mechanisms in industry affect us, for instance, we tend to concentrate on how all the potential actions of a body are reduced to simple and repetitive gestures. Charlie Chaplin’s depiction of the factory assembly line in Modern Times is a good example of this critique of control. In this scene, Charlie is forced to reduce the complex behaviors of his body to a limited set of functions useful for industrialization, such as the tightening of one bolt after another. The power of this critique comes from the implicit assumption that the proper coordination of body parts should occur at the level of the worker’s body as opposed to the level of the assembly line. Consequently, the industrialization of bodies is often criticized as causing a physical and social violence to the integrity of the body, and the use of grids considered to play an important part in this violence. Working in grids threatens the implicit control that organicism appears to offer. It splays mind from muscle, and disarticulates the fluidity of bodily movement because it displaces the regulative functions of the whole organism to the regulative functions of the assembly line. The violence to the workers’ bodies on the shop floor, then, came not just from controlling bodies, as all bodies need some form of control, but from conceptually breaking down each worker’s body to reengineer it with a purpose different from its preindustrial function, or as Foucault might say, to give the body a usefulness in disciplinary society.15 It is important to remember, though, that ideas about managing bodies are not transhistorical, and the conception of how bodies fit into society has changed over time. This is not just a reminder for the importance of historical study; it is a caution to avoid assuming that last year’s form of control is inherently better or more natural than this year’s. Consequently, it is instructive to think about how this form of bodily holism has changed in many areas of health and medicine, where our exercise increasingly depends upon breaking down complex movements into simple gestures, allowing one to build mass in specific muscle groups. The coordination of

100 · Warped Grids

these efforts happens at the level of the physician or athletic trainer instead of the shop floor manager, however. And, just as importantly, each conception of the body carries its own political implications. The violence to the industrialized body comes from the fact that it severs an implicit assumption about bodily control. Yet, as we also saw in chapter 1, an overemphasis on the integrative functions of an organic body could also work against inclusive pluralism by promoting an overly unified vision of life. Using grids as a mechanism for controlling life didn’t introduce the idea of control through regulation; it did, however, change how goods, bodies, and information flowed in the late twentieth century. Effective social operations meant controlling not just for what was happening, but also for what potentially could happen. This added a new layer to regulation that sought to find ways to regulate not just what one was, but what one would eventually become. The grid, in a sense, had to find a way to be more responsive.

Homeostasis and Security In his final lectures, Michel Foucault labors to demonstrate how the confluence of desire, regulation, and the circulation of goods led to a new form of power directed toward regulating the health of a society. This power differs from sovereign power in that it focuses on the power to “make live and let die” instead of the power “over life and death,” and it differs from disciplinary power in that the objects of power are populations instead of the bodies of individuals.16 Foucault adapts a concept from political economy, “biopower,” and he argues that its goal was to secure the safety of a population through the control of aleatory events: And we also have a second technology which is centered not upon the body but upon life: a technology which brings together the mass effects characteristic of a population, which tries to control the series of random events that can occur in a living mass, a technology which tries to predict the probability of those events (by modifying it, if necessary), or at least to compensate for their effects. This is a technology which aims to establish a sort of homeo­ stasis, not by training individuals, but by achieving an overall equilibrium that protects the security of the whole from internal dangers.17

Warped Grids · 101

The regulation of the vagaries of biological processes aims to turn groups of individuals into secure populations. This would be accomplished through the study of probabilities with the intent to create a dynamic equilibrium where chance events could inform rather than threaten a population. In the first three lectures of his series entitled Security, Territory, and Population, Foucault details how regulation works to maintain the security of a population. A close reading of these lectures helps to explicate what in chapter 2 I call the regulative properties of grids, and it places these properties into their political economic context. Tellingly, in an analysis of the distribution of grain in Europe in the eighteenth century, Foucault begins where the fictional scientist André Delambre began, on the importance of eliminating the scarcity of goods through the alignment of desires and resources. Since many events can lead to the fluctuation in the distribution of grain (weather, marketing, demand, and so on) and these can’t be scheduled, securing the circulation of grain involves “letting things happen”18 and being able to respond to events once they occur. This means that the goal of regulating grain is not to prohibit the occurrence of events (like laws do) or to normalize individual behaviors through prescription (like disciplines do), but to channel what could occur to meet a productive end: “[The essential function of security] is to respond to a reality in such a way that this response cancels out the reality to which it responds—­ nullifies it, or limits, checks, or regulates it. I think this regulation within the element of reality is fundamental to the apparatus of security.”19 Since this form of regulation is aimed at promoting the circulation of goods, what emerges is a laissez-­faire conception of freedom where being free means being open to “the possibility of movement, change of place, and processes of circulation of both people and things,” as opposed to an older conception of freedom based on the “exemptions and privileges attached to a person.”20 In this conception, the ability to circulate goods and people not only supplies demands; it exemplifies what it means to live in a free society. And significantly for the movie, a teleportation device is perhaps the ultimate technology for responding to the reality of aleatory events that threaten distribution. A teleportation device can immediately address any perceived shortages. Taken in its political economic context, André’s ­integrator/disintegrator is much more than a humanistic dream to eradicate all wants; it becomes the ultimate technological manifestation of the neoliberal dream of a regulation so efficient that it seems to disappear.

102 · Warped Grids

According to Foucault, desire was the “mainspring” of this neoliberal form of regulation, as it brings individuals with different interests into alignment to form populations.21 These populations are then managed based on what appears to be the manifest naturalness of their desires. For it is “the production of the collective interest through the play of desire . . . that distinguishes both the naturalness of the population and the possible artificiality of the means one adopts to manage it.”22 Although “desire” might be too narrow of a register to talk about the role of affect in political economy,23 the value of Foucault’s analysis is that it demonstrates how mechanisms of regulation linked the affects of a consumer economy with the emergence of population thinking.24 And although Foucault never directly talks about the development of advertising as one of these mechanisms, those of us living in the twenty-­first century do not have the luxury of ignoring it. As we saw in the last chapter, early mass market advertising helped give form to our entertainments and even informed how we looked at and talked about the world.

The Flexible Spaces of Security As Foucault recognized in his lectures on Security, Territory, and Population, the shift in emphasis from grids as fixed to grids as responsive moves one away from the use of a grid as a disciplinary system, as we saw on the shop floor, to the use of the grid as a security system that allows for the greater circulation of goods through a dynamic regulation of production practices.25 Foucault suggested that ensuring the circulation of goods in eighteenth- and nineteenth-century France went beyond restricting what could happen in a specific place to attempting to control for anything that might happen, such as chance environmental events like drought that could cause scarcity. The result was a responsive system of regulation that flexibly interacted with its environment. The emphasis then shifted from policing a “place,” in the way that some urban grids are used, to creating regulations that allow for grids to come into being in certain ways.26 Grids became responsive and thus allowed for the construction of new outcomes by limiting the potential of all types of relationships. It is not that the associative power of grids placed them beyond control; rather, that meaning and control reinforced each other to restrict the types of associations that could occur. A group of Japanese and Japanese-­influenced architects have argued

Warped Grids · 103

that the West draws a much stronger philosophical boundary between artifice and nature than the Japanese do.27 According to architect Kisho Kuro­kawa, “In Japanese architecture . . . traditional spatial elements of design such as ceilings, alcoves, and walls are each autonomous, that is, they are on the independent planes of a two dimensional world.”28 Composition with these elements then means placing them into different relationships that allow specific forms to emerge. One of the most obvious architectural examples of this is the use of light and mobile screens in Japanese architecture to variously break up spaces at different times. The mobility of the architectural elements forms a potential differential that allows for many ways to organize space. The fluidity of grids allows for this multiplicity of becomings created through specific juxtapositions. The exchange between indoor and outdoor spaces is modulated through a sequence of thin screens, such as paper sliding doors, wooden screens, and bamboo folding curtains that present a certain degree of transparency and permeability and are organized through spatial patterns. All of these devices create juxtapositions of heterogeneous elements instead of uniformity.29 So even though these architects were working with three-­dimensional spaces, the walls of the grids were permeable and adjustable enough to create a sense of the potential for multiple different outcomes in arranging the spaces. Flexible grids, then, can create a multiplicity of outcomes through their ability to juxtapose heterogeneous elements. As Salvator-­ John Liotta writes, both Japanese and Western culture “know the significance of becoming. Yet while Western aesthetics seeks concluded and perfect form and patterns, Japanese aesthetics also value transiency and imperfection.”30 And although I am suspicious of the hard and fast distinction between Japanese and Western aesthetics, Liotta’s point is useful in that it suggests how different spatial practices bring about different ways of considering how space can be organized. Central to this understanding is recognizing how ideas of control and regulation also began to be used in more fluid and ephemeral ways. As we will see in chapter 5, grids used this fluidity by layering grids on top of grids; this allowed them to be more responsive to a greater number of problems. For now, we will focus on the role of anomalies in grids, or what I term the “pests” that show the limits or inconsistencies of a specific gridded logic.

104 · Warped Grids

Pests in the Grid Regulation requires much work and not all regulation runs smoothly. Some organisms, for instance, are very difficult for humans to regulate and are relegated to the category of “pests,” In the movie The Fly (1958), a single housefly sets into motion the horrifying consequences of misregulation of fly and human bodies. There is a reason why a fly was chosen for the role of villain; human and fly interactions are difficult to regulate. From the perspective of many humans, flies are master minglers and they pose a special threat to those who want to keep things separate. All flies belong to the order Diptera along with other two-­winged insects like gnats and mosquitoes. Over 90 percent of the flies encountered in human habitations, however, are the common housefly, Musca domestica. The housefly is an especially persistent pest as it lives near humans and can be found in most places around the globe. As Steven Connor acknowledges in his book on the history of attitudes toward the fly, “More than any other creature, the fly has a reputation for hedonism. . . . The fly takes its pleasure promiscuously, restlessly, unswervably, unashamedly.”31 The wanton vitality of the fly is a constant reminder of the difficulty of regulating living things. It is this peskiness that creates the enabling conditions for the disaster that follows. The movie begins with a failure of regulation as the fly is seen entering through a hole in a window screen as the opening credits roll (see Figure 3.2). In doing so, the fly breaches the security precautions meant to keep flies and human spaces separate. Most types of flies, in fact, are masters of mixing things that humans try very hard to keep separate. The housefly’s smaller cousin, the fruit fly, Drosophila melanogaster, for instance, lives off the surplus of human agriculture. Its primary ecosystem, then, is created by the inability to efficiently circulate goods, people, and things. But the fly doesn’t stop with mixing things up; once mixed, the conditions for new types of outcomes are generated. For instance, the fruit fly is a well-­ known symbol for the fecundity of life. In some cases, humans have been able to harness this fecundity as a type of knowledge. As Robert Kohler has recently noted, the short doubling time and the numerous offspring of Drosophila melanogaster make it very useful for genetic studies.32 Consequently its ability to comingle and its generativity has made the fruit fly one of the most studied of all living organisms. What one quickly realizes when studying flies, though, is that it is much too human centric to think of everything in the grid as regulated and everything outside of the grid as

Warped Grids · 105

Figure 3.2. The opening credits from The Fly (1958). The fly enters through a hole in the window screen creating the enabling conditions for the teleportation disaster. The Fly (1958), DVD, 20th Century Fox, 2013.

unregulated. Multiple forms of regulation seem to be at stake, where each form of regulation not only privileges (selects for) specific types of lives but also generates new ways of living. A brief look at how flies order the world is instructive in understanding this point.33 Flies aren’t just willfully messing up human attempts at regulation; key differences between the physiology of flies and humans encourage the mixing of different types of regulative spaces. As the 1958 Encyclopedia Britannica Film The House Fly warns, “[A fly] cannot bite or sting but its physical structure and feeding habits make it a carrier of disease and death.”34 For instance, the fly has hairs on its feet that are sensitive to taste. This means that a fly can taste its food by walking on it. Humans, however, have their organs of taste more closely confined to their alimentary canal, allowing for the separation of the functions of eating and walking. This is a profound set of differences in the two organisms as it implies that humans try very hard to keep production separate from consumption, while flies need to mix these up to live. Is it any wonder that an organism that deftly mixes consumption and production would come to represent a special threat to humans during a period in the U.S. political economy where consumption began informing production in new ways? And it’s not just eating that is the problem. Flies also mix up human attempts to separate different types of excretions. When a fly larva hatches, it needs a carbon-rich environment to begin feeding. Consequently, flies like to breed in the organic things that humans throw away: trash, carrion,

106 · Warped Grids

and feces. A 1942 study from Georgia, U.S.S.R., even attempted to identify which type of manure attracted the most flies: “[G]ravid females deposited more egg batches on sheep dung, provided it was fresh, than any other type of manure available. After that, pig, horse, cow, and buffalo manure were chosen in the order named.”35 The reader is left to ponder where the nurturing properties of human feces might rank on this fly-­centered view of mammalian excrement. The fly’s physiological need to mix spaces of excretion and procreation has sharpened the fly’s reputation as a vector for disease. Writing seven years before the release of the movie, the medical entomologist Luther West claims that the house fly above all other types of flies is a problem for public safety: “In summary, then, it must be said that although the superfamily Muscoidea includes a number of human benefactors, it also contains many black sheep. Among these, Musca domestica is certainly the blackest.”36 The answer for West as for many others during the mid-­ century war on flies was greater education about the role of flies in public health and adoption of proper hygienic practice. “If it is possible to surround the rising generation with an environment relatively free from flies and fly-­borne filth, communities of the future will perhaps insist upon surroundings which are both pleasing and sanitary, and which they will not willingly relinquish, even under stress.”37 Some epidemiologists supposed that America’s hygienic future depended upon securing territory from the fly through regulation and then never ceding it.

Understanding Development by Regulating Affects Although Foucault emphasized the regulation of goods and people, or things that can circulate in a territorial space, a fly’s threat to regulation goes beyond macrospatial concerns of territory, security, and populations. In fact, one could argue that the horror of the film, The Fly, comes from an understanding of how a simple act of misregulation can replicate itself by generating new types of living things. This offers a deepened insight into the role of bodies and regulation: creating order not only makes the distribution of goods easier, it keeps bodies and societies from generating new types of pests. This is obviously true for the human and fly hybrids that the integrator/disintegrator misassembles, but it is true for all of us who live our lives in the grid. If all forms of life emanate from a limited set of biological properties, how exactly does life regulate development so

Warped Grids · 107

that the same properties can produce organisms as seemingly different as flies and men? Either these instructions must be coded for or they must be regulated differently during development. In the first instance regulation is relegated to turning genes on or off. In the second instance, regulation itself becomes the generative property. The fear of the 1958 movie The Fly is most assuredly the fear of regulation as a generative property. To understand this fear, it is necessary to see how the biological sciences explore an appreciation of the strangeness and familiarity of organisms. One of the reasons that fruit flies are so well studied is that scientists assume that all forms of life share similar biological processes. Studying a fruit fly allows one to gain knowledge of living things, like humans, without the ethical or methodological constraints of experimenting on humans. Yet, even those who have spent decades researching the little creature often find their appreciation for the similarity of basic process all mixed up with an otherworldly attraction to the strangeness of its form. This inability to regulate affects toward the fly turns what could have been a very routine article on genetic mosaics by geneticist Curt Stern into one of the strangest meditations on the relationship between fly and human ever published. Part love story, part naturalist’s fable, and part scientific argument, Stern’s paper demonstrates how attempts to regulate human feelings toward flies create the enabling conditions for understanding the role of regulation in all forms of life. Little about the dry title of the 1954 article, “Two or Three Bristles,” prepares us for the intimacy of its opening paragraphs. Stern begins where a lover might, by gazing into the face of the object of his appreciation and describing what he sees: “the head with giant red eyes, the antennae, and elaborate mouth parts.”38 Stern then lets his gaze fall down the body of the fly, balancing a tenderness in the detailed observations with the use of adjectives chosen to evoke an alien beauty: “At the arch of the sturdy thorax bearing a pair of beautifully iridescent, transparent wings and three pairs of ­­­legs. . . . A shining, waxed armor of chitin entirely covers the body,” and through this armor “[w]ith surrealistic clarity the dark colored bristles and hairs project from the light brownish surface of the animal, delicate but stiff, in rigid symmetry.”39 Even under close scrutiny, fruit flies mix things that many humans assume are separate. They mix surrealism with order and delicacy with fortification. Stern gently turns introspective by recognizing that his fascination for this small creature’s form is motivated by how different it appears from

108 · Warped Grids

his own: “The fascination of this view lies partly in its difference from the aspect which so many other animals, including man, exhibit. There is obvious life in the glow of a young skin, the moisture of lips and eyes, the relaxation and tension of muscles in rest and activity, and in the expression of emotions.” There “are no such dynamics in the insect, hardly any visible changes are evident even when old age approaches. No loosened skin nor weakened muscles disturb the exactly sculptured, eternal form.”40 These visible physical differences between humans and flies mark very different ways of inhabiting the world. Humans open ourselves to the world, lending us responsive and vulnerable to change. The fly, on the other hand, withdraws from the world, even armors itself against it, which makes it seem more ideal, mechanical even, and less responsive to its environment. For Stern, the key to these different ways of existing in the world is the organism’s relationship to moisture. The human’s large size guarantees that the loss of moisture won’t be fatal, so a drop of moisture on the lips of a human suggests a sensual vitality, a luxurious excess, and not an imminent loss of life. “All small insects like Drosophila have a large body surface in relation to their volume. Thus they must be protected from excessive evaporation or they will soon become dehydrated.”41 Even the hard chitin covering of the fly, as lovingly described by Stern, helps the small creature remain a victor in its constant battle against the loss of moisture. Although humans and flies are both animals and they both inhabit the same world, they do so by exploiting different physiological strategies. Tellingly, Stern sees a responsiveness to the environment as a human quality. Yet, all forms of organisms need to be responsive on an evolutionary scale, they need to be able to adapt to circumstances. In this case, it is the organism’s response to moisture that provides the physiological difference that makes the sensual difference in how the organisms are formed. Stern makes the point for regulation by changing the focus of his description away from the physiological differences between the two organisms toward the basic biological similarities. Despite the differences in forms, the two organisms still share similar genetic and cellular processes, and this is what makes flies such good model organisms for studying the genetics of humans. Yet, the insect is our brother in life. However widely separated we have become during the ages of evolution all are built of protoplasm, the basic life stuff. The vital functions of cells and organs

Warped Grids · 109

in insect and man are alike or very similar. The laws which govern the way in which the insects inherit the parental genes are identical with human heredity. The problems which the egg of an insect or of a mammal have to solve as it develops are much the same.42 Even something that appears as otherworldly in form as the fly could possess similarities to humans when it comes to biological processes. So although the title of Stern’s article seems to point only to a miniscule developmental anomaly in one of the smaller household pests, the number of bristles expressed on a body segment, this anomaly touches on the very big problem of how the diversity of living forms can come from similar processes. How can flies and humans be made of similar stuff and yet appear so different? The answer, of course, is in how that stuff is regulated. Regulation emerges as more than a set of rules but as a means of generating difference from similarity. Here again, a closer look at the script for the 1958 version of The Fly in terms of the generative logic of grids is especially revealing.

Understanding Bodies by Regulating Development The horrors of regulating flies made them an appropriate star for a monster movie, but the inability to keep fly and human spaces separate was only the prelude for a much greater threat: keeping flies and human body parts separate. It is here, once again, that we see grids work their regulatory magic. As we saw earlier in the chapter, grids don’t just control things—­ they create things as well. When the generative power of living things combines with the constructivist power of grids, small lapses of regulation can threaten to create monstrous consequences. This is what happens, of course, to the fly and the scientist while in the integrator/disintegrator in The Fly. The integrator/disintegrator works by splitting things up into constituent particles, making them easier to move, and then reintegrating them once they reach their destination. “For a split second,” explains André the scientist to his wife, “an infinitesimal part of a second, this [ashtray] disintegrated” as he demonstrates how the machine worked on a manufactured ashtray. “For one little moment, it no longer existed. Only atoms traveling through space at the speed of light. Then here, a moment later, integrated again into the shape of an ashtray.”43 When looking for a convenient example of how this integration might work, André appeals to the screen technology of the television, a much

110 · Warped Grids

more mundane regulator of Cold War desires, to demonstrate how complexity can be built from a series of small particles. “Take television, what happens? A stream of electrons, sound, and picture impulses are transmitted through wires and the air. The TV camera is the disintegrator, your set unscrambles or integrates the electrons back into pictures and sound.” Although fanciful when applied to objects like ashtrays, building things up from particles was a sleight of hand that chemical and electronic imaging technologies had been performing on surfaces for a very long time.44 But something goes wrong in reassembling the organisms. The ­integrator/disintegrator doesn’t recognize the complex regulation needed for two organisms instead of one. The result in the 1958 movie was a direct exchange, where the fly head and leg are swapped for a human head and arm. What results are two new types of organisms: the human-sized human/fly and the fly-sized fly/human (see Figure 3.3 for a picture of the human/fly and Figure 3.7 for a picture of the fly/human). There is something very industrial looking about this body-­part switch. It’s almost as if these two bodies were made from exchangeable parts and these two parts got switched at the factory. In this case, the misregulation would have come from middle management, where the wrong parts were supplied to the assembly line. This image of an industrial mix-­up is an especially apt way to describe how makeup artist Ben Nye Sr. manufactured the fly head from a hodgepodge of premade parts. For instance, he created the fly’s eyes by gluing plastic shells on a latex mask and painting them with a coat of luminescent paint; he made the fly’s proboscis with wood, clay, and latex and then attached it to the mask; he created the fly’s antennae from sculpted turkey feathers; and he had to special order a wig from Max Factor to simulate the hirsute skin on the back of the fly’s head. The actor, Al Hedison, had a plaster cast made of his head so that the mask would fit (a routine special effect for monster movies at the time), and a small wooden core allowed Hedison to move the proboscis with his mouth.45 Although the intention of the costume was to make Hedison look as if he was part fly, it achieved this effect by mixing consumer items in strange ways. One of the problems with movie special effects is that they often don’t age well. As scientific knowledge changes, what is considered a credible type of fantasy changes as well. It is not that these transformations need to be conceived of as scientifically accurate to be convincing. Still a basic knowledge of scientific processes informs how certain types of fictional

Warped Grids · 111

Figure 3.3. The human-sized human/fly from The Fly. Publicity shot distributed by 20th Century Fox.

transformations might take place. Even our entertainments must delicately regulate the relationship between science and affect. Filmmaker David Cronenberg, for instance, wanted an entirely new way of imagining what a fly/human hybrid might look like for his 1986 remake of the movie: [O]ne of the things that the producers and I spoke about and wanted to avoid was having Brundle [the scientist] turn into a 185-­pound fly. It would be silly if Brundle was just turning into this huge fly, and physiologically impossible, even given the fantasy elements of any sci-­fi horror film. It would have been as silly as the head switch in the original, and it wouldn´t even work as well as that since this isn´t the ´50s anymore. I wanted to make sure, as Brundle says, that he was evolving into something that had never existed before, a real fusion between an insect and a man that would embody elements of both.46 Cronenberg’s vision of the human/fly hybrid reflects a belief in the importance of DNA as a mechanism to code new proteins. The idea was to imagine what would happen if fly DNA was mixed with human DNA coding for proteins that were thought to be chimeric between fly and human proteins. The result would be a human becoming a fly at the molecular

112 · Warped Grids

level (Figure  3.4). Cronenberg’s conception of genetics reflects popular thinking at the time: genes were thought to primarily provide the instructions to create organisms. Cronenberg’s human/fly hybrid (there was no fly/human) was a result of mixing up these instructions creating a combination between fly and human phenotypic traits. On some occasions, however, earlier versions of special effects tell us as much about how scientific reality informs our dreams as the special effects in the later remake. For instance, although Cronenberg’s conception of how fly and human DNA might mix fits a 1980s view of how genes code for proteins, it tends to assume that the regulation for these genes during expression is also simply coded for, or part of the DNA. Thirteen years after the production of the film, a large consortium of scientists published a “rough draft” of the Human Genome Project that helped shape new ways of thinking about how genes inform development.47 One of the paper’s most stunning observations is that there were far fewer genes in the human genome than previously supposed. They thought they might find over one hundred thousand genes; they found around twenty-­six thousand.48 There weren’t enough genes to code for substances and then give the detailed instructions needed to turn these substances into forms and structures. “The modest number of human genes,” speculates Venter, and others, “means that we must look elsewhere for the mechanisms that generate the complexities

Figure 3.4. The fly/human hybrid at a middle stage of transition from David Cronenberg’s 1986 remake. The Fly, DVD, 20th Century Fox, 2007.

Warped Grids · 113

inherent in human development.” Mixing the genes from two different organisms wasn’t comparable to mixing the instructions for assembly of the organisms. What was needed was a better sense of how this relatively small number of genes was expressed. “The next steps are clear: We must define the complexity that ensues when this relatively modest set of about 30,000 genes is expressed.”49 Development, it seems, depended on how genes were regulated as much as on what they coded for. Furthermore, as other organisms’ genomes were sequenced, it became obvious that there weren’t enough differences between two organisms’ genomes to support a view of genes as coding for proteins as well as providing instructions for assembling these proteins into different organisms. For instance, recent genomic evidence has suggested that the sequence data between humans and flies are not that different; according to one source, humans and fruit flies share 60 percent of their genomic sequence and many of the conserved sequences are implicated in regulation and development.50 It is becoming apparent that genes not only code for substances; they do so in a way that allows for similar substances to be expressed at certain places and at certain times. As anyone using a grid to design something will tell you, the overall pattern of the design is not only dependent on what you put inside the grid but on how you arrange the relationships within the grid. Intriguingly, some variations in regulatory changes look eerily reminiscent of the “head switch” of the crude special effects in the 1958 version of the film. Several mutations have been identified that regulate where something is produced. In the fruit fly, for instance, a mutation in the Antenna­ pedia homeodomain protein can convert an antenna to legs (see Figure 3.5). This isn’t a difference in substance since both are made from the same substances, and it isn’t a difference in instructions coded in genes, since the genes that code for the development of appendages are remarkably conserved across species. Rather, it is a difference in where parts are assembled—­how the organism is regulated during development. The resulting effect is as if the antenna parts of the fly were accidentally swapped with the leg parts during assembly. Perhaps the 1958 human/fly costume was closer to reality than Cronenberg had supposed. These types of mutations are called homeotic mutants, a term derived from William Bateson’s 1894 book, Materials for the Study of Variation.51 Bateson suggests that homeotic variations were different from other types of variations in that they changed one thing to something else altogether,

114 · Warped Grids

Figure 3.5. Scanning electron micrographs of the head of a Drosophila melanogaster homeotic antennapedia mutant. This mutation shows the replacement of the fly’s antenna with metathoracic legs (legs usually found on the second thoracic segment). Image courtesy of F. Rudolf Turner.

“for the essential phenomenon is not that there has merely been a change, but that something has been changed into the likeness of something else.”52 These wholesale changes in the parts of organisms clearly looked different from the small changes in characters that became known as mutations through the early work in fruit fly genetics. As such, they have proved to be lightning rods for debates about the role of discontinuous change in evolution.53 Although this wholesale switching of parts doesn’t appear to be a major driver in the development of speciation, it has been extremely important for considering how the arrangement of parts is related to the overall form of an organism. As we will see in chapter 4 when we study another homeotic variation of the fruit fly, the bithorax complex, thinking of organisms as modules of a grid would come to dominate ideas on development in the late twentieth and early twenty-­first centuries. This, in turn, would further move conceptions of the regulation of organisms away from the teleological relationship between parts and wholes that we

Warped Grids · 115

saw structuring debates in German morphology in chapter 1. How does this understanding of the development of organisms change how we think about the political regulation of bodies in late twentieth-­century societies? Addressing this question requires returning to see how Foucault’s use of the terms “biopower” and “biopolitics” have inflected recent debates on the relationship of politics to life.

Understanding Biological Concepts by Regulating Bodies Clapperton Mavhunga has insightfully written on how colonialism in Nigeria depended on an especially exclusionary type of white supremacy that treated humans and nonhumans as pests to be exterminated. Flies and black Africans were pests that white settlers needed to immunize themselves against. As Mavhunga observes, “This slippage between the hugandanga (terrorism) of insects and that of humans was quite pronounced in the white Rhodesian imagination, with areas of tsetse fly and terrorist activity alike declared ‘red zones.’ Hence medical doctors immunized children against measles, entomologists chased fly and trypanosomes, the vets inoculated livestock, and the Rhodesian army vaccinated the countryside against ‘red-­black terrorism’ (communism and black insurgents).”54 Pesticides, vaccines, antiterrorist campaigns, and as we’ve seen in this chapter, screens, were all tools used to regulate pests and keep them in their place. Although Mavhunga’s reading of the eradication of pests in colonial Africa is not explicitly biopolitical, it bears a structural resemblance to a brand of biopolitics formulated most comprehensively by Roberto Esposito, where the development of modern communities occurs through an immunization toward the dynamics of exchange that gave birth to the community in the first place.55 In Mavhunga’s case, immunity was conceived as embodied, through the development of vaccines, as well as spatial, in the removal of populations deemed a threat to white rule. And although “immunity” is key to Esposito’s conception of biopower, Esposito’s conception of immunity, or immunitas, is inherently legalistic in that it means “granting an exception” instead of the more biologically based “protecting from.” According to Esposito, communities are formed through the circulation of gifts. Modern communities, however, are slightly different in that they also grant individuals an immunity, or dispensation, from the circulation of gifts that form communities. This occurs as immunitas and is the negative for community formation, or communitas. Consequently,

116 · Warped Grids

what is immunized is not a coherent conception of an individual but a dispensation from the obligations of communitas. This dispensation allows for the emergence of “individualistic modes” of organization within the community as it provides a protection from the excesses of gift giving.56 A modern conception of individual rights emerges when this dispensation is brought into the forms of community organization through legalistic and political proceedings. “What is immunized,” Esposito explains, “is the same community in a form that both preserves and negates it, or better, preserves it through the negation of its original horizon of sense.”57 It preserves it in the sense that there remains a form of communitas, yet the community has also been negated by adding individualistic modes of political engagement. There is much to like in Esposito’s version of biopolitics. One appealing trait is that there is no hard and fast distinction between the political and bare life as proposed by Giorgio Agamben. Agamben appeals to Roman law to suggest that there are two types of lives—­lives that have political worth, called bios, and lives that are excepted or excluded from political representation, called zoë. He defines the two lives in his Introduction to Homo Sacer, “zoë, which expressed the simple fact of living common to all living beings (animals, men, or gods), and bios, which indicated the form or way of living proper to an individual or a group.”58 Thus in Agamben’s conception of biopolitics, politics is something that is stripped from subjects in a process that has been occurring since the politics of kings and sovereigns, “and that the inclusion of bare life in the political realm constitutes the original—­if concealed—nucleus of sovereign power.”59 The political dimension of bios is what lifts humans from bare life into humanity. As Agamben argues, “[i]n the ‘politicization’ of bare life—the metaphysical task par excellence—the humanity of living man is decided.”60 Since Esposito’s understanding of biopower begins by recognizing the politically charged role of exchange in all social formations, immunitas, then, is a layering of a negative political dialectic over the initial acts of exchanges.61 Consequently, for Esposito, there are not two types of lives, political and bare life, there are two types of politics, a politics of exchanges and a politics of dispensation from exchange. The other interesting aspect to Esposito’s notion of immunitas, is that it isn’t strictly a protective form of immunity, like what one might imagine in the case of embodied, biological immunity. Although immunity does protect social organization, it does so by injecting a complex notion of individual dynamics into the role of

Warped Grids · 117

organization and its functions. This allows for a sophisticated way to envision how biopolitics operates as a form of exclusion even when it is acting as a means of individual differentiation. Esposito’s type of model is very interesting for thinking in terms of how a form of dialectical exclusion can occur in biopolitical formations. The model works well for thinking about oppositional forms of biopower, such as those that Mavhunga located, where the goal was to protect white colonial populations from some types of exchanges through the advance of western colonial liberalism and biomedical immunity. Despite the power of Esposito’s negative dialectical model, however, I am convinced that neither Agamben nor Esposito’s conception of biopower is a sufficient way to think about the multitude of interesting ways that biology operates in tandem with political oppression in contemporary society. One point of emphasis that gets lost with Agamben and Esposito’s overriding focus on the relationship of politics to power is the interesting relationship between biopower and changes in economic practice.62 As many historians of science have noted, economic and biological theories often shadow and support each other.63 This does not mean that living things passively reflect how they are theorized. In fact, the difficulty of theorizing about life is what makes it so interesting. The study of life is complex. It can’t be reduced to simple physical laws or a philosophy modeled on a single mathematical assumption. The very complexity of biology guarantees that living things utilize multiple strategies to keep living. One reason is that living things exist between chemical and social scales. Consequently, organisms both resemble and necessarily complexify attempts to describe them through either purely social or chemical terms. Thus, there can be no unified theory for life, let alone a unified theory of the relationship of politics to life, as life itself emerges through the interactions of very real differences across scales.64 This is one of the reasons why the concept of “biopower” is simultaneously analytically fruitful and intellectually fraught. There is value in thinking about how life is inherently political. This value is not reducible, however, to an easy formula for how life and politics interact. Consequently, it is not so much that we can’t identify what life is, per se, but that our identification will be inherently limited. This does not mean that we need to stop theorizing about the relationship of politics to living things. We should be careful, however, that our attempts to think in terms of life and politics create possibilities for thinking about multiple types of lives. If done with care, and

118 · Warped Grids

contrary to a politics of either political or biological determination, using biology to think about power can and should open our eyes to see how power informs multiple forms of lives across the complex interaction of different scales of existence (molecules, cells, bodies, and populations). Therefore, stressing the diversity of life forms should be an important aspect of biopower theories. It is useful to adapt biology historian Georges Canguilhem’s important insights on biological concepts: the history of biology can be viewed as a history of “conceptual progress.”65 A closer look at what Canguilhem means by “concepts” promises to bring analytical clarity to discussions of biopower even as it promotes multiple ways of thinking about how life operates. For Canguilhem, concepts aren’t overarching proclamations on the lawlike nature of biology; they are modest, historically informed generalizations on how things tend to react in specific circumstances. In a wonderful exposition on the role of “concept” in Canguilhem’s thought, Henning Schmidgen argues that Canguilhem’s emphasis on concepts draws him nearer to a halting and hesitating form of structuralism: “Concepts are smaller and more mobile entities than structures, and even if they display an interior stratification they keep a certain sedentariness and palpability.”66 Canguilhem’s writing is rife with examples for thinking about concepts in the history of biology, from detailed investigations of specific concepts, such as the reflex arc or cell theory, to a more provisional use of concepts such as vitality in his methodological musings.67 Schmidgen also warns against thinking about concepts as a “one-way street” where subjects use concepts to explain relationships. Concepts can also move the other way, where “[l]ife produces forms that prepare, that ‘invite’ the formation of concepts.”68 Canguilhem’s emphasis on writing a history of life based on a history of concepts suggests specific tactics for how the study of biopower can generate insights into the incredibly diverse relationships between biological and political thought. As we already saw, Foucault appealed to two major biological concepts when thinking about life: that biopower is “homeostatic” and that it is composed of “populations.” In Foucault’s thought, populations are the point of application of biopower (as opposed to individuals, citizens, or communities), whereas homeostasis described biopower’s fluidity, or ability to respond to unforeseen events. It is interesting to see how many interlocutors after Foucault have tended to ignore the important role of homeostasis while emphasizing the role of selection in thinking about the dynamics of collectivities. Agamben, for instance, intentionally avoids the

Warped Grids · 119

role of homeostasis in politics as he attempts to demodernize population thinking as a form of zoë. On the other hand, Esposito shifts the concept of population toward the social register of community and then modernizes the idea of community through his ideas of immunitas. Yet, Esposito also attenuates the power of thinking in terms of homeostasis in favor of a less supple form of regulation, the negative dialectic. Both theories of biopower, then, focus on the role of populations as a form of community and, even then, only in the very attenuated form of considering the role of exclusion as a mechanism of community formation. I notice, however, that the proliferation of the use of grids for thinking about life demands an increase, and not a reduction, in the types of biological concepts that we use to explain the relationship between politics and life. Not only have I gone to great lengths to avoid thinking of grids merely in terms of selection at the level of populations, I have also tried to broaden an understanding of regulation beyond simple homeostatic mechanisms. To understand why this is important, we need to first understand the importance of regulation as a homeostatic mechanism.69 This is why the volumes on neoliberalism from Foucault’s late work (such as the Birth of Biopolitics and Security, Territory, and Population) have been so important for understanding the development of European forms of biopower. They demonstrate how homeostasis can account for aleatory events while allowing for continuity in the overall effects of regulation. Homeostasis gives grids their suppleness and their ability to respond. The metabolic concept of homeostasis is very important for Foucault’s conception of neoliberal regulation as well as Canguilhem’s conception of the “normal,” where being normal means possessing the capacity to respond to changing circumstances as opposed to fitting a statistical mean.70 Foucault’s adept addition of homeostasis as a concept for social regulation allows for a more nuanced view of how bodies and social forms are coregulated. Yet, homeostasis is not the only important biological concept that we should include here. The idea of the grid is also a biological concept, and casting it as such (instead of as a purely abstract rational concept) changes how we consider how grids might operate as regulatory mechanisms.

Using Biological Concepts to Warp Grids The concept of the grid has been more useful to biology than simply thinking in terms of organisms as constructed on a grid (the topic of chapter 4).

120 · Warped Grids

Grids specifically, and surfaces more generally, are part of a whole series of strategies that organisms use to maintain heterogeneity in their environment. Organisms use grids to live. A good example is how grids can act as filters. Filters can play Maxwell’s demon in that they allow for the circulation of some substances while partitioning off others. This allows organisms to concentrate important resources by effectively creating specific and sharp differences. A cellular membrane can be viewed as a biochemical filter that allows some substances to be absorbed into the cell while others are kept outside of it. A spider’s web is an even more appropriate example for this chapter’s exploration into the life of a fly. The spider uses its web to filter and immobilize flying insects, allowing the spider to concentrate food that might be too diffuse to otherwise locate. Key to both examples is how living things use filters to create sharp differences in the concentration of substances. As Valerie Ahl and T. F. H. Allen note in their book on biological theory, Hierarchy Theory, “Surfaces will spontaneously form when there is a significant gradient in concentrations of information, energy, or matter. Surfaces amount to local places across where there are significant differences.”71 Although we may want to interrogate what is implied by their use of the term “spontaneous,” it is exceptionally important to note that grids and surfaces are important tools for creating localized pockets of heterogeneity. Yet, as we saw in chapter 2, grids are used to create order upon a surface (as in the page layout of a graphic designer) and not just across it (as in the operation of a filter). Organic conceptions of space tend to emphasize the relationship between the inside and outside of a form (see Figure 3.6a), setting up a dichotomous relationship between the outside of a surface and its inner depth. This is the relationship that we see with the grid as a filter, where a surface constrains the passage of certain items predicated on the parameters of inclusion (as applied by some form of immunitas, whether that be political conceptions of individuality or chemical affinities of a protein on a cellular membrane). This would be very similar to the effect of inclusion and exclusion in Haeckel’s illustrations of primates in chapter 1, where a branching tree of primates containing humans was made more palatable for a European audience by not including white Europeans in the tree. This type of difference is a product, however, of thinking in terms of whole forms, where it is important to notice what is included in the form and what is not. In our investigation of the relationship between image, animation, and biology in the late twentieth century, we will continue to note that opera-

Warped Grids · 121

tionally, surfaces must no longer only be seen as dialectically opposed to depth. In fact, in some cases it is instructive to think that surfaces might be a precondition for depth. For instance, multiple surfaces layered upon one another can build an experience of depth through the multiple surfaces. This is similar to how a cross section of a tree exposes the rings created by the tree’s growth. Each ring was once a surface before new surfaces were layered over it. This draws attention away from a simple idea of immunitas as a form of exclusion or exemption to a consideration of how the order on surfaces ends up creating exclusions by regulating the placement of things (see Figure 3.6a and 3.6b). Consider a cellular membrane. Clusters of proteins can act as selective gates to allow certain molecules to travel, or send a signal, across the membrane of the cell to its interior. This seems like a straightforward application of the filtering effect of Figure 3.6a. Closer examination brings up interesting questions about how these proteins came to be on the surface in the first place or why these proteins even cluster in a pattern. Understanding these questions could lead to an understanding of how other logics for creating order—­other grids, if you will—­influence the placement of molecules across the surface of the cell membrane. For instance, one can now profitably ask how the matrix of the cellular cytoskeleton influences the patterns of proteins on cellular surface. To sum up, in Figure 3.6a, the politics of the sphere is predicated upon exclusion. In Figure 3.6b, the politics of the sphere is predicated on how heterogeneous elements are ordered to create patterns on the surface. Both types of formal relationships possess a politics, and in practice they often occur together. It is very important to recognize, however, that these types of political engagements with spheres promote very different outcomes. The politics of exclusion creates political differences based on who and what is left out (Figure 3.6a). The politics of regulation creates political differences based on how elements are included (Figure 3.6b). The difference between these two forms of politics is immense. Returning to Curt Stern’s ode to the fly is instructive here. The hard exoskeleton of the fly created a rigid politics of exclusion, where moisture was imprisoned in the body. The mammalian ability to regulate moisture through its increased volume-to-surface ratio allowed for the suppleness of human skin. Regulation allowed humans to have a much more interactive relationship with their environment. Theories of politics and bodies need a more refined conception of how supple and unyielding forms of oppression in contemporary society operate together. Thinking in terms of biopolitics

122 · Warped Grids

Figure 3.6. Surfaces and Difference. A decrease in concentration occurs across the surface in 6a, often thought of as a politics of depth. The decrease in concentration over the surface in 6b is what I call the politics of forms. Image courtesy of Karl Benitez.

can help, but it would need to begin to recognize how a variance of biological strategies, or concepts, occur when power circulates through society. As Jeffrey Nealon recently argued, “It may seem banal to point it out, but within all the discussion of biopolitics and posthumanism, there seems a lot more discussion on the ‘politics’ end of biopolitics than there is rumination on the ‘bio’ part of the story.”72 A critique of life should not just be a critique of the concept of life, it needs to incorporate how concepts can help us see how lives creates differences. To appreciate the role of regulation in the world is not to deny the reality of the outside, as some might claim. Rather, as Michel Serres grasped, we now have to recognize how the outside is no longer the only place where pests lurk; they also lurk alongside, or even within us as parasites.73 This is why biology is going to continue to be an important site for critical political and theoretical engagement. Despite the death knell sounded by engineers and a few critical theorists, living things will remain intriguing as studies in how irreducible composites interact. Irreducible does not necessarily mean unexplainable, it just means that each explanation will require some form of pragmatic, experimental, or virtual engagement. Locating warped grids can be an especially important indicator for understanding how different forms of orders can permeate and change one another. Thus, thinking in terms of grids as a biological concept has the capacity to turn grids from idealized mechanisms for creating order into sensitive material mechanisms for discerning different types of orders.

Warped Grids · 123

Understanding the Politics of Warped Grids In a remarkable and surprisingly curt passage in the book Making and Unmaking the Grid, Timothy Samara offers a bit of advice on the politics of grids. “Amid discussions of race and gender, conservation, political empowerment, and civil rights, perhaps a simple conversation about where to put things—­the ‘mundane housekeeping’ of grid based design—­might have value again.”74 Taken on its own, this passage can be read as a politically conservative call to keep things in their place. More conscientiously, considering how grids might relate to one another, opens this statement to a more radical form of politics. Grids can give thinkers a way to contemplate how complex conception of order might possess potential strategies for radical politics. Impressively, some books on critical race theory seem to have come to similar insights. Although none of the books that I mention use the concept of grid, per se, they do appeal to mechanisms for thinking about how things are ordered beyond a dialectical politics of exclusion. In Habeas Viscus: Racializing Assemblages, Biopolitics, and Black Feminist Theories of the Human, Alexander G. Weheliye productively warps Gilles Deleuze and Félix Guattari’s idea of “assemblages” by racializing it. This allows him to see how political forms of oppression come into being through the creation of relationships other than exclusion. The idea is important for my work for two reasons. First, there is a dynamic to assemblages that goes beyond the mere fact of making connections. Assemblages work because of their allowance for heterogeneity among the elements that are being assembled. This is why an assemblage is important: the connection of two elements brings something to the table that either element can’t on its own.75 Much like what we saw with grids, the elements in an assemblage are similar enough that they can be put into relationship with each other; they are different enough, however, to make multiple forms of arrangements possible. This allows for an understanding of a racializing assemblage as “structured in dominance composed of the particular processes of bringing-­into-­relation.”76 Much like what we saw in chapter 1, where we needed to imagine living forms beyond organisms, we also need to rethink biopolitics beyond the categories of exclusion and selection if we want to use the concept to effectively think about twenty-­first-­century forms oppression. The liquid blackness project from Georgia State University is another significant project for thinking about how racial oppression occurs

124 · Warped Grids

through creating associations and focuses even more directly on the role of aesthetics in racial discourse. I believe this is one of the more interesting projects in critical race studies as it tries to understand how blackness as a concept morphs and changes and tries to “leverage, rather than condemn, this type of mobility.”77 Mobility by itself is not the necessary goal of the project, as liquid forms can be especially open to appropriation. Rather, it remains important to look at how blackness moves and is embodied, and thus how order helps to constitute an aesthetics as well as a politics. As Alessandra Raengo argues: If blackness is placed firmly in the middle, held at the center of our conversations, affective investments, aesthetic concerns, if it is therefore made accessible, discussable, touchable, usable, re-­purposable, then the focus might shift to new considerations: not what it represents, but what it does and can do, to its affective charge, and its sensorial reach; to the relations it facilitates, the fantasies it coagulates, and the sensible and sensorial configurations it orchestrates. One would therefore not be seeking a black aesthetic but rather to understand blackness as aesthetics.78 This, I believe, is a powerful approach to investigate how racial oppression manifests itself through and because of inclusion. It provides an especially important critical lens to investigate how desire, bodies, aesthetics, and regulations combine to create an inequity of place for all life in the grid(s). My point in offering these examples is twofold. First, I want to suggest that studies in aesthetics and form are already influencing the discussion of the regulation of bodies in interesting ways and that these studies can be used to understand how a politics of power orders specific conceptions of life. Second, I want to suggest that paying attention to how pests are generated can provide a sensitive indicator for finding warps, if not holes, in grids, as productive and interesting places where multiple ways of thinking about order influence one another.

Grids in Relationship with One Another Meaningfully, a viewer can use the distinction between exclusion and regu­ lation to understand how the horrific wayward regulation, or warping, of a grid by the integrator/disintegrator is rectified in the original movie ver-

Warped Grids · 125

sion of The Fly. The mutated fly/human pest (the one that has a human head and arm attached to a small fly body) generated by the neoliberal desire of perfect distribution, ends up being caught in a new type of grid, a spider’s web (Figure 3.7). This can be read as the intersection of two different stories about consumption. Although the fly/human pest is generated by the misregulation of one circuit of consumption (the creation of the ability to distribute goods for humans), it ends up being the main course in another circuit of consumption (the ability to productively arrange goods for spiders). This mixing of grids ended up being too horrific for midcentury screenwriters set on preserving an unwarped view of white suburbia. In the movie, the horrific play of power across the surface of the spider web is swiftly replaced by the finality of a politics of inclusion and exclusion when the police chief overhears the eerily high-­pitched cry of “Help me!” from the fly/scientist and smashes the repulsive little drama with a rock.79 This unambiguous return of the strong arm of the law changes the ending of the movie from a disaster of biopolitical regulation (where the spider gets to eat the fly/human), to an exercise of hierarchal sovereign power (where the police chief exercises the power over life and death). This is an especially important lesson about the politics of aesthetics in relationship to regulation. Laissez-­faire regulation of biopower is only tolerated when its ends are closely aligned to an assumed hierarchy of power. Aesthetics and the role of affect in society are effective mechanisms to achieve this end. Yet it is also a caution for those of us wishing to condemn all forms of flexible order as a product of neoliberal sensibilities. To do so diminishes the very challenges that pests present to grids. Some forms of generativity create new forms of order that allow for new, perhaps even liquid, arrangements between desire, aesthetics, and politics. Neoliberal economies, on the other hand, can only continue to thrive by reducing the fecundity of these pestilent cases to blatantly humanist outcomes. Or, in terms of the ending of The Fly, consumer-generated economies conserve their values by ensuring that the correct subject continues to be consumed. A similar appeal to a more straightforward type of power marks the future of the human-­sized human/fly as well. The consequences of the faulty reintegration of the scientist’s body parts are too much for the human/fly to bear, and he pleads with his wife to kill him before the fly part of his body takes over the human part. The wife does so by using a mechanical press to smash the limb and the head of the scientist into an unrecognizable pulp.

126 · Warped Grids

Figure 3.7. A scene from the ending of The Fly (1958) where the tiny fly/scientist hybrid is caught in a spider’s web. The Fly (1958), DVD, 20th Century Fox, 2013.

In an ironic twist, the industrial system that the scientist had hoped to transcend is what puts an emphatic end to the unfortunate consequences of his experiment. In both cases of extermination, justice appears to be served when the troubling forms of bodies are reduced from dynamic forms to simple substances. The easiest tool to accomplish this is to reduce complex patterns of interactions into simple forms of exclusions. In a world where new forms of regulation threaten to create new forms of life, it seems conceptually easier to return to an exclusionary form of politics where bodies, consumption, and production can seemingly be kept separate by reverting to an older and more comfortable mode of existence. The gruesome ends of the fly/human hybrids also provide a lesson about the important role of images in cultural regulation. This brings our discussion of the movie back to its publicity poster. Industrialization, laws, and analysis are very good at breaking objects and bodies down into pieces to control. These pieces, however, are often put back together through image-­based technologies. Imaging technologies allow us to envision how these pieces can fit together to create new types of orderings. They are key for understanding how things are regulated in complex conceptual systems. As Vilém Flusser writes in his book Into the Universe of Technical Images, “The whirring particles around us and in us must be gathered onto surfaces; they must be envisioned. We already have the visualizing power needed to do this.”80 Our mass-­produced images were our first integrator/ disintegrators. Understanding the politics for how things are put back to-

Warped Grids · 127

gether depends upon developing methods for thinking about the politics of surfaces beyond the dynamics of exclusion. Perhaps it is this stubborn attempt to envision the consequences of how a world is ordered that allows the story of The Fly, much like its prolific namesake, to keep breeding. The original 1958 version had two sequels, The Return of the Fly and The Curse of the Fly. As we have seen, David Cronenberg subsequently directed a retelling of the original in 1986 starring Jeff Goldblum as scientist Seth Brundle and Geena Davis as reporter Veronica Quaife. This, in turn, spawned yet another sequel, The Fly II, directed by special effects specialist Chris Walas and starring Eric Stoltz as the hybrid lovechild of Brundle and Quaife. There have also been comic books, operas, and much discussion of other remakes and sequels. The story, it seems, just can’t be exterminated. And this is the most important political lesson from our studies of pests in the grid(s). Controlling life doesn’t just mean eliminating it but can also mean promoting specific types of lives at the expense of others. Every grid we put in place not only regulates, it promotes the life of new types of “pests.” These pests are lives that are not wholly excluded but are discouraged from existing outside of their proper place; they must exist as a pattern whose very liquidity can be dangerous. Each of these different types of orders possess different “degrees of freedom” for developing new potentials. Understanding the generative properties of regulation forces us to ask if we are prepared to finally recognize that regulations not only breed political economies but new types of lives as well. In a world where regulation is as important as exclusion, we must also learn to ask under what terms inclusion occurs and what the consequences are. Lives depend on it.

This page intentionally left blank

4

Modulations Envisioning Variations

In the 1970s, grids advanced from being a means of displaying organisms to a means of conceiving how organisms evolve and develop. It was the modularity of grids, their ability to build complex forms from similar simpler elements, that biologists found useful. This resulted in new ways of thinking about how organisms self-­constructed during development as well as how organisms were related through evolution. Some scientists suggested that this new way of thinking might be a new discipline or field of study, calling it “evolutionary and developmental biology.” The long descriptive phrase was eventually shortened into the sleek and futuristic, “evo devo.” As Jessica Bolker has argued, “For evo-­devo, studying the developmental assembly and integration of modules is central to understanding how they and their interconnections may originate, break down, and change through evolutionary time.”1 This chapter will examine how two thinkers used illustrations to help articulate a modular theory of life: William Bateson and Edward Lewis. This narrow field of inquiry is not intended as a claim that these are the only two scientists who contributed to biological theories of modularity. On the contrary, part of the attraction of this way of viewing organisms is that it comes from many different disciplines in biology, such as comparative anatomy, development, and evolution. Consequently, modularity provides a robust framework for thinking of organismal form that helps synthesize many fields of research. As Manfred Laubichler has claimed: “Modularity is a fundamental characteristic of all biological organizations; it is thus a universal principle of biological form.”2 In this chapter, I intend to add to Laubichler’s insight by asking exactly how modularity works as a formal concept and why it might be different from other conceptions of form in the history of biology.

129

130 · Modulations

I am also interested in Bateson and Lewis because both scientists identify as geneticists. Bateson, in fact, is credited with coining the term “genetics” as well as other key terms in the discipline. I believe these two thinkers offer a bit of a corrective to overly simple criticisms of genetics as being a biological field dominated by a determinist outlook. When most people think of genetics, they think of a very simplistic notion of transmission genetics, where the gene supplies information that determines the construction of bodies, much like a blueprint helps determine the construction of a house. This oversimplification is unfortunate, as it downplays the very important role that anomaly and difference play in the history of genetics. It would be hard to conceive of genetics as a discipline if not for variations in body type, such as created through mutations (and other sources), as well as observations of the complex behavior of genes as witnessed in experimental practice. Some genes appear to move from place to place on chromosomes, and some genes were thought to affect other genes by being in close proximity to one another. Bateson and Lewis are two thinkers who eschewed easy interpretations of experimental results as they attempted to understand how heredity could produce a variety of bodily forms. They even used different aspects of modular thinking to explain how similar biological processes could create very different types of organisms. These different forms of modularity offered both scientists a way of considering how the development occurred as the product of a complex process of regulation and not molecular determination. Turn-­of-­the-­twentieth-­century biologist William Bateson believed that he could demonstrate how organisms were related to one another by showing how they varied. This insight encouraged Bateson to think about the development of the organism as a product of internally produced varia­tions, with their own laws, as opposed to an organism’s relationship to its environment. Variations between animals, in Bateson’s eyes, were created like the waves of the ocean creates ripples on a sandy beach. What we witness, the ripple on the beach, was the product of unseen forces of a larger process. This view of life stressed the importance of the segmental construction in animals, where organisms were thought to be constructed from a repetition of parts, each segment akin to a single ripple on the beach. Mid-­twentieth-­century drosophila geneticist Edward (Ed) Lewis had the interesting insight that variation could occur in a stepwise fashion and that he could use the tools of genetics to study the steps. This allowed Lewis to postulate a modular theory of development, where specific parts

Modulations · 131

of the fruit fly developed as a set of discreet stages, or modular genetic and biochemical interactions. It wasn’t just that these modules were repeated to create a complete organism, they had to be repeated in a specific sequence and in a specific location. Some variations between organisms could be caused by changing the steps of construction of an organism, as opposed to changing the whole blueprint of its design. To understand why these insights were so valuable, it is helpful to first explore in more detail why thinking in terms of grids and modules was such an important conceptual innovation.3

What Is a Module? “In the final decades of the twentieth century,” writes historian Andrew Russell, “experts in a wide variety of disciplines such as computer science, evolutionary biology, management studies, and educational theory introduced the concepts of modular design into their professional discourses and practices.” Sometime after the mid-­1970s, continues Russell, “Modularity became a way of seeing, knowing, and ordering.”4 The clarity with which Russell writes about modularity, however, masks the conceptual difficulty involved in puzzling out how modules worked in specific disciplines. In my research I have found that contemplating the history of modules opens a historian’s version of a wave-­particle paradox. The ubiquity of the occurrence of the concept in economics, design, biology, and informatics suggests that a broad historical multidisciplinary approach might be appropriate, so that readers can appreciate the wavelike surge of modular thinking across many disciplines. The concrete specificity of the circumstances of each discipline’s embrace of modularity, however, suggests a discipline-specific approach, where the reader can appreciate the particle-­like integrity of how modules were used to solve unique problems in each discipline. In fact, the more one stares at a module, the less one seems able to locate it with any specificity. Modules seem to be everywhere (in that they are distributed), nowhere (in that they are an organizational concept and thus part abstraction), and right in front of your face (in that they inform specific material practices). And all at the same time. Part of this paradox is built into the concept of modularity itself. According to biologists Gerhard Schlosser and Günter P. Wagner, a module is thought to be internally integrated, allowing it some form of autonomy from its surroundings, as well as actively part of relationships with other

132 · Modulations

modules, allowing it to coordinate functions with other modules despite its autonomy. “Two aspects of modules are emphasized by [structure-­or process-­oriented] perspectives, their integration concerning their internal relations (between their components) and their autonomy concerning their external relations (to elements of the context).”5 Modules are useful organizational concepts in that they allow a thinker to identify how an element can be unique, through its relative autonomy, as well as how unique elements can then be related to one another, through their ability to be integrated. In a sense, modules operate as a type of nonlinear bridge between specific elements and the arrangement of these elements into larger assemblages. In this sense, modules operate in the same fashion that we saw with grids in chapter 2. They allow for the partitioning of elements from one another, a moment of breaking down, and then the reassembly of constitutive elements into a newly integrated whole, a form of building back up. This chapter will show in detail how the emphasis on defining biological forms in biology in the late twentieth century was not built on a return to a Kantian relationship of the parts to the whole. In fact, modules provided the benefit of allowing for a conception of development not predicated on the creation of complex final forms from a single set of instructions. The forms of the living now appeared to emerge from the act of repetition itself, where stepwise iterations of a similar process combined to build a diversity of shapes, sizes, and developmental pathways without recourse to a normatively perfect form, but still tightly regulated in its method of development. What emerges is less a romantic symphony of flesh, bone, and chitin predicated upon proper stages of development, than a series of (neo)baroque variations, that through their constant repetition reveal the themes that generate their variance. Again, the insights on the properties of technical images provided by Vilém Flusser help link the visual practices of creating twentieth-­century biological images with the content of twentieth-­century biological theorizing on evolution and development. Modules, like surfaces of combined texts and images, promote the clarity of historical causality of textual thinking at the same time they allow for the more associative thinking promoted by images. For instance, imagine panels on the page of a comic book but imagine they were not created to tell a specifically linear story. Instead, these panels were created in a way that each one was meaningful and yet could be placed close to other panels to add to their meanings.6 As

Modulations · 133

we learned in chapter 2, the panels of a grid are often referred to as modules for this very reason. True to its nonlinear nature, modularity was not a concept that pristinely emerged from a single thinker and was then immaculately passed to others. Rather, it was a strategy that many disciplines adopted to imagine how complex objects could be built from simple pieces. This partially explains the difficulty in locating the origins of modularity. Some thinkers date the advent of the term “modular” to the introduction of system architecture of the IBM 360, a historical account popularized by Carliss Baldwin and Kim Clark in their book Design Rules: The Power of Modularity.7 Andrew Russell persuasively suggests, though, that to do so misses the history of modularity in other contexts and mischaracterizes how ubiquitous the concept of modules was in other computing platforms, let alone other disci­plines. “In each of these disciplines,” writes Russell, “modular systems called for standardized, interchangeable components (or modules) that could be recombined within a predefined system architecture.”8 Any field of inquiry that attempted to build complex assemblages using standardized parts would prove a fertile breeding ground for modular thinking. By the mid-­twentieth century the discipline of molecular development was one of these breeding grounds. Historians of biology have recently demonstrated that the desire for standardization that marked much of industrialized life also marked the history of biology. Scientists used industrialized terms for describing large scale breeding experiments, standardized pure-­line organisms to clarify the results of these experiments, and appealed to a modern conception of purity as a lowest common denominator for viewing the relatedness of all living beings. Thinking of organisms as modular helped biologists grapple with how diversity could be created from the use of standardized parts in much the same way that it did in other fields. The introduction of modularity in biology allowed researchers to understand how different developmental outcomes for an organism, such as why one organism develops arms while another develops wings, could be derived from similar additive processes. Thinking of organisms as composed of modules, however, demanded different ways of viewing how organisms operated. During the twentieth century, the evolution of organisms began to be envisioned as a type of tinkering based on creating new traits from the novel use of parts on hand, rather than on a singular grand act of design.9 During the twentieth

134 · Modulations

century, scientists increasingly began asking “how can small shifts in processes lead to large changes in organisms?” In fact, in some cases organisms no longer held the pride of place they once did. Modular molecular pathways made up cells, and cells made up segments, and segments made up organisms. Organisms, in turn, made up ecologies. All were thought to be composed of networks of interactions. The simple dialectic relationship between parts and the whole was replaced by a complex series of nested and loosely associated modules.10 The lessening of the focus on whole organisms didn’t mean that biologists stopped concentrating on issues of form or that the Kantian linkage between life and aesthetics had been riven. Rather, a very fluid aesthetics emerges based on adding together simpler forms, as opposed to an emphasis on the dialectical relation of the architectural integrity of whole organisms and the vitality of the materials of life. The organic formalism of biology, pursued with such vigor by Ernst Haeckel, progressed into an emphasis on the fluid nature of pattern formation. Architectural historian Sanford Kwinter called this “true formalism”: “What I call true formalism refers to any method that diagrams the proliferation of fundamental resonances and demonstrates how these accumulate into figures of order and shape.”11 Although I see different types of formalisms in operation in twentieth-­century biology (and thus would like to promote a variety of true formalisms), the arguments below demonstrate how repeated modules, not the constraints of wholes, contributed to a new way of envisioning organisms during the twentieth century.

What Does It Mean to Think in Terms of Modules? Given the historiographic paradox of modularity, in that it seems to appear everywhere and nowhere at once, I decided the best explanatory approach for this chapter would be to render the abstract concept of “module” in biology understandable and then to see how it develops in specific circumstances in the history of thinking about heredity and development. I once again look to the history of manufacturing and the concept of games to concretely explain the abstract idea of modular assembly. Consider the popular mid-­twentieth-­century board game, Cootie (see Figure 4.1). Origi­nally distributed in 1949 by William Schaper, the goal of the game is to build your own insect, or cootie, from a collection of manufactured plastic parts. Players roll a die to collect a part labeled with a corresponding number. The challenge of the game is that it requires players to roll a

Modulations · 135

Figure 4.1. The Game of Cootie. This is an original 1949 release of the game. Schaper games are notable in that they were one of the first companies to use predominantly plastic parts. Photograph courtesy of MediaCommons.

die in a specific sequence. You first need a thorax, which you receive by rolling a one, then you need to place a head on your thorax, which you receive by rolling a two. You continue adding the different body parts in order: antennae, eyes, proboscis, and legs. The winner of the game is the player who builds a cootie in the fewest number of die rolls. The Cootie example well illustrates how a relatively complex organism can be created from separate parts by regulating the order of construction. Consequently, it provides a concrete example for how development can occur through a sequence. It is much too simple of an example, however, for capturing how variation could introduce different outcomes using the same sequential logic. To visualize how variation works in development, it is a bit more accurate to imagine you are the manufacturer of the game and need to make each part from an unremarkable lump of plastic. As a manufacturer, you would then introduce a procedure to turn the lump of plastic into identifiable organs, such as thorax, head, legs, proboscis, and so on. The various parts of your cootie would then be created through a similar manufacturing process. However, this example still isn’t complex enough to explain how modularity is thought to work in bodies. Even though this new set of variations

136 · Modulations

allows us to envision how multiple copies of the same organism can be made from similar materials manufactured into different parts, it doesn’t provide enough information for us to imagine how different types of organisms can be created from the same parts. To take this next step, we need to imagine how each of these individual parts is subject to its own variations, where heads could change shapes, legs could change to arms or even wings, the proboscis could unfurl and shorten. The first series of variations is needed to introduce variation between segments on any organism (the head, the thorax, the abdomen), and the second series introduces variations to each of these sets of variations (changing legs to wings, for instance). The players of this new game can now create a great number of different creatures by following the same simple set of rules.12 It is this combination of two series of variations—­the variation that creates different parts from a single type of part and then the subsequent variations that each of these different parts can undergo—­that evolutionary and developmental biologist Sean Carroll has in mind when he dates the introduction of modular thinking in developmental biology to two exceptionally significant papers published in 1978 and 1980. The first is Ed Lewis’s 1978 Nature publication on the regulation of the drosophila bithorax complex during development, and the second is Christine Nüsslein-­Volhard and Eric Wieschaus’s 1980 Nature publication on mutations affecting segmentation during drosophila development (we looked briefly at one of the illustrations from this paper in chapter 2). These three researchers shared the Nobel Prize in Medicine in 1995 for the discoveries presented in these papers. As Carroll enthused in his Endless Forms Most Beautiful: The long drought in embryology was eventually broken by a few brilliant geneticists who, while working with the fruit fly, the workhorse of genetics for the past eighty years, devised schemes to find the genes that controlled development. The discovery of these genes and their study in the 1980s gave birth to an exciting new vista on development and revealed a logic and order underlying the generation of animal form.13 Although some may disagree with Carroll’s characterization of a mid-­ twentieth-­century drought in the studies of development (there were after all many very important papers published during that time, including some important conceptual developments in modularity and develop-

Modulations · 137

ment14), few will miss Carroll’s palpable excitement about the important role these specific experiments played for understanding the genesis of animal forms. What especially interests me is the similarity of process used in creating different animal forms, described by Carroll as the “logic of making a series of initially similar modules and then making them different from one another.”15 The idea of how thinking in terms of modules can add a new dimension of variance to thinking in terms of developmental sequences is Carroll’s point. The question is no longer “how did this structure come to be?” but instead “how can so many different structures come to be created from the same process?” This is the power of thinking in terms of modules. It no longer is an issue of how the form of a specific organism came to be but how a seemingly homogenous set of building blocks can lead to the creation of diverse animal forms. The rest of this chapter explores two important moments for establishing modularity as a mode for thinking about organisms. Focusing on these two moments allows me to highlight key conceptual innovations that led to a modular aesthetics of forms for practitioners of evolutionary and developmental biology. I begin with William Bateson’s late nineteenth-­century work on variation. Originally published as a critique of Darwin’s theory of natural selection, many of Bateson’s observations on variation weren’t widely appreciated until the late twentieth century. Two things emerged in retrospective importance: Bateson’s attempts to think of organismal forms in terms of the influence of different types of variations and his realization that animal forms can be made through a repetition of the parts. I then turn to Ed Lewis’s painstaking studies of the genetics of the bithorax complex in drosophila. Lewis worked on what he called “the levels of development” of related genes, where the protein products of genes were produced in a specific sequence based on the local control of where in the animal the genes were expressed. This allowed for an understanding of how variations produced by a single stepwise process could produce different outcomes in different tissues.

William Bateson and Variation The controversial turn-­of-­the-­twentieth-­century biologist William Bateson has found new life as a pioneer of important concepts in evo devo. Doubted by Darwinians for his emphasis on discontinuous change in

138 · Modulations

evolution,16 questioned for his mathematical ability by statisticians,17 mocked by the Morgan School of geneticists for his refusal to acknowledge the chromosomal theory of inheritance,18 and ridiculed by population geneticists for misunderstanding the species concept,19 Bateson now is treated more as visionary than villain in popular accounts of the history of evolutionary and developmental biology. Mark S. Blumberg, in Freaks of Nature: What Anomalies Tell Us about Development and Evolution, portrays Bateson as an alienated champion of the study of anomalies.20 Sean Carroll in Endless Forms Most Beautiful and in the multiauthored From DNA to Diversity lauds Bateson as a pioneer in thinking of animals in terms of modularity.21 Rudolf Raff points to Bateson’s pivotal role in thinking in terms of homeotic variation.22 Lewis Held Jr. refers to one of Bateson’s most speculative projects, his conviction that segmentation might be an effect of wavelike properties (a topic we will cover in-­depth in chapter 5) in How the Snake Lost Its Legs.23 All of these authors found passages in Bateson’s writings that foreshadowed their own concerns. What exactly is it that has elevated an embattled turn-­of-­the-­twentieth-­century geneticist like William Bateson, and not a morphologist like Haeckel, to the position of ur-­theorist of evolutionary and developmental biology?24 The short answer to that question is a growing appreciation of the role of “variation” in evolutionary and developmental biology. The concept of variation provided the foundation for William Bateson’s biology. As he writes in Materials for the Study of Variation, “To collect and codify the facts of Variation is, I submit, the first duty of the naturalist.”25 Bateson believes that a collection of known variations would give discussions on evolution an empirical basis by providing a “nucleus” of observations as well as help to elucidate the laws of descent. By providing a compendium of how organisms varied, Bateson hopes to identify the laws for how variations might occur. He supposes that this compendium of observations would correct theories of evolution that relied upon natural selection. The problem with natural selection, writes Bateson, is that its focus on the relationship of the organism in an environment, overemphasizes the role of gradual variations in evolution. Changes in environment are almost always gradual, surmises Bateson, so a theory of selection predicated upon an organism’s relationship to its environment mirrors the gradual changes of the environment where it lives. And although he recognizes the existence of gradual, continuous changes in organisms, gradual changes do not reflect the diversity of variations that Bateson witnesses in organisms,

Modulations · 139

nor do they reflect the discontinuous nature of speciation in general. Bateson thinks that changes in speciation are changes in type and not degree. We have seen that the differences between Species on the whole are Specific, and are differences of kind, forming a discontinuous Series, while the diversities of environment to which they are subject are on the whole differences of degree, and form a continuous Series; it is therefore hard to see how the environmental differences can thus be in any sense the directing cause of Specific differences, which by the Theory of Natural Selection they should be.26 Instead, Bateson suggests that the abundance of discontinuous changes must be due to some form of internal force that hasn’t yet been characterized. Bateson then proceeds to class all variations in organisms into two different categories so that he can better understand an organism’s internal influences on variation. He calls the first type of variations substantive, in that they constitute variation in the “actual constitution or substance of the parts themselves.”27 He calls the second type of variations meristic, as they change the number of segments, the placement of the segments, or the overall geometry of the organism. Bateson uses the example of a dog’s spine to explain the differences between these two types of variations. A substantive variation for a dog’s spine would change the form or substance of a single vertebra, whereas a meristic change would change the number or placement of vertebrae in the composition of the whole spine. However, Bateson spends little time describing substantive changes as he intends to cover them in more depth in a second volume of the Materials, a project he never finished. Consequently, the bulk of the book reads like a catalogue of meristic variations bookended by his theoretical observations on the importance of variation and speculation on how it occurs. Bateson concludes that the study of meristic variability leads to two major insights in the composition of organismal form. The first is the law of symmetry, that all animals are composed along an axis of symmetry. The second is the law of repetition, that animals tend to be composed by a repetition of their parts. The most important of these laws is the repetition of parts, as it is the most generally comprehensive. As Bateson notes, the development of symmetry can be viewed as a special case of repetition, where one side repeats the other in a mirrorlike fashion.

140 · Modulations

This phenomenon of Repetition of Parts, generally occurring in such a way as to form Symmetry or Pattern, comes near to being a universal character of the bodies of living things. It will in cases which follow be often convenient to employ a single term to denote this phenomenon wherever and however occurring. For this purpose the term Merism will be used.28 Meristic changes were especially revealing changes in that they suggest how the law of repetition effects animal forms. Changes result because organisms create discontinuity by being composed of discrete sections. The key for understanding Bateson’s theory of variability is appreciating that the sectioning off of an organism during evolution and development is creative. Creating sections provides discontinuities to what would otherwise be a homogenous substance: “The existence of pattern implies a repetition of parts, and repetition of parts when developed in a material originally homogeneous can only be created by division.”29 Although conceiving of organisms as repeated parts is not original,30 no other biologist had elevated this observation into an insight on how evolution might occur. The importance of Bateson’s arguments are not that he introduces the concept of repetition in biological forms, but on his insistence that repetition is the primary organizing principle for variation followed by his extensive documentation of how these variations present themselves. Repe­tition creates new segments as it organizes an organism’s form. Bateson used many strategies for making his argument. The most obvious is his encyclopedic approach. Weighing in at over six hundred pages, Materials reads like a catalogue of discontinuous variations in different forms of animals. To make his point, Bateson attempts to convince his readers that variation “is cosmic not chaotic.”31 In other words, he needs to give enough examples to demonstrate that there is some type of order behind the variation. He does this by interweaving textual descriptions of organisms with tables of data, graphs, and illustrations. Bateson’s use of images is especially illuminating for understanding his presuppositions about animal forms. Many illustrations provided by Bateson present two or more examples of variance of a specific character. This illustration strategy allows readers to see how variations of characters create a pattern of differences, what Bateson refers to as the discontinuous sequence of variance of a character, rather than just illustrating characters in themselves. As Bateson notes: “The Study of Variation is essentially a study of differ-

Modulations · 141

ences between organisms, so for each observation of Variation at least two substantive organisms are required for comparison.”32 For Bateson, variation is always a form of difference that only appears through comparison. Take for instance the illustration of the butterfly, Satyrus hyperanthus [sic] (Figure 4.2). Here he presents a lithograph of the variance of eyespots on butterfly wings as changes in placement and numbers of spots. This is an interesting illustration for several reasons. First of all, it is one of the few cases in the Materials where Bateson uses a lithograph to depict whole organisms. This makes the illustration look like a natural history drawing, where an attention to the details of form and markings lends verisimilitude to the illustration. (Note how each set of antennae are rendered differently, giving the impression that these are illustrations of specific specimens.) The illustration resembles a display case with four specimens pinned to a mat. This visual emphasis on the fidelity of representation is further reinforced when Bateson refers to the second example in the illustration as “the most frequent form,” suggesting an emphasis on the variance between

Figure 4.2. Figure 77 of Materials for the Study of Variation. Here William Bateson compares four illustrations of Satyrus hyperantus, a common “ringlet” butterfly. William Bateson, Materials for the Study of Variation Treated with Especial Regard to Discontinuity in the Origin of Species (London: Macmillan and Co., 1894), 294.

142 · Modulations

specimens, as opposed to the more starkly binary depiction used later by geneticists, where variations are listed as either mutant or wild type. Despite the attention to detail in this illustration, in most cases, Bateson uses much more abstract images to demonstrate the repetition of parts. Take the following illustration of the placement of pig’s nipples (Figure 4.3). Here Bateson relies only on depicting overall patterns of nipple placement, effectively eschewing any representation of the overall form of the pig. This gives Materials a very different aesthetic from that found in Haeckel or even other less artistically oriented morphology texts. Bateson tends to collect illustrations that emphasize patterns of traits as opposed to forms of organisms. In fact, Bateson tends to eschew thinking in terms of an individual anything. As we saw above, Bateson thought that variations occur in series as opposed to single occurrences. Consequently, variation for Bateson was a creative propagative force and not the deformation of an ideal type. For Haeckel, the individual represents a dynamic archetype. The individual for Bateson is a single instance of a series of variations as variations constitute the forms. This shifts the explanatory strategy from demonstrating how parts can create archetypical forms to documenting enough variations that one can glimpse the laws for how patterns are organized. “Each individual

Figure 4.3. Figure 34 from Bateson’s Materials, where only the overall pattern of the placement of nipples in young pigs is emphasized. William Bateson, Materials for the Study of Variation Treated with Especial Regard to Discontinuity in the Origin of Species (London: Macmillan and Co., 1894), 191.

Modulations · 143

and each type which exists at the present moment stands, for the moment, therefore, as the last term of such a series. The problem is to find the other terms.”33 Patterns, not forms, structure Bateson’s biological world. It is this aspect of Bateson’s work that resonates so strongly with his son, Gregory, when he develops his father’s insights on the organization of biological variation into a seminal principle of cybernetics, named Bateson’s law after his father. This law states that when an asymmetrical lateral appendage is “reduplicated, the resulting reduplicated appendage will be bilaterally symmetrical.”34 So, if an individual was born with two feet on the left leg and no feet on the right, the two feet on the left leg would not be two left-­shaped feet. They would be a left-­shaped and a right-­shaped foot. According to Gregory, bilateral symmetry is the most likely outcome because it requires less information to construct symmetrical structures than asymmetrical structures since the same information could be used to construct each side. In creating this law, Gregory Bateson elevates his father’s desire to find the cosmic regularities informing biological variance as bold proclamations on the nature of information itself. Consequently, whereas the father, William Bateson, asks what happened to introduce symmetry, the son, Gregory, recognizes that bilateral symmetry requires less information. Bateson the younger read Bateson the elder’s work on biological variation as a problem in cybernetics. As Gregory explained: In the language of today, we might say that he was groping for the orderly characteristics of living things which illustrate the fact that organisms evolve and develop within cybernetic, organizational, and other communicational limitations. It was for this study that he coined the word “genetics.”35 For Gregory Bateson, William Bateson’s work in genetics was already an experiment in information and cybernetics. His father just didn’t have the vocabulary to describe it as such. Although there is some danger in seeing modern conceptual problems in past formulations, Gregory Bateson’s vision of his father as a proto-­ cybernetician is illuminating. The claim could be read as a totalizing tendency of information, where everything can be read as information. From this perspective, genetics has always been an information science. The problem with this view is that it ignores the historical and material specificity through which ideas can be formulated. There are reasons why

144 · Modulations

a relationship is called “information” in one era and “variation” in another era.36 Ignoring this difference in terminology downplays the historically unique circumstances behind the formulation of an idea. A more interesting reading is to think about how biology and information theory have in some ways been involved in a mutually compatible project. Most likely this was a project shared by other professions as well. An example of this is my argument from chapters 2 and 3 that understanding how complexity could be built from similar components was a challenge already found in advertising and popular culture and not just as an explicit goal for cybernetics. The knowledge that numerous fields were experimenting with how form could be built from repetition is an especially important insight for understanding the relationship of biological thought to other social developments. Some scholars have too often portrayed biologists as passive participants in a society of ever-­increasing control based on the imperatives of cybernetics, control, and communication.37 This type of thinking minimizes the biological contributions to information theory, cybernetics, and systems theory. It also suggests that those in the fields of cybernetics were somehow removed from a set of prerogatives that informed many aspects of twentieth-­century thought, such as the role of affect in the development of flexible accumulation. As I have been arguing, biology, as much as computing, contributed to making life calculable by helping to develop a concept of regulation that was flexible and open to change. It didn’t do this through reductive strategies, where computing reduced biology to a science of control; it did this through seeing how things were rebuilt from parts that had already been reduced. Perhaps this is one reason why William Bateson’s work has received so much attention from researchers writing about evo ­devo. Bateson provides insights into how variation could occur through “regularity and ‘lawfulness.’ ”38 It is not so much, as Carroll claims, that Bateson provided the first writings on modularity. As we saw, Bateson never really articulates a view of segmentation as autonomous as well as integrated. At best, what we have in Bateson is a theory of variation predicated upon the process of segmentation and not a coherent theory of modularity. When Bateson does reach for an analogy to describe the genesis of patterns in animals he appeals to harmonics, not manufacturing, and not to abstract ideas on control.39 What Bateson does supply is a way to think about how variety could suggest order. This is a very important consideration for thinking in terms of the regulation of aleatoric events in life and political economy as explored in chapter 3.

Modulations · 145

Bateson’s conviction that understanding variety reveals order is one of the reasons why geneticist Ed Lewis repeatedly references him in Lewis’s later writings. Journalist Stephen S. Hall even reported that Lewis had a picture of Bateson on the wall of his laboratory, along with Alfred H. Sturtevant, Lewis’s graduate supervisor, as well as Sturtevant’s supervisor and creator of the Caltech genetics group, Thomas Hunt Morgan.40 This is a strange homage for Lewis to pay, as Bateson’s refusal to believe in the chromosomal theory of heredity made him the subject of ridicule for the Caltech fly group. Understanding why Lewis eventually found Bateson’s work evocative is important for understanding why thinking of animals in terms of modules changes how one conceives of the way development is regulated.

The Mod ’70s—­Variations Are the Theme In 1978, Ed Lewis published one of the most frustratingly abstract, exquisitely detailed, and conceptually elegant papers in the history of molecular development. Miranda Robinson, an editor of Nature, contacted Lewis to see if he might publish a paper where he would “put your opinions, ideas, and facts all in one article.”41 When Robinson contacted Lewis, he hadn’t published a paper in thirteen years. His submission, simply entitled, “A Gene Complex Controlling Segmentation in Drosophila,” ended up condensing thirty years of experiments into five dense pages.42 The paper isn’t like many others. It does not present an argument, or even a specific set of experiments, driven by a hypothesis. What it does present is a vision, a series of images explaining outcomes of how genes might operate during development. As one of his biographers writes shortly before Lewis’s death: “Here, for the first time, a framework for thinking about how genes control development is presented.”43 By mating and characterizing an immense number of fruit flies, Ed Lewis explored how a sequential arrangement of gene expression in one multigene cluster might direct the segmentation of all fruit flies. As Alexander Varshavsky, one of Lewis’s colleagues at Caltech, observed, “What Ed was able to do, in the absence of molecular understanding, was to come up with a basic concept of how genes work . . . in terms of a logical relationship, like a diagram of an electronic circuit.”44 Lewis’s contribution to understanding development, his vision, is how a series of simple steps could make complex and varied outcomes. Although he had been working for decades on the logic of how stepwise changes in gene regulation could occur, it wasn’t until

146 · Modulations

the 1970s that Lewis’s work became known beyond the confines of a close circle of drosophila geneticists. Understanding Lewis’s newfound popularity depends upon understanding the context within which he worked and published his ideas. During the 1970s, there was a growing interest in the genetics of eukaryotic, or complex, organisms. Some researchers gained confidence that they had figured out enough about the role and structure of nucleic acids to systematically study how they functioned in complex organisms. One of the best-­known cases for this was the career of Nobel laureate Paul Berg, who began working on a eukaryotic virus, SV 40, in the middle of his career. Berg began studying a eukaryotic virus hypothesizing that he could use the viruses’ simplicity as a probe to elucidate the relatively complex structure and expression of the eukaryotic gene.45 Berg hoped to have the same types of success that early twentieth-­century molecular biologists had when they began using bacterial viruses to understand the intricacies of bacterial genetics. Ed Lewis was not one of these scientists. While others had only recently begun to study the genetics of eukaryotic organisms with strategies successful to bacterial genetics, Lewis eschewed simple analytical strategies in favor of working the old-­fashioned way, by breeding and screening hundreds of thousands of mutant flies. Lewis used established strategies in classical genetics that opened new possibilities for thinking about the role of genes in development. Because of this, Lewis appears to be one of those innovators in the history of biology who come to new understandings by pushing well-­established tools to the absolute limit. This is an especially important insight when we consider how much the field he helped to breed has changed in the last forty years. Today, visual images of embryos stained with multicolor fluorescent markers bound to monoclonal antibodies saturate many copies of Nature or Cell (see, for instance, Figure I.1 in this text’s introduction). These tools produce stunningly beautiful images vivid with color and intricate in detail. Lewis had no such tools and had to rely on the much older visualization technologies of looking at flies under a microscope, scoring them in notebooks, and occasionally staining their enlarged salivary chromosomes. Yet visualization was still an important quality for Lewis; in fact most studies of Lewis appeal to his strong visual sense as a keen skill in his work.46 It was his ability to provide an overall vision of how genetic regulation could occur that earned Lewis his Nobel Prize. Consequently, Lewis’s

Modulations · 147

use of diagrams and images was especially important for communicating this vision to an audience beyond drosophila geneticists. So, although Lewis’s tools were far from trendy, he supplied a way of thinking about genes that those armed with newer techniques in biomolecular manipulation, DNA sequencing, and molecular visualization could easily exploit. It is from this perspective that Lewis helped not only to visualize the logic of molecular development but also to render the logic of molecular development visualizable. One type of meristic variation described by William Bateson ended up being especially important for Ed Lewis’s studies. This was the “homoeotic variation” that we covered in chapter 3. Often referred to today as Hox mutants, these variations changed a character or organ into something that looked like an entirely different character. Through the experiments of Ed Lewis and others, these mutants provided dramatic insights into the modularity of organism construction and how similar modules can become different from one another.

Figure 4.4. Segmentation in larval and adult drosphila. Segments T2 and T3 are the thoracic segments that produce wings and halteres. Scott F. Gilbert, Developmental Biology, 10th ed. (Sunderland, Mass.: Sinauer Associates, Inc., 2013), Figure 6.7. Reprinted with permission from Oxford University Press.

148 · Modulations

Take, for instance, the homeotic variations that Ed Lewis researched for his 1978 Nature paper, the bithorax gene complex. Figure 4.4 is an illustration taken from a well-­known textbook in developmental biology that shows how the segments of a fruit fly larva develop into the segments of an adult. One of the important characteristics of flies as winged arthropods is that they have one, instead of two, pair of wings. For some reason, flies have developed the ability to turn off the expression of one of these pairs of wings. In drosophila, specifically, the wings develop on the T2 thoracic segment as shown in Figure 4.4. Behind these wings, on the T3 segment, are found a pair of winglike disks called “halteres” that are thought to stabilize the flight of the fly while providing little to no actual lift. Mutations in the bithorax gene complex, however, transform the halteres of the T3 segment into winglike appendages. This is how the complex got its name, bithorax complex, as the fly now looks like it has two winged thoracic segments instead of one. A stunning example of this is pictured in Figure 4.5, from the cover of a 1983 issue of Nature. In this case, the T3 segment looks entirely transformed into a T2 segment. The halteres have been changed into wings. If the cooties in the game of Cootie had winglike appendages as well as legs, the haltere piece for the game would be switched with an extra wing piece. The original bithorax mutant, however, wasn’t as visually dramatic as the fly on the cover of Science. The fly with two complete sets of wings had to be constructed by combining specific mutations effecting the halteres. This point is important as it suggests why Lewis found these genes so fruitful to study. First isolated in 1915 by Calvin Bridges,47 the original bithorax mutant, later known as bx (for bithorax), had a T3 segment only partially transformed into a T2 segment. This means that the halteres are like typical halteres in the part of the segment toward the back of the fly, but they seem to be substituted for a true winglike segment in the half of the haltere toward the front end of the fly (Figure 4.6). This mismatch in parts of the fly wings produces a haltere/wing structure sized in between that of wing and haltere and shaped with a graceful curve. In 1919, Bridges isolated another mutation to the halteres which he named bxd (for bithoraxoid). This mutation was the inverse of bx in that the side of the enlarged wing was switched. A mutation in bxd changed the back part of the haltere into a wing while the front part remained a haltere. A third mutation was identified in 1934 when W. F. Holland mailed to Bridges a mutant fly Lewis would later call, ubx (for ultrabithorax).48 Although flies with two copies of the ubx mutation would die as larvae, flies with only one copy of the mutated gene produced slightly enlarged

Modulations · 149

Figure 4.5. A completely transformed mutant of the bithorax complex where mutations have complemented to make a fly with four true wings. Science 221, no. 4605 (1983). Reprinted with permission from AAAS.

halteres. This mutation particularly interested Bridges as it appeared to exist on a different gene than the other mutations but caused similar types of effects when expressed.49 Bridges began referring to this group of mutants as the “bithorax cluster,” since there seemed to be a cluster of closely related genes that caused the mutations.

150 · Modulations

Figure 4.6. A reproduction of Calvin Bridges’s first illustration of bithorax. Notice the halteres transformed into a more winglike structure but not yet looking like complete wings. Calvin B. Bridges and Thomas Hunt Morgan, The Third-­Chromosome Group of Mutant Characters of Drosophila Melanogaster (Washington, D.C.: Carnegie Institution of Washington, 1923), 138.

The theory that genes could, in fact, occur in clusters on the chromosome is interesting in itself. This observation troubled the law of independent assortment of heritable units as outlined by Mendel, where genes were thought to assort to offspring independently. Granted, the chromosomal theory of inheritance already troubled this distinction, as genes on the same chromosome tended to be inherited together, yet in this case, the associated genes appeared to have a similar biochemical function as well. This was an especially interesting complication. To understand why a cluster of genes with similar functions might be an interesting problem, one needs to understand the view of the gene as it was emerging in the early to mid-­twentieth century. Understanding genes at that time involved characterizing their behaviors biochemically and genetically. The quickest way to visualize a change in a gene was through the effects this change made on the phenotype of the organism, through the visual identification of individual mutants. Once identified as a mutant,

Modulations · 151

the relative position of the gene responsible for these changes could be mapped by crossing the mutant fly with other flies. Correlating a specific biochemically induced phenotypic change with a specific genetic location suggested that one had located the change and defined its effects on the organism. The assumption behind this approach was that the gene operated through a simple linear causality, where a single gene discreetly coded for a single enzyme. This assumption of simple genetic causality was best exemplified by the approach of G. W. Beadle and E. L. Tatum when they declared their hypothesis of one gene–­one enzyme, where a single gene coded for a single enzyme responsible for the phenotypic change.50 What Bridges is identifying, however, is anything but simple. First of all, multiple genes appeared to control a single function, haltere formation. This meant that the mutations for haltere formation mapped to different locations on the chromosome. Even more importantly, although these mutations mapped separately, they always appeared to be located near one another. This was a phenomenon that Barbara McClintock described in 1944 and named “pseudoallelism.”51 More surprisingly, Lewis noted that some of these mutations seemed to act synergistically when they occurred near one another. In 1951, he used this observation to refine McClintock’s observations, calling these mutations “position pseudoallelism.”52 So although the genes seemed to work the same way, they were found on slightly different parts of the chromosome and, when found together, seemed to augment one another’s effects. Clearly something was going on that the simple one gene–­one enzyme formula failed to explain. Bridges appealed to cytology, or the visualization of the chromosomes of flies, to understand what this something might be. The salivary glands of the fruit fly, as well as a few other organisms, have chromosomes that have undergone several rounds of replication without any division. All this extra genetic material enlarges the chromosome, making them easier to visualize than chromosomes from other regions of the same organism. Bridges used this technique to demonstrate that pseudoallelic mutants often demonstrated duplicate banding patterns on chromosomes. Bridges argued at the time that these double bands demonstrated gene duplications, suggesting that pseudoalleles might be a single gene now duplicated onto contiguous sites on the chromosome. This was an exciting insight for Bridges as it suggested a possible role that mutation could play in evolution. If one of the newly duplicated genes mutated to provide another function, then the cells carrying this mutation

152 · Modulations

would now have two functions, the mutated function and the function of the original gene. This would allow an organism to acquire a new function without losing the possibly critical function of the original gene. Lewis began using the bithorax mutants in the late 1940s to test this hypothesis.53 In 1951, he published a mechanism for how this might work in a thorough analysis of mutant crosses.54 He described the relationship between these pseudoalleles as a difference in “levels of development.” Lewis initially imagined that development might occur through a series of biochemical steps. Genes would produce products that would transform a substrate into something new and different. The linked genes of the bithorax complex were especially interesting as the genes would successively transform a single substrate. An analogy would be an assembly line, where workers build a single product by sequentially adding different parts. Mutations to genes in the complex would disrupt this sequence at a specific stage. Consequently, a change in a protein created by a mutation early in the process would also affect the proteins created later in the process. Lewis suggested that the levels of development for the bithorax complex might look like this: Substrate → bx → Substance A → ubx → Substance B → bxd → Substance C In this diagram, the gene where the bx mutation occurs is responsible for turning a chemical substrate, known just as Substrate, into a different substance, known as Substance A. The genes where ubx and bxd mutations occur were then responsible for turning Substance A into Substances B and C successively. Lewis believed that these interactions occur because they are localized to a very small portion of the nucleus. Having no understanding that genes are translated from nucleic acid into protein in the cytoplasm and thus a distance from where the gene appears in the nucleus, Lewis imagined that gene products would mostly effect the proteins produced by nearby genes. All the components would then be found in high enough densities to allow for their interaction. The further away the gene, the less gene product would be available as a reactant. In this 1951 paper, Lewis offers a model for how gene products can act in a sequential manner on a single initial substrate to transform that substrate into a series of related but different proteins. Each step of the process was a “level of development” for creating the final product, which he called “Substance C.” At this point in

Modulations · 153

his thinking, changes in expression of gene products weren’t due to any regulation of genes, per se, but to changes in concentrations of biochemical reactions. By his 1978 paper, Lewis had changed his argument believing that proteins produced by genes could directly regulate neighboring genes. This simple conceptual leap involved a much deeper reevaluation of the role of gene expression in development. In his earlier work, Lewis thought it was the proteins that did the important work of creating cellular variation. In the 1978 paper, Lewis envisioned that genes and proteins worked together to create a series of switches where each switch was sensitive to all previous switches. This meant that proteins weren’t just transforming a bit of cellular stuff, they carried information about what the next step in the sequence should be. It also meant that genes weren’t just continually on, they were exquisitely sensitive to some of the proteins that surrounded them. This distinction added a new level of control in genetic regulation as it revealed whole new possibilities for genes to act in different ways in different cellular environments. In effect, Lewis suggested that this type of sequential process not only led to predetermined (or programmed) outcomes, it could also be a mechanism for variation. A protein involved in creating halteres in one thoracic segment could create wings in another segment depending on what other proteins were expressed. The mechanism of regulation became the source for variation. By the time he published his 1978 paper, Lewis had elaborated this simple sequential scheme into a vast and complex picture of how genes might interact in development. During his career, Lewis became known for his meticulous construction of drosophila mutant stocks and his ability to think in abstract, even visual, terms, about how they relate. His colleagues often referred to Lewis’s 1978 paper as a “picture of ideas”55 for the way that Lewis imagined how gene expression could occur: Ed’s preoccupation with making perfectly transformed flies reflected perhaps an attempt to achieve mastery of the genetic system or an esthetic sense for perfection. It also revealed Ed’s appreciation of the power of an image. Before the description of the homeotic mutations, one might have imagined that the network of developmental decisions would be hopelessly complex, that no understanding would come before a complete description. Homeotic mutations made obvious that there is a decision tree

154 · Modulations

and at least some logical generalizations to the program of animal development. This optimism motivated many other geneticists . . . to search for other master regulators, and it underlies the interpretation of the phenotypes they found.56 Lewis’s 1978 paper offered an encompassing vision for how development of an organism could result from many smaller decisions. This vision was best communicated through images. Flusser’s work on technical images is especially useful for understanding the significance of this claim. First of all, the insistence of referring to Lewis’s work on visual terms plays to the overall phenomenological difference between images and lines. Lewis used abstract images to present an overall series of relationships, a “decision tree” as opposed to the strictly linear causal interpretation of Lewis’s 1951 paper. As we saw in chapter 2, Flusser thought that “This space and time peculiar to the image is none other than the world of magic, a world in which every­thing is repeated and in which everything participates in a significant context.”57 This shifted the interpretive emphasis from discerning a single directly linear causality to discerning an emergent series of patterns. As we also saw in chapter 2, Flusser believed technical images were different from pretechnical images in that they incorporated texts, or instructions, into these new formal relationships. This guaranteed that the patterns that were produced referred just as much to the informational and industrial system that produced the patterns as they did to the actual material structures of the fly. This wasn’t a rebuilding of the fly as an organic form, it was a general way of considering how variation could be built from similar mechanisms, essentially the same goals as we discussed in chapters 2 and 3 for twentieth-­century industry. We will cover Lewis’s work as a form of envisioning in much greater detail below. Although he never used the term “modular” in his paper, Lewis’s overall schema treated drosophila development as a modular system in the same terms as we have seen modularity defined. It was the combined autonomy and interdependence of gene products that helped determine the variation of segmentation in the fruit fly. And, as others would argue in the  developing science of evolutionary and developmental biology, the same mechanism could be extrapolated for explaining how a diversity of forms could be produced by similar mechanisms across species.58 Two key illustrations in Lewis’s 1978 paper are especially important for communicating his vision of development (see Figure 4.7 and Figure

Modulations · 155

4.12). A close look at these figures reveals just how radical Lewis’s vision of development was becoming along with his increasing reliance on illustrations to communicate it. The first is a diagram of the stepwise dependency of gene expression of the bithorax gene region (Figure 4.7). This is an expansion and complication of what Lewis calls “levels of development” in his 1951 paper. This figure moves from a stylized realism of the fly into the abstraction of stages of modular assembly in a single illustration. The second illustration (Figure 4.12) moves the other way, suggesting how genetic and chemical signals can work to build an organism. In this figure, Lewis uses his model of development to explain the variation of his mutant flies. The little bit of organicism present in the first illustration drops out in favor of a schematics of modular construction. What emerges from the combination of these two images is an interesting recapitulation of the history of genetics in the twentieth century. The first illustration demonstrates what I call “genetic rationality,” or how a reliance on information processing technologies helped create a vision of heredity as controlled by particulate genes. This moment was an analytical moment of breaking things down into smaller components. The second illustration shows how these particles can work together to create an organism. This moment was a constructivist moment where ideas of grids and modules helped Lewis envision how organisms could be rebuilt. Taken together, these illustrations transform a concept of normal “form” into a radical affirmation that variations are the theme. Let’s begin with part a of the first illustration (see Figure 4.7; part a is found in the upper left corner. I’ve enlarged and labeled it in Figure 4.8). On the right side of the illustration you will recognize a schematically rendered but recognizable adult fruit fly. Lewis places a schematically rendered larval form of the fruit fly on the left side of the insert. He then places a much more abstract depiction of fly segmentation in between these images. Accompanying this depiction are labels and lines that chart how the general pattern of segmentation, as found in the abstract form, expresses itself in the larval and adult forms of the fly. It is difficult to see, but Lewis takes care to label the phenotypic features important for identifying the different mutants of the bithorax complex in the adult and larval flies. In the larval form these include such features as the mandible hooks and dorsal sense organs of the head, the pattern of sensory bristles in the thorax, the patterns of fine spikes and pits found on the underside of the thorax, the more course spiking found on the underside of the abdomen,

156 · Modulations

Figure 4.7. Figure 1 entitled “Genetic Control of Segmental Levels of Development in Thoracic and Abdominal Segments,” E. B. Lewis, “A Gene Complex Controlling Segmentation in Drosophila,” Nature 276 (1978): 565–­70. Image courtesy of Springer Nature Publishing.

and at the very tail of the larva fly, a final posterior spike. These are juxtaposed to the more recognizable features of the adult fruit fly, including legs, wings, and halteres. Lewis then uses this highly abstract version of segmentation to describe how his mutants map levels of development in each of the segments of the fly. As I have illustrated with arrows in Figure 4.8, Lewis begins by mapping two different axes in the developing fly: the creation of distinct modules arranged longitudinally along the fly’s body (what Carroll referred to as making “a series of initially similar modules”) and then demonstrating how these modules develop during the morphogenesis of the fly (what Carroll called “making [the modules] different from one another”). The larger part of the illustration provided in Figure 4.7 (listed as part b in the original paper) adds another level of complexity to the illustration by mapping how mutations of the bithorax gene complex effect the level of development of phenotypic features for each of the segments. The key to understanding this added level of complexity is the small insert at the top right of Figure 4.7, excerpted as Figure 4.9 below.

Segment Differentiation

Stage of Development

Figure 4.8. Detail from upper left corner of Figure 4.7. There are two axes of differentiation in the illustration. Moving from the top to bottom in the illustration demonstrates how the flies are similarly segmented in the adult and larva state. Moving from left to right demonstrates how the phenotypic effects of genetic expression in these segments change as the fly matures from larva to adulthood. Figure 4.9. Detail of insert on upper right of part b in Figure 4.7. This illustration abstracts an overall expression pattern for each thoracic segment as a series of levels of development of expression. It is offered as a key for making sense of the complex diagram that dominates Figure 4.7.

Sequence of Gene Expression

158 · Modulations

This insert shows that each segment in the thorax has four discrete levels of development that correspond with each of the four genes found in the bithorax complex (these are listed as S0, S3, S4, and S5). Consequently, the expression of Ubx+ (S0) is seen as initiating the expression of the bxd+ gene (S3), while the bxd+ initiates the iab-­2+ gene (S4), which initiates the iab-­3+ gene (S5). The expression of the genes occurs in a stepwise fashion where the expression of each gene (except for the initial Ubx+ gene) requires the expression of the gene immediately preceding it on the chromosome. Remarkably, the sequence of expression of these four genes is the same as their position on the chromosome. This phenomenon, known as “colinearity,” is not shared with other regulatory genes. In the center of part b of the illustration in Figure 4.7 and moving from top to bottom is the abstract rendering of drosophila segmentation shown in part a. Each segment is now drawn with the small circular indicators of levels of development as found in the key presented in Figure 4.9. Each level of development is indicated in these diagrams by shading the circles in the segments, as shown in Figure 4.10. The darkened circles indicate that the segment strongly stimulates the next level of development while the blank circle suggests that no production of the next level exists in that segment. Lewis also included hatched circles to suggest intermediate levels of production. In a wild type fly, the level of expression of the bithorax complex genes increases as one moves from the head toward the tail on the thoracic and abdominal segments of the developing fly. In other words, in the segments labeled MS (mesothorax), MT (metathorax), AB1 (abdominal 1), AB2, and AB3, a wild type fly expresses the sequence of the genes in the gene complex in the same sequence that segments are found while moving from head to tail. The MT segment, for instance, only expresses the product of the Ubx+ gene while the abdominal segment closest to the tail, the AB3 segment, has expressed products of all four genes. The important lesson from this part of the illustration is that new products are produced when moving from anterior to posterior of the fly. This is where the modularity of the arrangement becomes especially important. The development of each segment depends on its own internal level of development to complete the needed sequence of expression as well as the overall position of the segment in the organism. Lewis then supplements this already complex picture of development with an even more complicated horizontal axis that maps how each of his mutants can be explained because of the segmental stepwise expression of levels of development. Abdominal segment 1 (AB1) in Figure 4.11 is a great

Modulations · 159

Levels of Development Increases With Each Segment

Figure 4.10. A detail from part b of Figure 4.7. This illustration details how the level of expression works in most flies. Beginning with the thoracic segment and continuing through the abdominal segments of the fly, each segment displays the expression of a higher level of development.

example of how this works. This example shows the same causal sequence outlined in his 1951 paper (with the addition of the iab mutants) applied to each specific segment of the fly. The overall effect is a highly abstracted grid of how his mutant stocks relate to one another in their levels of development as mapped across an abstract depiction of fly segments. The horizontal axis maps the relationship of levels of development in specific mutants, while the vertical axis suggests how the levels of one segment inform levels in the anterior segment. This illustration makes several interesting points, and I want to emphasize two of these. The first point is epistemological. Lewis has moved from

160 · Modulations

Levels Of Expression Shown By Each Mutant

Figure 4.11. The level of expression of thoracic and abdominal segments as expressed by Lewis’s mutants. Each gene is graded for level of expression of phenotypic traits, where the solid circles represent the highest level of expression (the most like the wild type fly in this case), the hatched lines are lower levels, and an empty circle suggests no expression at all.

a representation of an identifiable fly to an abstract grid depicting how fly segmentation occurs. This grounds his highly abstract vision of development in the recognizable physiological structures of flies.59 The same relationship between concrete and abstract existed in Lewis’s experiments as well. Lewis mapped his genes by following concrete phenotypic changes in his mutant flies. Lewis needs his highly abstract schema to envision how these concrete changes are related to one another. Lewis frequently mentioned the importance of abstraction for his style of reasoning. He credits this insight to reading Bertrand Russell’s, The Scientific Enterprise.60 The second point is formal. Lewis treats each segment of the developing fly as a module with loose constraining relationships to other segments,61 where the overall position of the segment suggests how that segment should develop, and also to internal autonomy, where the levels of development of each segment depend upon the sequence of the expression of genes in that segment. This allows Lewis to provide what might be called an exploded view of a developing organism that emphasizes how each of the different modules can be plugged into development at different times. Although Lewis is thinking of these genes as “master control genes,” or genes that control the expression of other genes, he is, in effect, shifting toward thinking about development in terms of the modular construction of genetic circuits. This is a profound shift in emphasis compared to how he was thinking about gene expression in his 1951 paper. In his earlier paper, genes produce enzymes that directly modify proteins. To return to our earlier analogy of the assembly line, in the 1951 paper, gene products acted as workers on an assembly line where each station on the line would perform a function to transform the parts into a final product. In the 1978 version

Modulations · 161

of the paper, the functions of these workers could be swapped, allowing different products to be created in different circumstances. For instance, one could swap a heater for a light and change how the circuit functions, even though the rest of the components of the circuit remained the same. The implications of this modular view of organisms is brought home in the second major illustration of Lewis’s paper, reproduced as Figure 4.12. In this illustration, Lewis takes the exploded view of the fruit fly and rearranges the modules to show how the Hox genes helped to create several different phenotypic outcomes during development. Across the top of the illustration, each of the phenotypes is listed by its genomic designation and labeled a through h, where h is wild type. At first glance, the stylized fly embryos look similar in design to the wild type embryo in the first diagram, Figure 4.7, with the genotypes listed as levels of developments in each segment. A closer look reveals, however, that Lewis has added a new level of detail in the illustration by schematically indicating phenotypic structures such as the appearance and location of sensory bristles, pigmentation, and spiracles for gas exchange in addition to information about wing, haltere, and limb development. The fly embryos are arranged in the grid moving from loss of all functions on the left to wild type on the right. Lewis is building the fly back up from the highly abstracted state as depicted in part b of Figure 4.7 by showing how the different modules relate to one another when plugged into the developing flies. The fly he builds, however, no longer resembles an organically integrated organism (as found in Figure 4.8); it now looks like a series of modularly constructed fly larva. Taken together, these two illustrations (Figure 4.7 and Figure 4.12) present a comprehensive view of development where organisms are created through the variation of the modules that comprise them. These modules are made from similar processes and can even use similar rules for their expression, but small changes of when and where they are expressed can create large changes in phenotypic outcomes. In a remarkably understated passage even by the standards of technical papers, Lewis points to this conclusion with a simple stark sentence: “During ontogeny the above rules presumably result in each segment having a specific array of BX-­C substances, at the right time, at the right place.”62 This emphasis on building a fly back from an abstract model reflects the fine structure of Lewis’s experimental strategy as well. Most geneticists start with a wild type or “normal” fly and then map single mutations of the wild type through crosses. The idea in these cases is to use deviations from the normal to map normal functions. Lewis pioneered a technique that

162 · Modulations

Figure 4.12. Figure 6 originally entitled “Stylised Phenotypes Detectable in First Instar Larvae or Mature Embryos.” E. B. Lewis, “A Gene Complex Controlling Segmentation in Drosophila,” Nature 276 (1978): 570. Larva a in the illustration is the loss of function mutant for the whole bithorax complex, DfP9, as indicated by the blank circles denoting the larva’s level of development. Larva h is a wild type with complete expression of the bithorax complex. Larvae b–­g are mutant variations of expression of bithorax complex genes. Image courtesy of Springer Nature Publishing.

was the opposite. He established a loss of function mutant for the whole bithorax complex (larva a in Figure 4.12) and began adding wild type genes back into the organism. This allowed Lewis to pinpoint specific gene functions when a specific segment of the gene was added back. Commentators have suggested that Lewis was performing “add-­back genetics,”63 where the goal was to observe genetic effects by recreating the bithorax complex through genetic crosses. Vilém Flusser’s musings on the role of envisioning in the creation of technical images wonderfully explains the goals and products of developmental genetics as it was practiced by Lewis. Creating a technical image wasn’t so much about placing a figure on a surface in relationship to another figure, it involved arranging particles on surfaces in informative ways. Flusser called this process envisioning, where the manipulator

Modulations · 163

of the image had the “capacity to step back from the particle universe into the concrete.” He distinguished envisioning from imagining as envisioners worked with data rendered into bits, particles, texts, and codes while imaginers worked with analog forms and images. “The gesture of the envisioner is directed from a particle toward a surface that never can be achieved [because of its abstraction], whereas that of the traditional image maker is directed from the world of objects toward an actual surface.”64 This means that technical images inform differently from traditional images. Traditional images inform through their reference to objects; technical images inform by creating unlikely configurations of particles. “Envisioners press buttons to inform, in the strictest sense of that word, namely, to make something improbable out of possibilities. They press buttons to seduce the automatic apparatus into making something that is improbable within its program.” Workers on technical images, juxtapose different data streams to create images that couldn’t traditionally occur to better inform themselves how these data streams relate. It is instructive to view molecular development as a form of envisioning. Geneticists during the early twentieth century worked hard to standardize, rationalize, and then atomize a theory of heredity into the science of genetics. This work involved turning concrete bodies into manipulable abstractions and then mapping these abstractions in relationship to one another. Lewis was one of the first geneticists to develop a strategy for thinking about how to move the other way, from abstract genetic designations to concrete bodies. He did this by envisioning how the abstract data gathered from his bithorax mutants could be arranged to create a coherent theory of development. This involved using unlikely combinations of data (his mutants) as a means for achieving the very improbable series of correct steps for creating a normal fly. One of the most astounding implications of Lewis’s envisioning is how it transforms the concept of “normal.” Most people assume that a normal genotype exists as a predominant type in a population. Mutations, then, occur as deviations from this norm. At the level of practice, however, genetics rarely operates with such a simple notion of normal. In fact, even the most cursory history of genetics reveals that there could be no conception of a normal gene without a conception of mutation (or some other form of variance).65 Genes have always been a type of abstraction that only becomes concrete through their ability to correlate different experimental results. A normal individual then emerges as an abstract and unlikely but never fully concrete entity. For instance, Lewis can only envision how a

164 · Modulations

normal bithorax gene complex works by mapping out the possible permutations of mutants that disrupt this process. A normal individual then is the one abstraction that helps make sense of the other abstractions. A normal genotype isn’t an actual physical presence as much as it is a “vision” for how all variations might relate to one another. Perhaps this explains why Bateson was such an important touchstone for Lewis.66 It was Bateson who supplied Lewis with a model for thinking about the role of variation in development. For instance, Lewis often quoted this observation from Bateson’s Materials: “So long as systematic experiments in breeding are wanting, and so long as the attention of naturalists is limited to the study of normal forms, in this part of biology which is perhaps of greater theoretical and even practical importance than any other, there can be no progress.”67 For Bateson, understanding how exceptional cases occur could reveal the multitude of ways that evolution and development might operate. Ed Lewis begins with this same assumption for studying how species evolved. What he soon realizes is that discontinuous changes might suggest a way to begin mapping the intricacies of development as well. Despite the difference in the types of biology in which both researchers engaged, they both shared the conviction that variations are not just anomalies in forms. The variations suggest a means for understanding ordering principles. This ordering principle in Bateson’s hands was the productive capacity of variation. For Ed Lewis it would be a sophisticated plan for how genes can initiate the expression of other genes. Or as Lewis more dryly concluded in his 1979 paper: “The BX-­C genes are assumed to control the organism’s thoracic and abdominal segmentation pattern by producing substrates which in turn regulate other genes that determine segmental structure and function.”68 Envisioning development depends upon the concrete existence of anomalies more than on the existence of an ideal normal type. Figure 4.12 is also an informative illustration for how different Lewis’s formal conception of an organism was from that of a morphologist like Haeckel. Where Haeckel imagined a dialectical relationship of a different type of energy directing forms into elaborate organic structures, Lewis envisioned how particles (genes) could regulate other particles (genes) to create variations. Perhaps a musical comparison might help clarify these very different conceptions of form. The musical analogy for Haeckel’s theory of development would be the symphony. Here the overall form of the work is important, and the parts, such as the phrases and their development, contribute to that form and can build dramatic tension. The more appropriate

Modulations · 165

musical analogy for Lewis’s conception of development is a minimalist set of shifting variations searching for a regulatory theme. Here the interest of the listener is held by the unfolding patterns of tones over shifting time signatures and not by the drama of closure and release. Haeckel’s development plays an organism like a Beethoven symphony while Lewis’s development plays an organism like Philip Glass’s “Music in 12 Parts.”69 Lewis’s modular vision of development appealed to scientists for several reasons. First, his vision offered a way for geneticists to view how the product of one gene could work directly on other genes as either positive or negative modifiers of gene expression. Second, the genes of the bithorax complex were colinear to the order of expression. The sequence genes were expressed in during development was in the same order that genes were found lying on the chromosome (as depicted in Figure 4.9). Although colinear expression wasn’t preserved in other genomic regulatory systems, the fact that the bithorax system was co-­linear provided a straightforward component to an already complex model. The third reason is that Lewis’s vision provided an experimental strategy whereby the development of drosophila segmentation could be studied by characterizing mutations. Nüsslein-­Volhard and Eric Wieschaus explored this strategy for their landmark 1980 paper, “Mutations Affecting Segment Number and Polarity in Drosophila.”70 Lewis, Nüsslein-­Volhard, and Wieschaus shared the 1995 Nobel Prize in Physiology or Medicine “for their discoveries concerning the genetic control of early embryonic development.”71 The fourth and final reason was that the paper gave those trained with the newly minted molecular techniques of sequencing and recombinant DNA analysis an experimental system on which to focus. In 1984, W. McGinnis, et al. published the sequence for a regulatory domain, or a specific sequence of DNA responsible for regulating the expression of genes, called the homeobox of which the bithorax genes are members.72 Other researchers went on to show that homeobox genes were found in animals, plants, and fungi. So even organisms with very different body plans utilize similar molecular mechanisms during development. Modules might not just be organism specific, then; all organisms now appeared to be made from similar components. This is a strong conceptual shift away from thinking in terms of the dialectic of wholes and parts. Too often, when we think of “building things back up” in the history of biology, we tend to appeal to the organicist logic of Kant, Goethe, and Haeckel (among others), where form is derived from

166 · Modulations

the dictates of the whole organism. Modular thinking explains how variation can be built from the bottom up, where small variations in the sequence of expression can lead to large physiological outcomes. As we saw, this is what Carroll referred to as a “logic of making a series of initially similar modules and then making them different from one another.”73 At the end of his career, Lewis fully understood how this modular view of life organized more than the development of drosophila; it provided a way to think of how form could emerge from process in all animals. Take for instance, this poster that Lewis hung on the wall of his laboratory. Made from a picture of a drosophila and an inexpensive paper human skeleton used for Halloween decoration, Lewis then mapped the expression of corrolated regulatory gene systems from head to tail on both images (see Figure 4.13). The similar patterns of labels across the torso of the two organisms stress how viewing animals as modules helped Lewis and other researchers to envision how the diversity of animal forms could be derived from a similar process. It all depended on how the process was regulated.

Figure 4.13. Photograph of the working model of Hox organization found on Ed Lewis’s laboratory wall. See also Figure 1 in James F. Crow and Welcome Bender, “Edward B. Lewis, 1918–­2004,” Genetics 168, no. 4 (2004): 1774. Photograph from Ed Lewis Papers, permission courtesy of the Archives, California Institute of Technology.

5

Drawing Together Composite Lives and Liquid Regulations

It might have been the most important lecture he gave in his life; it certainly had the highest profile. In 1995, Edward B. Lewis shared the Nobel Prize in Medicine with Christiane Nüsslein-­Volhard and Eric Wieschaus, for “their discoveries concerning the genetic control of early embryonic development.”1 Recipients of the Nobel Prize are required to present their research a few days prior to receiving their award. These lectures are the usual length of a plenary presentation, around fifty minutes. Ed Lewis’s lecture was entitled “The Bithorax Complex: The First 50 Years,” and he told the history of research on the bithorax complex and the implications of this research for understanding development in drosophila. Remarkably, Lewis played a homemade animated movie for over twenty-­two minutes of his forty-­seven-­minute address. In the movie, Lewis combined photographic images of bithorax mutants with cutout animations from colored paper templates. Taking advantage of Caltech’s proximity to the Los Angeles film industry, Lewis got advice from a professional animator2 and chose to animate his complex model of development to help others better understand it. Animating his model allowed Lewis at least three specific improvements: he depicted his complex model of development as a moving image, he easily shifted between scenes incorporating actual mutant flies with animations of developmental models, and he created composite images where images could be overlaid with other images or text. These three improvements allowed Lewis to depict the role of molecular signaling during development in interesting and engaging ways not easily possible in a static illustration or a written description. Lewis could now use animated images to rapidly shift scales between organisms and molecules, or he could inter­ weave photorealistic images with abstract models in order to suggest how something might occur. He could also layer images to strategically 167

168 · Drawing Together

combine the specificity of labels and diagrams with the associative power of images. One technique especially gives animation the flexibility to incorporate all three of these capabilities. It is called “compositing” and its flexibility makes it an especially important filmic technique for the twenty-­first-­ century entertainment industry. A composite image is any image where a part of one image has been replaced with a part of an image from a different source. In practice, composite images are created in several different ways. They can be created through purely physical means, where actors on a soundstage act in front of either a projected or painted background. They also can be created through purely electronic means, where one computerized image is layered over another. Or they can be used in hybrid images with the use of green screen or chroma key techniques, where actors are filmed over backgrounds of a single color (like a screen of green) and special filters can then remove these colors from the film. This allows directors to place actors in front of fantastical backgrounds during the film’s postproduction. The important point is that many contemporary images are composited images. Animation, movie special effects, even the more mundane layering of texts on top of images (think here of the “ticker” or “chyron” of a newscast scrolling at the bottom of the main image) are all composited images. As Jeff Sauer recently wrote for Computer Graphics World, “Compositing has exploded, driven primarily by the rapid growth of computer processing power, and it has become a regular, indeed expected, part of postproduction.”3 Despite the film industry’s enthusiastic use of computers for compositing, the use of composited images predates the introduction of computers in film production. Set design in theater, for instance, could be considered a form of a composite image as the actors depict a scene in front of a background. Yet, it is with the introduction of movie cameras that we get a proliferation of various ways of using compositing to build different types of worlds. Backgrounds in films were not only painted, and thus static, they could be projected as well. The composite image was the combined image of two different movies. This allowed for the film director to not only depict two different types of worlds (the world of the actors and the world of the background) but also two different sources of movement (the movement of the actors and the movement of items in the background). This little trick enabled a new universe of methods to show various rates of change in a single moving image. Compositing not only stimulated an

Drawing Together · 169

industry of movie special effects, it also provided novel ways of depicting change over time in biological moving images. As we saw in chapter 1, it is hard to depict how living things change with a single image on a page. This is why scientists often used specific forms to suggest movement, such as Haeckel’s use of the open curved line in the Kunst-­Formen der Natur, or why they used a sequence of images to suggest change over time in a single organism, such as Haeckel used in his grids in Anthropogenie. These techniques weren’t just interesting flourishes in biological illustrations, they made important biological arguments. As Jacques Monod remarked, biology more than most other sciences is especially keen on understanding how change occurs over time: “In contrast to most aspects of physics, biology incorporates time as one of its essential parameters.”4 It’s not so much that other sciences don’t profitably utilize time as a construct, it’s just that the history of biology is marked by a need to understand how change occurs. Many debates in studies in heredity, speciation, metabolism, morphology, ecology, and development were shaped by how researchers thought that change might happen. As medical illustrator Eric Hazen suggests, “Visualizing change over time is essential to understanding biological events such as embryogenesis and development.”5 Finding the best way to envision changes over time, though, often depended on adapting the visualization technologies that scientists had readily available. For Ed Lewis working in the pre–­personal computing era, this meant utilizing animation’s unique ability to depict the passage of time. Put in the terms of regulation explored in this book, compositing is a marvelous, albeit limited, tool for ordering durations of time.6 Yet, as we already indicated, compositing gives animators other benefits than the depiction of rates of biological change. Drawn images can be layered over photographic images and text can be layered on pictures. It is in my emphasis on compositing that this chapter may depart from common sense ideas on what animation is or how it achieves its goals. For many media theorists, animation is primarily a way to put inanimate objects into motion. Animation achieves its “illusion of life” by animating what might seem like a previously inanimate media environment.7 Critical theory has even begun to embrace the idea of animation to visually describe a world constant in its restlessness and vitality. And although putting subjects and objects into motion is an important part of what happens in animation, motion is the effect, not the enabling condition that makes animation such a ubiquitous part of today’s media diet. In this belief, I follow Thomas

170 · Drawing Together

Lamarre as he argues in his book, The Anime Machine: A Media Theory of Animation, “In the analysis of animation, priority should fall on compositing (the space within images that becomes spread across frames) over character animation (movement across frames).”8 Lamarre’s key insight is that the depiction of movement in animated images demands layering images on top of one another to depict change. Animation can depict movement, then, because of its facility in creating composite images. The depiction of change over time is an effect of drawing together multiple durations of past, present, and futures in a moving image. When a character kicks a ball in an animated film, the animator carefully depicts changes in the rate of the ball’s movement in relationship to the character’s leg and the movement in the background. The world doesn’t exist as a static state that is put into motion; it exists as a composite experience of all the different rates of change drawn together in a single image. This key insight, that animation depicts multiple rates of change, suggests to me that animation is a complex tool for regulating flows. It can fruitfully be viewed as akin to other ways of understanding flows, such as those elaborated in the study of fluid dynamics or those applied to the distribution of goods in a business. Although we often think in terms of a world composed of subjects and objects, this is a product of our scale of the phenomenological experience of our world. In order to understand how the very small, the very swift, or the extremely large affect our experience of the world, it is often useful to move from the register of seeing how objects are placed on a grid in order to see how flows of objects or particles constitute different types of grids at different times. So, although animation may initially appeal to biologists for its promise to put ideas and images into motion, it ends up being particularly useful as it works in a sophisticated way of building things like grids back up, by offering differential regulation of how change occurs over time. This chapter will use Ed Lewis’s decidedly low-tech animations as an analytical fulcrum from which to pry open new understandings of how biologists use animation to depict a constantly changing world. My reliance on the low tech might also strike some readers as a departure from common sense. Viewers today can find amazingly beautiful and extremely detailed animated movies of cellular processes freely available online. These films use sophisticated animation software, such as Alias Systems’ Maya, to produce stunning visual effects. Relying too heavily on computerized techniques as a form of analysis of scientific images can obscure

Drawing Together · 171

as much as they reveal, however. For instance, one of the most important developments in the history of twentieth-­century animation is the role of labor in producing moving images. Putting a world into motion requires hard work. Skipping directly to a world of heavily computerized animation obscures how difficult it was to adopt animation to computing environments in the first place.9 Interestingly, understanding the innovations used to save labor in producing celluloid or filmed animation can, if done carefully, increase an understanding of how software and computers create animated content. Many of the conceptual innovations of cel animation can still be found as common operating heuristics in computerized animation programs. An analysis of Lewis’s movie also gives me a perspective from which to create my own composite analysis of how bodies were made calculable before they could be computed. Lewis is an interesting case in this regard as he assumed that teaching himself animation was more important for his science than teaching himself techniques in computation. Lewis saw animation as a viable alternative for putting ideas into motion before computing. Granted, much of this choice probably had to do with the techniques available to him, but this is exactly the point. His location near the Los Angeles film industry is important as well. As many other researchers have argued, visualization techniques in biology are often bound to visuali­ zation techniques in the entertainment industry.10 This reminds us of an incredible diversity of resources for biologists in mid-­twentieth-­century biology. Change in biological conceptions wasn’t just the product of the introduction of computing to biology, and animation is a great example for understanding this. Contrary to fields such as molecular modeling, where computers were embraced relatively early in its history, animation implicates a different model of scientific development, the development of the film and entertainment industry. As molecular animator Janet Iwasa commented, “Unlike molecular visualization software, like Chimera and Pymol, 3D animation applications were not designed to display or analyze molecular structures. Rather, most animation software was created for the game and entertainment industries, and the software capabilities reflect the needs of this target audience.”11 This chapter’s focus on compositing is especially meant to demonstrate how much of our ability to visualize animated biological processes is due to visualization of spectacles in our filmic entertainments. As Flusser has argued for all forms of envisioning with technical images, distinctions between fact and fiction often recede

172 · Drawing Together

in importance to how different types of data are assembled, or in this case, composited. To state it provocatively, our envisioning of the workings of twenty-­first-­century bodies depended on the development of an industry for producing special effects. Debates shift away from the rationalistic (“are these animations real?”) to the pragmatic ( “are these animations useful?”). Finally, and perhaps most importantly, I have already argued that there is a politics to how forms are used to depict living things; focusing on animation, however, deepens these arguments by suggesting that there is also a politics to how these forms are composited to create images with depth and movement. As we will see, although animators in the biological sciences are concerned with the use of specific forms, they are even more interested in combining different forms to depict specific outcomes. Biologists must continually ask themselves if what is being animated, and how it is being animated, helps them understand other sources of data in new ways. Thomas Lamarre has argued for the importance of understanding the multiple effects that composited images can produce. The animation stand provides a number of ways to deal with the gaps between layers of the image. It allows for techniques of compositing that help to suppress the sense of a gap between layers, because movement within the image might undermine the sense of the stability of the image or of the continuity of movement across images. I will refer to this suppression of the perception of movement between layers as closed compositing. But there are other uses of the animation stand. It is also possible to composite layers of the image very loosely (open compositing), which imparts the sense of a truly multiplanar image. There is also “flat compositing,” in which the play of layers remains palpable but comes to the surface of the image, which I will later call the superplanar image.12 These forms of compositing, “open,” “closed,” and “flat” provide a very careful reminder that not all forms of holding images together need to be  considered as the consequence of a single type of “gaze,” although there are clearly times when this is at play.13 For biologists, this means that the goal of using animation shouldn’t always be closed compositing, where all potential gaps between images are smoothed into flashy special effects cele­brations. In fact, it is only the rare biological event that smooths molecular dynamics, cellular dynamics, and organismic dynamics into a

Drawing Together · 173

single logic of biological change.14 Most of the time, these different types of series of events bend and warp each other in interesting ways. This brings us to the political point about the use of animation to understand the regulation of all types of bodies. As I argued in our investigation of pests in chapter 3, understanding the use of grids in biology often involves understanding how grids interact, or how different forms of ordering meet and inform one another. Every time two different orders of change meet they create localized pockets that don’t fit squarely into either order of change. These localized interactions create what I’ve called “pests.” The political value of understanding how pests are produced, is twofold. First of all, it is important to recognize that not all meetings between things, peoples, or regulations need to be reduced or synthesized into one another. This was the problem of organic holism, where the synthesis of the parts was thought to create a well-­ordered organism. Most of the time it is important to hold different logics, patterns, and rates of change together without the assimilation of one into the other. Second of all, these places/moments of warping and change can be incredibly fruitful places for generating new types of lives. They breed dynamics not found in either of the original orders, hence their pestiness. This productivity isn’t inherently “good,” “bad,” “light,” or “dark” but arises through differential forms of combination that need to be analyzed in situ to more fully understand their consequences. Realizing how life is always a form of composite life, where elements are sometimes firmly bound and sometimes loosely associated, is key for understanding the dynamics of regulation in relationship to living things. Noticing the moments when a sense of order is warped, haunted, or strained is an especially sensitive indicator of the depth, complexity, and vitality of life in the grid.15

The Return of the Fly: Lewis’s Movie The movie Ed Lewis shared at his Nobel lecture consisted of six short handmade scenes. Lewis crafted each scene to make a specific point about the evolution and development of drosophila. In order to do so, however, he needed to mix a surprising number of cinematic techniques. Lewis uses live action images of drosophila to establish what developmental anomalies of the fly look like, and he animated some scenes to explain his complex models for how molecular interactions regulated genic expression. A brief overview of the scenes of the movie gives a sense of how Lewis used certain

174 · Drawing Together

types of cinematic conventions to make scientific points. The first scene is entitled “Life Cycle of Drosophila (abridged)” (see Figure 5.1) and is filmed using living flies under a dissecting scope. Using a voiceover reminiscent of classics in the natural history genre, such as Jean Painlevé’s documentaries,16 Lewis narrates the life cycle of the fly beginning with gestating eggs mounted in mineral oil and ending with adult flies. During this narration, Lewis takes time to point out developing structures in the larvae such as the “black jaw hooks used to burrow into food and the posterior spiracle used for breathing.” Each of these features will be important in his later more abstract depictions of the development of the fly. Lewis then shifts to describing the adult structures of the fly and ends with a powerful shot of a gravid female fly laying an egg. He observes at the end of this scene, “As this sequence ends, the life cycle starts over again.” The cyclical structure of the narrative and the parenthetically modified title “(abridged)” point to the shortcomings of filming flies, where the camera can only capture the briefest episodes in the drama of even the briefest of lives. It is appropriate that Lewis turns to animation to depict the unfilmable durations of the evolutionary history of the fruit fly in a scene entitled “Evolution of the Fly.” Lewis begins this sequence with a progenitor arthropod, a relatively undifferentiated “worm like organism.” The use of animation allows him to suggest the continuity of an evolutionary pathway in a few minutes of film. He then animates how this organism evolves as it develops limbs, a proboscis, and antennae. In addition to showing how flies might have evolved, Lewis uses this scene to develop a visual iconography for other animations sprinkled throughout the film. Establishing these visual conventions allows viewers to more easily move between depictions of actual physiological features and Lewis’s stylized animated depictions. The highly condensed and decidedly guided depiction of evolution lends this scene the didactic quality of an educational film (see Figure 5.2). If the first two scenes play like documentary and educational films, the third scene, “Mutants of the Bithorax Gene Complex,” is best described as a horror film. This description is especially apt as the subjects of the scene are Lewis’s collection of drosophila homeomutants, or mutants where whole body parts are missing or misplaced. As we saw in chapter 3, this type of mutation fueled stories on the horrors of misregulation in other forms of entertainments for centuries. Lewis augments this theater of the

A

B

C

Figure 5.1. Stills from Scene 1, “Life Cycle of Drosophila (abridged),” from Ed Lewis’s Nobel Prize lecture movie. A. A drosophila larva, B. a close-­up of an adult male drosophila, and C. a female drosophila laying an egg. Courtesy of the Archives, California Institute of Technology.

Figure 5.2. Stills from Scene 2, the “Evolution of the Fly,” from Ed Lewis’s Nobel Prize lecture movie where Lewis animates the evolutionary history of the fly through different stages of invertebrate evolution. Courtesy of the Archives, California Institute of Technology.

176 · Drawing Together

grotesque by filming live mutants that are sometimes fixed to the glass slide. We will cover this scene in greater detail below. Lewis returns to his animation stand to create the last three scenes of his movie. All three scenes illustrate how mutations in the bithorax complex explain the genetics of molecular development. The scene “Late Effects of the Bithorax Gene Complex” shows how mutations to the bithorax complex affect the anatomy of adult flies, such as the shift from a pair of halteres (small wing-­like stabilizers) to a second pair of wings. The next scene, entitled “Early Effects of the Bithorax Gene Complex,” animates the hard to visualize changes in the cuticle and spiracle of the developing larva. In his concluding scene, the “Regulation of the Bithorax Gene Complex (a model)” Lewis uses animated effects, such as changes in scale within the frame, to show how his model of development explains how molecular interactions change physiological features. I will cover this last scene in more detail below as well. Although it isn’t the most riveting twenty minutes of film you will ever watch, and the camera and animation techniques look amateurish to screen savvy twenty-­first-­century viewers, Lewis’s choice of animation as a medium reveals important assumptions about how best to understand molecular genetic regulation. First of all, animation only enhances the importance of modularity when thinking about evolution and development. Lewis’s use of paper cutouts to create organisms from standardized shapes lends to the feeling that these shapes can easily be swapped out for other shapes, as this is basically what Lewis does when he makes his film. It is also important to understand how much Lewis wanted to animate his model. This is not the typical career choice of a researcher of Lewis’s stature. He took time to train himself in a new technique and performed all the hands-­on work himself. Animation even became an integral part of his later scientific career, as many visitors to Lewis’s lab, for instance, commented on seeing his animation stand set up in a corner of his laboratory.17 As we have seen, Lewis’s model for genetic regulation was particularly complex, and the control that animation offered to help depict these changes must have been especially attractive. One of animation’s most important qualities, I will argue below, is that its use of composite images allowed Lewis to navigate between changes in representational practices, molecular and organismic scales, and durations of change. Understanding exactly how animation provides control, however, depends on understanding how animation differs from other types of filmmaking.

Drawing Together · 177

The Ontology of the Technical Image: From Representation to Projection One of the most dramatic effects Ed Lewis used in the construction of his film was the shift between photorealism, where flies were directly filmed with a movie camera, and animation, where he used cutouts and compositing to create scenes that were either difficult or impossible to photograph. Lewis tended to use photorealism when he wanted to document a phenomenon, such as the life cycle of the fly or the phenotypic effects of specific mutations, and he used animation when he wanted to present his model of regulation or when he was working at a scale that couldn’t be photographed. So for Lewis, each of these photographic styles had specific epistemological values and each of these values was important for scientific practice. He used photography to enhance the empirical value of the film while he used animation to put abstract models into motion. This isn’t necessarily a straightforward choice. Lewis wasn’t using photography as a primary source of data. Very few pictures of flies, for instance, survive in Lewis’s archives. He could have easily animated all the sequences of the film if he wanted to. Instead, Lewis thought it important that we see what he sees when he looks at a fly under a dissecting scope. There was something about photographing a fly that appealed to Lewis for this specific purpose. Photography has long been used to present the world as it is thought to be. André Bazin, perhaps, expresses this most elegantly in his influential essay on “The Ontology of the Photographic Image.” According to Bazin, “The objective nature of photography confers on it a quality of credibility absent from all other picture making. In spite of any objections our critical spirit may offer, we are forced to accept as real the existence of the object reproduced, actually ‘re-­presented, set before US’ that is to say, in time and space. Photography enjoys a certain advantage in virtue of this transference of reality from the thing to its reproduction.”18 According to Bazin, photography’s special qualities privilege it for capturing and representing the real existence outside of the camera. Because of this, photography is more aesthetically suited to a form of true realism, where “the aesthetic qualities of photography are to be sought in its power to lay bare the realities.” As “[o]nly the impassive lens, stripping its object of all those ways of seeing it, those piled up preconceptions, that spiritual dust and grime with which my eyes have covered it, are able to present [reality] in all its virginal purity to my attention and consequently to my love.” For Bazin

178 · Drawing Together

then, photography was a remarkably creative act that helped to present a world as real and thus reveal how nature “imitates the artist.” Photography creates its sense of realism through its special mechanical properties in capturing the world. In this it is different from other forms of creativity. Other thinkers see this authority as bought at the expense of a relationship to time and space. The photograph only participates in its own depiction of the world as a world that has already occurred. This is what Bernard Stiegler grasps at when he suggests that photography is a “mirror reflecting the past.” The mirror of photography is good at capturing “what happened” but since this is always something that has just passed, this reflection is always an “adieu” to what has just occurred.19 The photographic surface gains its credibility as a form of realism by removing the world that it pictures from the depth of space and its capacity to change over time. These reflections on the ontology of the role of the photographic image add depth and complexity to Flusser’s portrayal of envisioning technical images examined in earlier chapters. Flusser also saw that photographs were more than simple mirrors of the world; this is especially evident in his distinction between traditional and technical images. “Traditional images  are mirrors. They capture the vectors of meaning that move from the world toward us, code them differently, and reflect them, recoded in this way, on a surface.” He then contrasts: “Technical images are projections. They capture meaningless signs that come to us from the world (photons, electrons) and code them to give them a meaning.”20 Photographs intended to provide a sense of realism of the world then, are envisioned in such a way as to present the image as real. The meaning they provide, however, is anything but quotidian. As we explored in relationship to the use of anomalies in genetic research in the last chapter, especially important technical images are often those that present improbable events. This is why mutants are so useful (especially when envisioned in a realistic fashion). They signify that despite the reality of the image, some aspects of reality are at first fantastical and need to be explained. The combination of the ontological realism of the photographic image and the capturing of rare events was especially revealing for envisioners. Bazin, too, suggested that this appeal to fantasy was a part of the camera’s ability to present the real. He suggested that this was the reason the surrealists appealed so heavily to photography, as they “had an inkling of this when they looked to the photographic plate to provide them with their monstrosities.”21 The mechanical act of photography

Drawing Together · 179

creates its sense of reality not by simply representing time and space as it flows, but by locating improbable events that, according to Bazin, produce “an image that is a reality of nature, namely, an hallucination that is also a fact.”22 As we will see in this chapter, the reality that is also a hallucination is a hallmark of many scientific films. An understanding of how realism finds facts in unlikely occurrences is especially useful for understanding why the third scene of Ed Lewis’s movie, “Mutants of the Bithorax Gene Complex,” might best be described as a horror film. This description is especially apt, as the subjects of the scene are Lewis’s collection of homeomutants, or mutants where whole body parts are replaced, added, or lost. Homeomutants have long starred in spectacles of extraordinary bodies such as carnival sideshows and horror films. Two-­ headed reptiles or cyclopean mammals, for instance, are homeotic variations in Bateson’s use of the term. Lewis’s scene, then, already draws from a well-­trod set of expressive techniques for displaying the hallucinatory regulation of exceptional bodies. Three mutations take center stage: the conversion of halteres into an extra set of wings; the conversion of a nonlegged abdominal segment into a legged segment, giving the resulting flies eight legs instead of the usual six; and the reduction of a legged segment into a nonlegged segment producing a fly with only four legs. The mutation that adds an extra pair of wings (the bithorax mutants as covered in chapter 4) might best be thought of as a type of superpower, as it vividly recalls the transformation of mild-­mannered Clark Kent into Super­man. What was once the most normal of flies now seems to be blessed with the extraordinary superpowers of an extra set of wings!!! This type of change especially appears to highlight how extraordinary bodies appear to augment, rather than diminish, an important physiological capability. The addition of an extra pair of legs, might at first glance, also seem like an augmentation of capabilities, except Lewis chooses to film this mutant with the fly glued on its back to improve the audience’s view of its abnormality. Thus, although the fly has a new pair of legs, these legs look especially delicate as they uselessly flail in the air. Some augmentations are clearly too much of a good thing and can weaken the heroic subject. The final shot sharpens the pathos of the monstrosity narrative as it documents a four-­legged fly left to awkwardly stumble across the slide. This is perhaps the most piteous of all the mutants as the fly is clearly struggling to accomplish what his sibs could easily perform. This brings up the other genera that Lewis’s images directly evoke, the

180 · Drawing Together

Figure 5.3. Stills from Ed Lewis’s movie, Scene 3 entitled, “Mutants of the Bithorax Gene Complex.” Courtesy of the Archives, California Institute of Technology.

creature features from the 1950s, 1960s, and 1970s. In these films a mishap often turns an ordinary pest into a threatening monster, such as radiation creating giant ants (Them!), a teleportation experiment that goes awry (The Fly), or a chemical substance that makes giant spiders (Village of the Giants). As I have already suggested in chapter 3, one of the consistent themes among these films is the threat that industrial scales pose to the regulation of bodies. How can a society effectively regulate the immense power of splitting the very small, protect itself when contacting the cosmically far, or effectively distribute the vast expanses of the remarkably uniform if it can’t effectively regulate itself? Although it would be a mistake to assume that Lewis’s film intentionally evokes these themes, as that would work at cross purposes to his goal of promoting scientific inquiry, it would also be a mistake to fail to recognize that Lewis’s film is animated by the same hopes and fears as these other films. The challenge that mutations pose for scientists and screenwriters is to explore how change can best be regulated so that abnormalities are no longer considered abnormal. By appealing to these visual conventions, Lewis presents a problem that his science can hope to solve. He also strengthens the overall goal of much of the biological sciences: if problems are adequately researched, aleatory events may be regulated properly in the future.23 As we learned in chapter 4, bodies aren’t just optically envisioned, they became physically and conceptually envisioned as well. The same process

Drawing Together · 181

for envisioning technical images, creating patterns from particles and grids, was used for envisioning how real bodies were made. In Flusser’s terms, technical images are not just mirrors, they are projectors that help to encode a world with information. Fruit flies in particular have been exceptionally fruitful projectors for understanding how bodies are envisioned. In Bazin’s terms, the reality of the image has caused the distinction between “image” and “object” to disappear. As I and others have argued elsewhere, bodies first needed to be standardized before their variances could be explained. First, the heritable material of the fly was identified as a series of manipulable particles with chemical properties.24 Each mutant is produced in a way that allows for the visualization of rare occurrences, anomalies almost impossible to observe in real time fly watching. Genetics is also a form of technical imaging of the body, and drawing too hard and fast of a distinction between imaginary and real bodies threatens to destroy how deeply technical images have allowed scientists to envision and manipulate bodies as technical objects. Images are no longer removed from the real in a way that they previously were thought to be; as Flusser has observed, image and object, fantasy and reality, are now inextricably mixed together. This points to one of the more interesting aspects of our paradox of control that requires further explication. It was animation’s ability to control the effective placement of images in relationship to one another and to interweave image and text that allowed it to merge imagination and reality in new and interesting ways. This view of control is contrary to how most people align control and precision. Control is often thought to give analytical precision by providing greater certainty, such as the certainty to distinguish fiction from reality. I have not found this to be the case with animation, or with biology, where greater control of images can lead to a greater appreciation of unlikely or even fantastic events. The more control lent to the envisioner, the easier it is to present unlikely scenarios. This occurs because of the way that animation puts images together.

Waves of Change It is not especially difficult to make a film like the one Lewis made. If one has the knowledge, a few specialized pieces of equipment, and some experience with photography, it just takes work and patience. This is one of the more interesting aspects of animation—­it can be accomplished in very

182 · Drawing Together

many ways. Animated films can be made with a minimum of technical assistance, through simple techniques like stop-­motion photography, or they can be made with incredible investments in large pools of talented labor, such as the traditional animation shops of the twentieth century (think Disney), or they can be made by investing time and money in sophisticated software and film manipulation programs. Despite the immense variety of ways that animated films can be made, most of the techniques appeal to a remarkably homogenous set of concepts and principles. Understanding these concepts will further our investigation into what animation might tell us about the role of regulation in thinking about life. Perhaps a good starting place for understanding how animation works is to draw continuities with the work we have already studied with grids. One can describe a movie as a grid put into motion. When one trains a camera on a subject and presses the shutter release, the camera takes a series of photographs, where the frequency of the shots in the series are based on the frame rate, or frequency, of the shutter. The camera arranges a series of photographs as still shots on a strip of film. If the strip of film is developed and played back at the correct rate, the persistence of vision of the human eye will stitch these snapshots together to form a continuous stream of motion. A movie camera is especially good at capturing specific images from a complex stream of fluid experience. The use of grids to break down complex images into single shots is an act of partitioning, as we learned from our studies of grids in chapter 2. Yet, as we have already learned from our study of Flusser’s writings on the technical image, this act of partitioning movement into a frame also involves the partitioning of the image of the frame into a matrix of chemical reactions, where specific variations of light and color are broken down into another matrix of dots and particles that create a single image when arranged correctly. “The photographic universe is made up of such little pieces, made up of quanta, and is calculable (calculus = little piece or ‘particle’)—­an atomized, democratic universe, a jigsaw puzzle.”25 So, with films we have two acts of partitioning, the first is partitioning a dynamic world into a series of snapshots and the second is partitioning a view filled with light, shadows, colors, and forms into a series of particles on the film. Consequently, the constructivist logic of grids, what Foucault called “usefulness” (chapter 3), occurs at two different moments in filmmaking as well. In the first moment, the array of chemicals on the negative of the photographic film is transformed into a series of forms, shadows, colors,

Drawing Together · 183

and textures projected on to a screen. This leads to several creative choices in film photography. Certain shots highlight the granularity of the film, while other shots highlight time of exposure or light sensitivity. Dark room operators, as well, can increase or decrease exposure times or even bring out the details of an image by exposing only a portion of that image. The important point here is that although the camera does capture an image, it only does so through what Flusser called its “program” or the built-­in capabilities of the camera itself. “Consequently it is true that the choice of the ‘object’ to be photographed is free, but it also has to be a function of the program of the camera.”26 Thus, envisioning, or the creative act of working with technical images, requires working with and against the program of the camera to create images that are meaningful to viewers. Movie cameras add a second moment of constructivism to this process in that they assemble a series of single frames into a moving image. This provides another level for the potential of creative manipulation by the film directors. They not only get to manipulate the look and feel of the image in the frame, they also have the opportunity to arrange how the sequence of images are aligned on the film through editing. This is an especially important point in that it helps to define a crucial component for film criticism and for the aesthetics of twentieth-­century film production—­how a sequence of shots creates specific feelings in its viewers. For some commentators on film history, this ability to manipulate the relationship of frames to one another opens new ways for how viewers envision the way bodies process the relationship between time and space.27 For filmmaker Walter Murch, editing is an important constitutive practice to making a film. “Editing,” argues Murch in his popular, In the Blink of an Eye: A Perspective of Film Editing, “is not so much a putting together as it is a discovery of a path.”28 Editors often don’t just splice together shots, they help create the narrative of the film from a collection of filmed scenes. Although animation and cinema rely on editing to shape the pace and direction of a movie, animation also adds a third level of creative manipulation to filmmaking. Unlike a cinematographer who concentrates on how best to capture an image, animators need to create the world anew for every frame. For example, imagine one of the simplest forms of animation, line animation directly on the film. An animator confronted with a strip of blank celluloid must draw the elements in a way that allows for the continuity of form and location within each frame for the element to persist beyond the fleeting durations of the projected single frame. This requires

184 · Drawing Together

that animators approach the composition of a frame in a very different way than filmmakers. Even the smallest of changes in the placement or shape of an object on the film causes movement or flickering when projected on a screen. If this flickering is too great, it can disrupt the continuity of the animated objects in the scene. So, while filmmakers can assume the continuity of the world that they film, animators must create specific techniques to give their animated worlds a sense of continuity. In animation, the immediate task is not, how does one put this world into motion (by directing the subjects in the film) and then capture it on film (by cinematography) but, rather, how does one best create stability in a world prone to constant change? This additional focus adds an extra level of control for animators to regulate how their images change over time and offers those of us interested in the regulation of life an insight into how transformation can be regulated.29 The need to create stability in animation images is so important that a student of animation can fruitfully think about the history of animation as the development of techniques for regulating how images change over time. Stop-­motion animation, for instance, creates movement from a sequence of still shots. The animator carefully places the object in a new position with the change of each frame of the film. The technique gives the image the realism of photography, while the movement of objects within the frame appears to happen by magic. Objects can jump quickly from place to place or smoothly glide across a scene, all while seeming to require no external mode of locomotion. Other animators have used tracing paper to help replicate static elements in scenes without having to recreate the content of each frame in painstaking detail.30 Each choice of a technique for regulating change lends the animator’s final movie an aesthetic dimension, a tangible quality, for how change occurs in the world presented. Just as Flusser argued that photographic images emerged from the codes programmed within the camera as an industrial and postindustrial object, the change wrought by animators also reflects the political economic realities they live in. A closer look at two examples, William Kentridge’s use of charcoal drawings and the development of cel, or celluloid, animation as popularized by the Disney studios, demonstrates how deeply intertwined animation techniques are with metaphysical and political economic assumptions about how the world changes. The South African fine art animator William Kentridge animates his work by drawing directly on paper with charcoal and then capturing the

Drawing Together · 185

drawing at specific stages with a photograph. He changes his world by erasing and redrawing the dynamic parts of the image, allowing him to avoid having to redraw the whole image for each frame. His method also presents change over time in a specific way.31 In Kentridge’s world, each change comes with a trace of what has just occurred, as each new line is drawn over the erasures of the line that previously appeared. This gives Kentridge’s images a ghost-­like quality where the past continually haunts the present through blurred and smudged remains of previous images. In the opening scene of “Tide Table,” for instance, Kentridge uses his technique to transform an abstract wave of a tide table printed in a newspaper to a wave of water that crashes on a beach (see Figure 5.4). This wave recedes revealing a herd of cattle grazing on seaweed in the surf of the beach. Kentridge pits the liquidity of change against the abstraction of grids.32 Kentridge’s poetics work especially well with his scene’s subject. Trained in the fine arts, Kentridge offers an example of how animation’s ability to depict movement can bring industrialized images to life. It is also key here

Figure 5.4. Still from William Kentridge’s “Tide Table” 2003. In this image, the graphed wave function representing the tides from a newspaper tide table begins to transform into an ocean wave. Image courtesy of the San Francisco Museum of Modern Art.

186 · Drawing Together

that a wave is the transformative element of the critique. There are few better analogies than waves for thinking about how forces can turn static objects into fluid subjects. First of all, oceans provide origins for many secular creation stories.33 The sea of life creates, destroys, and marks seemingly permanent structures with its force. As Haeckel’s theories of protoplasm suggested, the bodies of living things are in a sense a strange type of shore. Organisms package seawater, sun, and earth into fleeting structures marked by the fickle proclivities of the chemistry of carbon. Life, it seems, is literally and figuratively a beach, and waves provide a compelling way to think how materials, objects, and energy interact to create patterns. William Bateson continually appealed to waves as a means of envisioning how life regulates forms. He called it an idea that was “always at the back of his mind,” and, according to his wife Beatrice, one of the “causes of his hesi­tation to accept the view that [hereditary] ‘factors’ may be, in any literal sense, transmitted as material particles.”34 For when William Bateson saw how organisms were organized he tended to see waves instead of particles: Hardly by any effort of imagination can we see any way by which the division of the vertebral column into x segments or into y segments, or of a Medusa into 4 segments or into 6, can be determined by the possession or by the want of a material particle. The distinction must surely be of a different order. If we are to look for a physical analogy at all we should rather be led to suppose that these differences in segmental numbers corresponded with changes in the amplitude or number of dividing waves than with any change in the substance or material divided.35 When Bateson saw bodies, he saw patterns. When he saw patterns, he thought of the ripples on sand or the spotting of clouds in a mackerel sky, both the products of unseen forces. What was important for Bateson then, was how these forces formed patterns, an inclination that seemed to put him at odds with a growing body of experimental research in genetics that tended to view bodies as composed of characters formed through particles found on the chromatin of living things. As he wrote to statistician G. H. Hardy in 1924, “We have had some absurd attempts—­mostly from ­biometricians—­to apply mathematics to biology, but as I said, my hope is

Drawing Together · 187

still that I may live to see mathematics applied to biology properly. The most promising place for a beginning, I believe, is the mechanism of pattern.”36 Bateson hoped a study of patterns would help him not only to understand how things were made but also how they were arranged. For example, Bateson wasn’t simply interested in the fact that organisms had repeated segments, he was also interested in how and where the segmentation appeared. He asked questions such as why some tissues appear segmented while others do not (see Figure 5.5). Bateson understood that an investigation into the materials of heredity weren’t enough on their own, there needed to be an allied investigation into how forces arranged materials in specific ways, or how they were ordered. “Our tissues therefore are like a beach composed of sands of different kinds, and different kinds of sands may show distinct and interpenetrating ripples.”37 The key was to start thinking how these unseen forces produced the visible ripples of living flesh. Yet commentaries on the metaphysics of life, at some point, are also commentaries on the political economy that birthed them. This is obviously the case for the scene from Kentridge’s “Tide Table.” Kentridge trained as a fine artist and brings that sensibility to his animation, however

Figure 5.5. A normally unsegmented structure like the tusk from an elephant can appear segmented at times, “showing how easily a normally unsegmented structure may be converted into a series of repeated parts.”38 Image from William Bateson, Problems of Genetics (New Haven, Conn.: Yale University Press, 1913), 37.

188 · Drawing Together

animation came to prominence as an industrialized and not a fine art form. Twentieth-­century animation developed into a highly s­ pecialized production system that required extensive divisions of labor, large amounts of money, and large workforces dedicated to producing and distributing it as a final product. Kentridge is clearly working against these tendencies by producing images wrought by a single hand and a reliance on the art establishment to help distribute his work. The opening image doesn’t just depict an abstraction of a wave as a tide table provided by a newspaper. Newspapers embody many principles of industrialization. They are mass produced, inexpensive, and are all produced in a similar way—­using grids. Grids allowed newspaper publishers to simplify and control the layout of their pages, allowing them to create varied printed surfaces from a remarkably homogenous collection of parts. The inclusion of special features like tide tables (as well as comics, as we saw in chapter 2) helped publishers sell newspapers during days when headlines weren’t that exciting. A tide table in a newspaper not only serves as a means for informing its readers about the state of the ocean at the beach, it also helps regulate the distribution of the paper itself. The brief little scene in Figure 5.4 neatly packages Kentridge’s aesthetics, his thoughts about regulation and life, and his critique of industrialization. The scene uses an ocean wave to wash away the abstraction of industrial life, replacing its grids with the fecund meeting of the forces of earth, sea, and sun. What Kentridge’s movie seems to ignore, however, is that waves don’t just wash away order, they provide a more fluid form of order than a grid. As Bateson glimpsed, the energy that invigorates a wave not only has the power to destroy, it also has the power to create. In fact, the deeper one looks into how animation orders change, the more waves one tends to find. A closer look at what Kentridge was reacting against, the studio production system of animation, is especially informative for understanding the way animation often provided a model for how the fluidity of living things was a consequence of the most highly regulated forms of order.

Layers of Control: Regulating Change through Composite Images When most people think of animated films they think of Walt Disney and two-­dimensional celluloid animation. Cel animation, as it became known, heavily borrowed from industrial processes in order to balance continuity and change in the moving image. Cel animation standardized animation

Drawing Together · 189

techniques, partitioned tasks into specialized roles, and then created hierarchies of labor to manage workflows. Almost all technical developments of cel animation were intended to more effectively regulate change over time. In cel animation, animators draw directly on transparent cels, or sheets made from cellulose nitrate or the less flammable cellulose ace­tate. The transparency of the cels allowed animators to layer drawings on top of one another when they wanted to change the position of a part of the drawing. A historically important demonstration of this process is provided by E. G. Lutz in his 1920 book, Animated Cartoons: How They Are Made, Their Origin and Development.39 Although this demonstration isn’t the most visually stunning, it is one of the most historically important as it is rumored to be the book that Walt Disney used to train in the art of cel animation. In Lutz’s illustration below (see Figure 5.6), he starts with the background scenery drawn on one of the sheets. Notice the two holes at the top of the sheet. These holes are registration marks and are used to quickly and precisely align the celluloid sheets on top of one another. Ideally, the cels that change the least frequently would be placed near the bottom of the stack. In the image below, Lutz chooses the trees and grass of a park as the background image.

Figure 5.6. The background image for Lutz’s animation. Illustration from E. G. Lutz, Animated Cartoons, 70.

190 · Drawing Together

Lutz then builds a depth and movement in his image by layering other cels over his background image. Immediately on top of his background image is the relatively static image of the torso of the two men (see Figure 5.7). He then layers the most changeable of elements of the compo­sition, the gesticulating hands and the speaking mouths of the subjects, on top of the torsos. Each frame of the film, then, is created as a composite of the layers of all the different cels. When the animated film is shown, the scene appears to come to life as parts of the composition move at a different rate from other parts. In the simple scene presented by Lutz, the only mobile elements of composition are the hands and mouths of the subjects. There is nothing that precludes the application of this technique to more complex scenarios, however. One could imagine a background where the trees and grasses gently swayed in the breeze as well, albeit at a different rate from the hands. This is the type of control of relative motion that the use of layers, added to the power of registration, can bring to an animated world. It can provide control over multiple rates of change over a series of frames. Many studios developed differentiated labor systems in the construction of celluloid animated films. Some of the most experienced artists, for instance, just drew the “key frames,” or the frames where motions started and ended, while other less expensive animators, called “tweeners,” would draw the transitional frames between each of the key frames. In the United States, these jobs became highly codified through the unionization of labor. It was in this sense that the political economy contributed to this labor intensive, industrialized, and exquisitely controllable form of animation. Partitioning of the labor process helped to build greater control into how images changed. Although Lutz’s illustration is useful for showing how drawings can be combined, its focus on the frame makes it hard to get a sense for how the actual movements are propagated across the frames. Kentridge’s use of waves to depict motion in animation is more than an evocative analogy, it can serve as a heuristic for understanding how single frames of film can be used to regulate multiple durations of change. One first needs to understand how a wave on the water is propagated. Those who study waves often break them down into different types depending on the medium of the propagation and the direction the energy moves in relationship to the propagating medium. In an ocean wave, energy is propagated across a liquid through an orbital shaped wave. This means that although a wave will pass from one shore to another, the particles of water that propagate that

Figure 5.7. Here Lutz demonstrates his use of cels as layers to be added on top of the background. Illustration from E. G. Lutz, Animated Cartoons, 71.

192 · Drawing Together

wave effectively return to where they began. Imagine a cork bobbing on the ocean (see Figure 5.8). When a wave passes underneath the cork, the wave doesn’t take the cork with it. Rather, the cork, like the water it floats on, moves in a slow orbit back and forth. The energy of the wave effectively moves transversally across the water while the particles that propagate that energy only move in an ellipse.40 Movement in animation works in a similar way. Although each of the frames appears static when viewed individually, the eye sees the movement that takes place across the frames when the film is projected. The way the images are ordered in relationship to one another across the frames is what shapes the movement, not just the single image in a frame. Much like the moving wave to the relatively static cork, it is the depiction of the force conveyed by the object and not the object itself that is displayed. Also, like an ocean wave, many motions are not simple motions; they often exist in constant association with other waves. The wake from a boat, for instance, combines with the ocean wave that rocks that boat to create a complex combined wave form. The compositing of layers in animation allows one to compile together changes with different periodicities, much like the combination of the high-frequency crests of a boat’s wake with the longer durational crests of the ocean wave. These superimposed complex waves can be mathematically broken down into a series of component waves. Take for instance the diagram in Figure 5.9 which shows a complex wave form on the bottom, and then breaks this complex wave form into a series

Figure 5.8. The motion of an object floating on a wave. Note that the cork on the wave moves in a small elliptical pattern while the wave travels across the ocean. Image courtesy of Karl Benitez.

Drawing Together · 193

Figure 5.9. A Fourier Transform analysis of a complex wave form. The analysis allows an investigator to break down complex wave forms into the more basic constitutive simple waves that comprise it. Image Courtesy of Karl Benitez.

of three simpler waves of different periodicity and amplitude. This act of breaking a complex wave form into multiple sine and cosine wave forms is a mathematical analysis called a Fourier Transform analysis.41 Now imagine that we wish to chart the different periodicities of movement within an animated film. The final animated movie one sees—­the scene of two men talking in the Lutz illustration, for instance—­would be like the complex wave form at the bottom of Figure 5.9 as it is composed of movements with multiple durations. The topmost layer of the Lutz animation is the frequent movements of the hand and mouth gesticulating, the second layer is the periodicities of bodies of people in the park, the third layer would be the periodicities of the movement of the background images. Although this is a fairly rough demonstration since these periodi­ cities covered by each of the layers of movement are complex wave forms in themselves, the idea remains the same. By breaking a scene of complex movements into manipulable individual layers, one can effectively break

194 · Drawing Together

a complex wave into less complex and more easily regulatable simpler waves. The telling difference of course is that Fourier Analysis partitions complex waves into simpler ones while animation composes simple motions together into a complex composite using grids (for the registration of the images), layers (the composite of cels), and frames (using a motion picture camera). So, although cel animation creates moving images, it does so by providing a mechanism to regulate the changes of the components of the image. Therefore, a great number of computer graphics and animation programs have adopted the mechanisms of control provided by cel animation, even though they no longer use celluloid sheets to enact them. A graphics program like Photoshop for instance, uses layers in its compositions to allow illustrators to isolate specific compositional elements. Visual effects programs like After Effects or animation programs such as Maya generate composite images by layering many simple timelines to create a composite complex timeline. They also preserve the distinction between “key” and “tween” frames in an animation workflow. The difference, of course, is that the computer automatically tweens the movement between key frames, effectively replacing the lower paid illustrators. As Thomas Lamarre has argued, animation is less about a simple production of movement than bringing together different types of movement in a single image. Understanding the use of composite images in animation changes everything.42 Yes, animation is a useful entertainment and pedagogic device because it depicts moving things. But it doesn’t achieve this motion solely by transferring energy to a static world. Animating an image isn’t like kicking a ball; more fundamentally, animation draws together different rates of change of a world already in motion. The use of composite images allows modern animators to regulate different rates of change in a dynamic and often complex visual world. Understanding how animation works can even help one perceive how the movie camera, as Flusser stated, is a product of the codes of industry. It is easy to see how the regulation of time works with animation, since editing can be used to manipulate the passage of time in the final film. Yet, film construction has its layers too. Scene construction, the compositional dynamics of mise-­en-­scène, and the importance of actor placement can all be viewed as the juxtaposition of different layers to create a final scene. Since cameras are so efficient at packaging together a large amount of information in a single shot, we tend to overlook this construction. In ef-

Drawing Together · 195

fect, they flatten all the dynamics of different layers into a complex image. Understanding this process helps us see how the composition elements of creating a shot within a scene are similar to the layers in an animated film. The relationship between form and time is created by composing a series of overlapping sequences. It is easy to downplay this important element for some types of filmmaking processes, where the camera is turned on and simply captures a scene, without much thought on how the elements within that scene relate. Studying animation, therefore, offers a way to unpack some of the assumptions that go into why certain forms and certain temporalities are drawn together to create a final composite image. In fact, it is possible to invert the prevalent assumption that animation is a special case of cinema, in that they both use a camera, to claim that for some purposes at least, cinema can best be viewed as a special case of animation.43

Composite Motion in a Cel(l)ular World Viewing animation as a way of controlling different rates of motion changes how we view Ed Lewis’s animation. It’s not simply that he is using animation to show movement, he uses animation to combine three different rates of change into a single image: the rate of change of the background (in this case static), the rate of change of molecules interacting, and the rate of change of an organism developing. During the entire movie, he even depicts evolutionary change as well (in Scene 2, for instance). This allows Lewis to make complex temporal comparisons in his scientific argument. For instance, he can demonstrate how some forms of molecular processes, such as the expression of a gene, change developmental processes, such as the formation of bristles on a larva segment. Lewis isn’t accurately representing specific rates of change, as these processes are impossible to directly witness or film. Animation allows Lewis to draw together different relative rates of change to suggest that these might be interrelated (or in terms of the analysis of chapter 3, these two forms of order might warp each other). Viewers can then see how molecular change, organismal change, and evolutionary change can interact to create a complex account of biological change. More recent animators, working with a full suite of computerized tools, have used animation’s ability to depict relative frequency of motion to suggest how order emerges from the frenetic pace of molecular kinetics.44 In the hands of a master animator, such as Drew Berry, the scientific

196 · Drawing Together

exploration of movement and change informs existential speculation about the role of stability and change in cellular spaces. Trained with a master’s degree in cell biology and possessing extensive experience using computerized graphic tools in medical illustration, Berry now manages the production of biomedical animations at the Walter and Eliza Hall Institute of Biomedical Research in Melbourne, Australia.45 A MacArthur Fellow, Berry has received awards from scientific institutions, such as the publication Nature and the National Science Foundation, as well as from the entertainment and technical industry with an Emmy and a prize from the animation software suite Maya. Berry embraces this recognition, split as it is between entertainment and scientific sources, as he sees himself mainly as a movie maker who uses scientific data to create engaging educational films. As he stated to Guardian reporter Stephen Curry, “The goal of my work is to show non-­experts—­the general public aged 4 to 99, students of biology, journalists and politicians, and so on—­what is being discovered in biology, in a format that is accessible, meaningful, and engaging.”46 Yet, edu­cating for Berry means walking a fine line between entertainment’s need to captivate and science’s need to inform. So although he freely claims that “my work is not intended to be a lab-­bench-­calculated model for research use, it is an impressionistic, artist-­generated crude sketch of phenomena,” he is also at pains to stress how his animations are built from scientific data. “I would then assert that the animations are firmly founded on real data and are as accurate as I can possibly make them.”47 Animation’s ability to control different rates of change is key to Berry’s ability to balance his seemingly conflicted desires to entertain and inform. In a scene from his movie on programmed cell death, or Apoptosis, Drew Berry uses animation to suggest how the frenetic and random movement of atoms can lead to macro molecular and cellular change (see Figure 5.10). Berry’s choice of scene is saturated with existential drama, as programmed cell death is a type of death initiated through cellular mechanisms. This makes it distinct from other types of cell death due to violence or injury to a cell, which is a type of cell death known as necrosis.48 Programmed cell death, then is part of a body’s ability to balance stasis and change as well as a major mechanism in generating organismal forms. Most human hands, for instance, lack webs of skin between the fingers, as these cells were somehow “programmed” to die off during development. In Figure 5.10, I have abstracted three frames from a scene that lasts only a few seconds. Drew Berry uses this scene to demonstrate his conception

Figure 5.10. A brief scene capture from Drew Berry’s film Apoptosis. In this scene Berry utilizes differing rates of movement to depict the initiation of a molecular cascade for programmed cell death. Molecular images from Apoptosis animation by Drew Berry, The Walter and Eliza Hall Institute.

198 · Drawing Together

of how a “death ligand” from a T Cell binds to receptors on the surface of a diseased cell, triggering apoptosis. In these reproductions  the healthy cell is the mass at the top of the image and the diseased cell is the darker mass at the bottom of the image. The dark mass at the center of the image is the “death ligand” and the “death receptors” are the movable lighter elements in between the cells. Once bound, the receptor initiates a molecular cascade within the diseased cell that dissolves the cell’s scaffolding proteins and leads to the dissolution of its integrity (a process known as “blebbing”). Unfortunately, what is most amazing about this scene is what I can’t easily demonstrate with a screenshot. Berry has carefully combined different rates of motion among elements within the scene to lend the scene a sense of dissolution and dissolve. Using what is known about biological kinetics, Berry portrays the smaller molecules that make up the membranes of the cells as moving at a much faster rate than the more complex larger mole­ cules of the death ligand and the death receptor. Berry effectively bundles different stages of change to make an existential point about cellular and molecular dynamics. In Figure 5.11, I have used the diagram from Figure 5.9 to roughly depict how rates of movement in the animation can be conceived as a combination of different wave forms. The overall effect is to suggest that biological order emerges out of the turbulent kinetics of molecular movement. This transition to stability is elegantly reversed in the film as the enzymatic cascade of cell death breaks down macromolecules, liberating the kinetic potential of their constitutive molecules. From a cellular point of view, death is not the cessation of all activity into existential nothingness, but a loss in the variety of the frequencies of change. It is also important to recognize that Berry isn’t depicting real rates of change. Molecular movement happens much too quickly and at much too small a scale to be perceived by humans. Representing molecular movement, as Flusser suggests about all forms of envisioning, is an inherently absurd proposition. Berry uses the technical capacities of animation to envision the different scales of movement. The reference point is not unaided human vision but another more familiar form of technical image, time-­ lapse microscopy. “To represent cellular dynamics I strive for the look and feel of movements observed under time-­lapse microscopy. The molecules and their Brownian motion is an emulation of molecular dynamics simulations, created with the various dynamics and video game engines in 3D software such as Maya, but with speeds manipulated and slowed down to

Drawing Together · 199

Figure 5.11. A rough depiction of how different rates of change in Berry’s Apoptosis interact without collapsing together to create a single complex wave. The top module (A1) denotes a frequency of atomic movement in the animation; the middle module (A2) denotes the frequency of molecular movement; and the bottom module (A3) denotes the frequency of cellular movement.

make the action watchable for an audience.”49 Technical images inform by packaging together truth and fantasy, data and games, and not just by separating them. They provide an impossible point of view from which envisioners can collect numerous strands of data into a compelling image. Their value it seems, come from their ability to draw together observations from a wide number of sources. This doesn’t make them “right” or “wrong,” but it does open them up to criticism about which parameters need to be portrayed more sensitively. Even though Berry recognizes the absurdity of his animations, he remains quick to claim that his animations remain “as accurate as I can possibly make them” in that he has built into his movies observations culled from long hours of literature review. Berry literally thinks of his movies as a drawing together of different types of data into an understandable format. The analogy he uses is the more plentiful and less contentious act

200 · Drawing Together

of drawing together findings provided by articles that review a specific topic of the scientific literature. “My work,” argues Berry, “is most akin to a ‘review’ paper in the literature, presented in visual form.” In effect, viewing the molecular dynamics of a cell’s interior requires the absurdity of technical images to be informative. I think it is important to stress what we mean by “absurd” and its relationship to representational knowledge in this case. It isn’t that representation isn’t important; Berry’s animations obviously use representational cues in their composition. However, the idea of discerning what is true through an assumption of the primacy of representation is absurd. Yet, not all absurd points of view are allowable, just the ones that are derived from sources of data that can be fruitfully composited in a complex but internally consistent technical image. What emerges in importance is veracity, or faithfulness to the data source, as well as robustness, how well the data fits with other pieces of data. Animation then carries an epistemological value in that it allows researchers to see how data fits together in a composite image. This point has been repeatedly stressed by biomedical animator Janet Iwasa, and in doing so, she has drawn a continuity between animation and another scientific convention, the illustrated model.

Envisioning as a Dynamic Scientific Model In the sixth and concluding scene of Lewis’s movie, “Regulation of the Bithorax Gene Complex (a model),” Lewis animates his model of gene regulation (see Figure 5.12). Lewis uses two important qualities of animation: he controls motion to show how molecular interactions can lead to developmental realities through single discreet steps, and he presents multiple scales (organismic and molecular) in a single scene to demonstrate how molecular interactions affect physiological structures. Although Lewis’s use of cutout animation dates the look and feel of the scene, the pedagogical goal of the scene is amazingly prescient. As we will see, much of the appeal of current biological animations are that they can be used for the more speculative aspects of biological practice, such as the presentation of hypothetic models, as well as provide a medium for coming to grips with the scientific and existential issues involved with interactions across atomic, molecular, cellular, and organismic scales. The scene begins with an establishing shot showing how some regulatory molecules in the bithorax complex are likely expressed in a gradient

Drawing Together · 201

from the most concentrated in the head area to the least concentrated in the tail area of drosophila larva. The larva’s only features, at this point, are its segments and the gradient depicted as a decrease in shading moving from head to tail (Figure 5.12a). The second shot from this scene replaces the illustration of the gradient with a depiction of the rudimentary physio­ logical features of a wild type developing larva. Layered on the bottom of the frame is a genetic map of the bithroax cluster, with arrows indicating which gene segments seem to be implicated in specific segmental changes. Remember here that the bithorax cluster has the remarkable property of colinearity, where the sequence of genes on the chromosome mirrors the sequence of segments that the genes effect. Lewis is suggesting that the head-to-toe gradient in the developing larva could help explain the co­ linear order of genes in the bithorax cluster (see Figure 5.12b). The scene then changes scale as it zooms in to demonstrate how the effects of the gene regulation work in each segment. Lewis first shows the Thoracic 2 (T2) and Thoracic 3 (T3) segments with their embryonic features (see Figure 5.12c). He then enlarges a single cellular nucleus from each segment so that viewers can envision how proteins might bind to the DNA regulating the expression of the genes in the cluster. Lewis zooms in by substituting larger size paper cutouts for smaller ones as the magnification of the nucleus increases (see Figure 5.12d). It turns out that zooming in, or all movement into the depths of a scene, is a difficult effect to achieve in animated films. The layering of animated films makes movement across and up and down a layer (over the grid and not across it) easier than moving through the image (moving across grids). Lewis’s use of cutouts is a simple and efficient way to change the apparent magnification of a scene without changing the perspective of the viewer. In twenty-­first-­century animated films, computerized animation programs, such as Maya, would change magnification by changing the position of the virtual camera in a cell. This effect forces the perspective of the viewer to change as the focus shifts across the landscape of a scene. This type of “fly through” effect is much easier to accomplish with computerized animation than with cutouts or cel animation.50 The next shots from this scene demonstrate how each of the transcription products of the bithorax genes contribute to the development of the tracheal trunk or a specific cuticle structure, such as how the bristles band on each segment (see Figures 5.12e, f, and g). The key point from this set of illustrations is that Lewis thought that bithorax complex proteins act as

202 · Drawing Together a

b

c

d

e

f

g

Figure 5.12. Frame captures from the sixth and final scene of Ed Lewis’s animated movie. The scene is entitled, “Regulation of the Bithorax Gene Complex (a model)”. In this scene, Lewis uses animation’s power to speculate about interactions over time to explain how his model works. Courtesy of the Archives, California Institute of Technology.

genetic repressors, that when the gradient of the proteins lessens toward the anterior of the larva, the repressors become less effective, thus allowing the larva to express more structures. A mutation to these proteins then eliminates the suppression necessary for the development of a wild type embryo in a specific segment and can create the jarring effect of having a structure produced in a segment where it should not occur, such as what we have previously seen with homeotic mutants.

Drawing Together · 203

It is important to once again stress that Lewis chose animation to explain his complicated model of molecular development. Animation allowed Lewis to work at scales too small or too abstract to visualize as well as to quickly change views between molecules and organisms. It allowed him to do so at a time when most types of computers couldn’t handle the complexity of the moving images he wished to create. Although Lewis uses animation to put different cellular elements into motion, he does so to explore the feasibility of how his ideas on molecular development might work. He is using his animations, in other words, to present a dynamic model, significant because it can demonstrate how change occurs over time. Lewis’s diagrammatic rendering of the fruit fly larva is also important in that it doesn’t pretend to be scientific realism. Rather, what Lewis presents is intended as a hypothesis whose value comes from its ability to explain, by drawing together, several seemingly disparate phenomena. Lewis appreciated how animation as a form of temporal regulation is especially fit for the demands of conjectural reasoning. Other researchers have been drawn to animation for these same reasons.

Putting Ideas into Motion: Regulating the Fluid Nature of Thought As a postdoctoral researcher at University of California at San Francisco, Janet Iwasa recalls the moment that she became convinced that animations could be used for scientific research. The key for Iwasa was thinking of animation not as a representation, per se, but as a model or hypothesis. In her article “Animating the Model Figure,” Iwasa compares animations to the more conjectural model figures often found at the end of a scientific paper.51 These figures appear in the paper after the scientist has presented and evaluated evidence and tries to determine what this evidence means. For instance, Figure 6 from Lewis’s 1978 paper is a fitting example of a model figure (see chapter 4, Figure 4.12). This illustration isn’t intended to represent Lewis’s data, but instead to offer a means for speculating how the data best fits with his and others’ observations. As Iwasa notes, “While the main purpose of creating a model figure is typically to communicate a hypothesis to a wide audience, the process of constructing the model figure is often illuminative. When we start to draw a model figure of a process that has only been imagined, we must go through an active and creative mental exercise of concretely defining the properties of the model.”52 Consequently, it isn’t so much that the presentation of the figure is important

204 · Drawing Together

but that creating the figure in the first place can be generative. Iwasa’s insight is that working with animation can extend this generativity to not only thinking about how things are but also to how things change over time. The epistemic value of animation comes from how it draws together seemingly disparate observations into a form of fluid, or changeable, view of how the world might operate. This form of fluidity comes from two levels of thinking about change. The first level is that animation is a time-­based medium and is especially fitted to envisioning how things can change. Again, Flusser’s concept of envisioning is an important word here as animations provide a way to envision unlikely scenarios by their ability to draw things together. Although scientists are interested in how things are represented, they are mostly interested in data. Where does it come from? Is it accurate? How does it fit with other data sets? A model in these terms is not a form of rendered reality but a means of combining data streams, either technically, by feeding data points to make a computer graphic, or conceptually, by seeing how things fit together in a moving model figure. Iwasa suggests that part of creating composite images is that the method offers a way for researchers to evaluate how time-­based processes fit together. This ability comes from animation’s ontology as a time-­based process as well as the ability of composite figures to draw together disparate elements. Animation, it turns out, is especially suited for exploring what William Bateson called “the mechanism of pattern.”53 The second level of fluidity important in Iwasa’s work is her emphasis on the conjectural basis of modeling. In this case, animation is useful because it is abstract and ephemeral. Conjectures are useful as they focus issues instead of solving them. The idea is not necessarily to see what is, but under what conditions something can possibly occur. This emphasis on ephemerality has influenced the aesthetics of Iwasa’s animations. For instance, on her webpage Iwasa presents animations as a series of provisional drafts instead of as a fully rendered masterpiece of cellular ontology. Consequently, she adopts an aesthetics of ready-­made shapes, surfaces, and colors for the elements of her films. Take for instance, “HIV Integration: Early Draft” in Figure 4.16. This lack of attention to the existential thickness of molecular existence, its texture and complexity, lends some of Iwasa’s animations the look of a stop-­motion animation made from preformed pieces. This places them somewhere on the technical animation

Drawing Together · 205

scale between Drew Berry’s sophisticated use of visualization software and Ed Lewis’s use of cutouts. In the scene presented below Iwasa uses stock shapes to depict an HIV virus’s ingress into a cell (Figure 5.13a, b, and c show the entry of the virus through a cellular membrane and viral travel through the cell), its co-­option of the cell’s replicatory mechanisms (Figure 5.13d and e), and the egress of the newly produced virus through the nuclear membrane (Figure 5.13f ). Still, she retains some of the ability of the animation program Maya to set some of the more cinematic elements of the film, such as shot and scene. Iwasa has been especially creative in thinking about how to make animation available to more researchers. Mastering the complex software used for digital animation is time consuming as well as unnecessary to depict the more straightforward animations that Iwasa wants to promote in science. As Iwasa explains, “Three-­dimensional animation software is well known for having a steep learning curve, requiring time and dedication to learn how to do even the most mundane tasks, as well as patience and tenacity to work through the inevitable bugs, crashes, and other technical issues.”54 Iwasa has even developed a software program for biologists in order to encourage them to begin animating their models. Fittingly entitled Flipbook, this program is intuitive to use as its interface is uncluttered, the tutorials are helpful, and the means of controlling one’s drawings are straightforward. The basic concepts are the same tools as more sophisticated animation programs built on the use of grids and layers, but the presentation is highly streamlined. In order to facilitate its use of existing data, the program links directly with Worldwide Protein Databank (http:// www.wwpdb.org/), and researchers can download known protein structures by their designation (see Figure 5.14). Once the proteins are downloaded, the program renders the protein into a space-­filling model and offers researchers options in protein placement, texture mapping, shading, and shape construction. All components are placed on an animation stage that compiles images to put the structures into motion and then easily share them with others on a website. Iwasa used the free and open source software Blender as a base for the project. Blender is a cross-­platform animation software produced for the entertainment, design, and gaming industries and was created to encourage a community of game designers and animators. The developers at Flipbook then adapted the program for the creation of animations containing biomolecules.

206 · Drawing Together a

b

c

d

e

f

Figure 5.13. HIV integration: early draft, Janet Iwasa. In this sequence of stills, Iwasa suggests how a viral particle might enter a cellular milieu and appropriate its ability to translate RNA into new viral proteins. She has sacrificed some of the detail of her molecular world to more easily depict how things might interact and change. Image courtesy of Janet Iwasa, University of Utah. Original Publisher, Science of HIV website (http://scienceofHIV.org), 2015.

It is Iwasa’s focus on the role of time in molecular movement and the importance of provisional hypothetical statements that helps us gather together ideas presented in this chapter on the use of animation as a form of fluid control. In a sense, the term “regulation” as we have used it in this manuscript has always been time dependent. For Foucault, regulatory systems needed to respond to aleatory events to move from systems of control to systems of security (see chapter 3). With animation’s exquisite capability of controlling multiple frequencies of change in a single image, we see this ability come to fruition as an enabling condition for envisioning biological change.

Drawing Together · 207

In the discussion of biology and animation techniques, it is easy to miss that the act of creating an animated composite as a form of drawing together is already a model of time. It is a model built from a collective experience of a past (as opposed to a purely phenomenological one), navigated through the capacities of a technical present (through the codes of technologies used to envision technical images), and predicated on offering a vision that is only partially glimpsed and open to revision (as is the nature of all modeling). This does not mean that there is not a real behind the fantastic images of biological animations. But the real is forged through the difficult process of understanding scientific data, making a movie, and then offering this vision up to others as a form of data for revision. This also does not mean that all images that follow this process are equally useful. As we will see, with all fluid forms of regulation, the very act of bringing something together should, in most cases, warp the logic of the grid that is used to bring it together in the first place. Some forms of bringing together do this more than others.

Figure 5.14. Molecular Flipbook interface with space-­filling model of Homeobox protein HOX-­B1. Notice the suite of tools on the left of the screen, the time bar for the animation at the bottom, and the abstract grid in the background.

208 · Drawing Together

Composite Lives: Living the Weird and the Robust In my discussion of animation, I have centered the role of compositing in the biological sciences by emphasizing how animation creates compelling visions of the world by gathering together disparate elements. It can be difficult to discuss the role of control in visualization because visuali­ zation inherently draws together the ability to regulate and the ability to know in new and challenging ways. We saw this with grids, as grids helped to ensure that placement within an image became as central to meaning making as what was being pictured. We see this again in the composite animated image, as drawing together involves judgments based on how things should be ordered and how they should best interact over time. It is animation’s ability to regulate fluid or dynamic states that makes it such an interesting tool for biological thought as this ability gives thinkers a means to interrogate how movements relate with one another. Just as grids simultaneously enabled construction (by being used as a scaffold) as well as analytical clarity (by being used as a tool for partitioning into smaller parts), composite animations extend this nonlinear conceptual advantage through time. Too often, however, our animations of the world (as well as many other statements and images) are offered as a fait accompli, as an authoritative picture of what is occurring, instead of the restless probing of an experiment in envisioning.55 It is important to keep in mind, however, that the view of the world that animation offers is a view based on the tools and dispositions of a specific political economy and is not unfettered. This insight needs to be embraced by scholars who study the development of scientific ideas; it is because of animation’s relationship to a political economy and not despite it that animation’s study can teach us much about contemporary biology. Two points from this chapter deserve a final emphasis as a coda for the whole book: recognizing the act of drawing together is an epistemic value and drawing together allows for the envisioning of rare and unlikely events. Despite numerous claims over the past century and a half, biology is not a unified science.56 Biology uses techniques and insights from physics, chemistry, cellular biology, natural history, evolution, comparative anatomy, structural biology, genetics, and on and on. The virtue of biology is not that it has unified these perspectives, but that it has found ways to traverse them when it needs to. For instance, structural biologists need to have mechanisms for moving between the dynamics of molecules and

Drawing Together · 209

the dynamics of cells. These perspectives should never be presented as removed from debate about the reliability of data, conventions for representation, or on the reliance of genre to create arguments. Most importantly, the use of these mechanisms should encourage further experimental inter­ ventions, observations, and careful study. A growing group of biologists today understands the need to see things pragmatically and empirically, by building things to see what works. This makes a study of scientific process even more urgent than it currently is. To use a hackneyed but understandable metaphor, if scientific observations illuminate the world it studies, it only does so by illuminating the observer as well. To observe is to draw together, to make a composite, to traverse a chiasm,57 of a world constantly in interaction, an interaction that includes the observer as well. This brings us back to Flusser’s claims about the role of envisioners in creating informative and unlikely combinations. Drawing elements together should not just be a reduction, or even a synthesis where two disparate elements create a novel third element. As animation has taught us, there are many ways to draw things together. Spatially we can have seamless composites, such as closed compositing, where all elements in an image appear propelled by the same force. Or we can use open compositing, where elements are held together in ways that their meeting produces strange occurrences, such as jumps in placement, vibrations, or the morphing of shapes. These are the little pests, that when produced, generate their own types of order, not just through a synthesis in the creation of a whole new organic order but as resonating perturbations, vibrant waves of patterns formed on flesh, celluloid, and screens. These unlikely events present a challenge to understanding in that they suggest that the normative conditions that one thought were occurring, the ordering of a grid for instance, are not always enough to explain the outcomes that one is witnessing. All orders have a para-­site (next-­to) even as they have an outside. Analyzing composites is a way of understanding how being next-­to creates a world of depth and motion.

This page intentionally left blank

Epilogue Toward the Nonsynthetic Care of the Molecular Self

The beating hearts of lives lived in the grid carry within them a beautiful paradox. Regulations increase the weirdness in the world even as they promise to bring control and order. It isn’t so much that gaining greater control of the world exhausts its strangeness by turning it into a series of clock-­like events. It’s that the world is bifurcated; it is both fundamentally weird and fundamentally ordered at its very core. Consequently, envisioning the world shouldn’t just mean making it more calculable, more designable, or easier to engineer; nor should it mean rejecting all forms of building and all types of constructive systems of knowledge as a deepening of control. It does suggest, however, that we proceed provisionally with two fundamental convictions. We should recognize our ethical duty to evaluate systems of order to see how they privilege some lives over others. We must continue to become aware of how internal differences are generated with the use of grids and other heuristics for ordering the world. This requires participants to move beyond a politics of inclusion to ask how the ordering of those included also matters. Recognizing inequity alone isn’t enough. We also need to recognize the profound importance of anomalies and exceptions. Exceptions are not so much blind cries at the grid, although that may be included as well; rather, they are an indicator that other grids are present. Anomalies not only suggest that the world is more complex than one presupposed, they also indicate heterogeneous spaces and times where life, in all its pestiness, can bubble forth. A key strategy for supporting lives in the grid, then, is to recognize how life itself is a composite of multiple spaces and times. Vibrating atoms in my body bind to create the temporary stability of long chain molecules. These molecules have an enormous amount of chemical and morphological diversity, allowing them to interact in an amazing number of ways. Yet, 211

212 · Epilogue

stability is never what it seems. Cells within organs continually reproduce and die. And my body is the product of the relationship of many other bodies. These include relationships that began way before I was born and will persist long after I die. Each element within the grid is the product of a heterogeneous collective. My life incorporates several types of bodies (animal, plant, bacterial, and viral) with their own life durations and life cycles. What we call “life” isn’t a single metaphysical principle that death negates, it is a teeming collectivity that warps and changes its regulations through its interactions across vastly different scales of time, space, materials, and forms. Explaining life, then, requires skills in very different types of “compositing” to appreciate and understand the variety of ways the world is collected. And this is what we see in how biology is practiced today—­the use of a variety of tools for envisioning how living things exist. Biologists continually use stories, animations, models, calculations, propositions, theories, observations, arguments, diagrams, molecular sequences, photographs, and movies to understand life in its complexity. Recognizing how a composite image or statement is gathered together, then, is especially important for understanding how ontological diversity contributes to lives in the grid. As we saw in chapters 1 and 4, a body is not just ordered through the demands of its final form, it is also ordered through its interactions of molecules, cells, and modules. Form, it turns out, is the consequence of this ordering and not just its cause. These interactions need not result in a higher synthesis, where a new novelty of function comes into being. Molecules can make cellular organelles, for instance, but they more often don’t. And even if they do create a new entity in one organism, the same molecule might create something very different in another organism. Each element in this heterogeneous soup can have multiple functions. As recent biology has taught us, genes not only code for proteins, they bind proteins as well. Practicing the delicate art of understanding lives in the grid should ensure that we understand how things are held together without feeling compelled to synthesizing all interactions into a final product, form, or category. This view of understanding living things deserves special attention in a world where knowing is productively conflated with making. To use the term Flusser introduced in chapter 2, understanding the world increasingly means envisioning it. And as we discovered in chapter 4, envisioning the world is no longer a task that takes place on a page or is confined to a

Epilogue · 213

screen, it now occurs through the manipulation of materials and organisms. This suggests that animation is only the tip of the iceberg. The rise of synthetic biology, the study of biomimicry, the renewed enthusiasm for studies in developmental biology, and even the exploration of the relationship between biology, art, and design now embrace, even if only implicitly, this new form of doing as knowing. Understanding the implications of this discourse requires attenuating the current emphasis on “innovation” that animates fields such as synthetic biology, biological design, or bioengineering, in favor of a discourse of concern and solicitude. This can be through an emphasis on the importance of maintenance,1 or it could be through a focus on the importance of an expanded, nonhumanistic conception of care.2 Caring for living things should begin by recognizing the importance of holding heterogeneous elements together, before rushing to reduce them to instrumental ends or synthesize them into a new type of function. As William Bateson wrote and Ed Lewis frequently quoted: “So long as systematic experiments in breeding are wanting, and so long as the attention of naturalists is limited to the study of normal forms, in this part of biology which is perhaps of greater theoretical and even practical importance than any other, there can be no progress.”3 Weirdness does as each weirdness is; its value is that it can’t be fully anticipated. Studying biology should also mean cultivating more creative and less oppressive approaches to studying lives in the grid.

This page intentionally left blank

Acknowledgments A project that gestates for as long as Biology in the Grid develops numerous caregivers. Some of these caregivers actively participated in shaping the pages that follow; some provided needed catalysis, reflection, and perspective. All have unselfishly supported me in some way. Those who read the manuscript have, perhaps, sacrificed the most. Academic writing in general, and my writing in particular, takes a bit of time to dress up and make presentable. These folks braved my prose at early stages and need to be especially congratulated. My longtime friend and collaborator Robert Mitchell wrestled with preliminary chapters and helped me, as always, to slow down and clearly argue a point and provide enough examples to make my arguments convincing. His friendship and astute eye continue to enrich my life and work. My dear friend Dakota Gearhart took the time to share ideas with me at formative stages of the project. From her I learned to value the weird and unexpected and to pursue an artistic vision with integrity, imagination, and humor. My friend and one-­time student Adam Nocek provided a unique blend of sympathetic collaborator and catalyst to further the scope of my research. Good students (meaning smart and caring as well as challenging) are always the best teachers. And speaking of good students and good friends, Nathaniel Mengist read all or parts of every chapter and gave me his unique and valuable insights throughout the entire manuscript. I especially owe to him my increasing reliance on en dashes as well as many wonderful moments discussing important topics over beverages. Kanna Hudson, who helped me on my first manuscript, had the courage to once again enter the fray and help me check my references (which I find particularly tedious) and improve my prose. Francesca Hillery helped me identify key sources and was an important sounding board for ideas. Cary Wolfe especially provided insight and perspective as I developed this manuscript. Doug Armato offered the keen eye that years of directing an innovative academic press can contribute. Karl Benitez used his fertile and creative mind to help me illustrate a few tricky points with clarity and simplicity. William 215

216 · Acknowledgments

Rorabaugh read an early draft of the complete manuscript and gave me much needed encouragement and advice. And, although I never forced her to read actual pages of the manuscript, I did often accost my daughter, Kamla Thurtle, with hastily read passages in my need for feedback. She always helped me with her uniquely honed critical judgment and sensitive wisdom. Two forums at the University of Washington were especially important for refining early drafts of the material. Elizabeth Porter and Ryan Callo invited me to present the content of chapter 1 at the Legal Methods Workshop, and Jenna Grant invited me to present chapter 2 to her class on Visuality and Medicine. The Simpson Center for the Humanities supported me with a Society of Scholars fellowship that allowed me to present material that would later become parts of chapter 5. The members of the Ontogenesis Workshop in Santa Fe (Stuart Kaufmann, Gaymon Bennett, Cary Wolfe, Adam Nocek, Michael Epperson, and Sha Xin Wei) patiently listened to an early draft of chapter 4. Arthur and Mary Louise Kroker invited me to present a very early vision of part of the project at the Pacific Centre for Technology and Culture. Claudia X. Valdes invited me to present the formative work for chapter 5 at the University of New Mexico. Sarah Street gave me an opportunity to present some of the work from chapters 2 and 5 at a conference on Intermediality at Bristol University. Even more individuals contributed to this book by working with me in ways that have transformed what I think about and why. These include (in no particular order) Bruno Clarke, Rebecca Cummins, Tyler Fox, Brian Reed, Caroline Simpson, Cynthia Anderson, Laura Harrison Marquez, Nara Hohensee, Rodrigo Valenzuela, Erin Clowes, Nick Bar-­Clingan, Chandan Reddy, María Elena García, Suzanne St. Peter, Stacey Moran Nocek, Rahul Gairolia, Third Andresen, Jeanette Bushnell, Orit Halperin, Ron Broglio, Christina Wygant, Nancy White Iff, Amy Peloff, Joel Ong, Stephanie Andrews, Claudia X. Valdes, David Ribes, Leah Ceccarelli, Alison Wiley, Robert Brain, Adam Frank, Elizabeth Wilson, Timothy Lenoir, Kate Hayles, George Behlmer, Anand Yang, Lynn Thomas, Tyler Volk, Maynard Olson, Christophe Bonneuil, Jean-­Paul Gaudilliere, Keith Harris, Henning Schmidgen, Jens Hauser, Louise Whiteley, Anu Taranath, Annie Dwyer, Terry Schenold, Sarah Kremen-­Hicks, Charles Richter, Uma Malhotra, Erik Parr, Joh Howard, Gabriel de los Angeles, Tony Lucero, Matt Sparke, Melissa Liu, Adam Zaretsky, Elizabeth Busch­mann, Stephanie Andrews, Steven Oscherwitz, the Molecular Shadows Salon,

Acknowledgments · 217

Stephanie Maxwell, Joe Davis, Tim Cahill, Finnbogi Petersson, Martin Jarmick, Theron Stevenson, Vic Aque, Axel Roesler, Carrie Bodle, Allison Kudla, Jentery Sayers, Genevieve Tremblay, Vicente Rafael, Ed Taylor, Ginny Ruffner, Eleanor Toews, Rich Doyle, Kristian Blak, David Giles, Wendy Wiseman, Susan Squier, Giorgia Aiello, John Vallier and the Media Arcade, Elizabeth Schiffler, Wanda Gregory, Marta Lyall, John Palios, Mr. White, and Sophie. I’ve had incredible students and can’t list them all. Thank you for educating me. Two folks deserve special thanks for making my time at the University of Washington productive and enjoyable. Kathleen Woodward, the director of the Simpson Center for the Humanities, created a world-­class center for the study of the humanities on the UW campus and I’m grateful to have been a participant. John Toews, longtime director for the Comparative History of Ideas program, provided the most treasured of all resources: an intellectual home. Finally, a heartfelt toast to the Big Time Brewery and its remarkable cast of characters, including but not limited to Rick, Margaret, Sara, and Roy.

This page intentionally left blank

Notes Introduction 1. There is a robust literature on world building in fiction, see for instance the work of Marleen S. Barr, such as Feminist Fabulation: Space/Postmodern Fiction (Iowa City: University of Iowa Press, 1992), or more recently, World Building: Discourse in the Mind, ed. Joanna Gavins and Ernestine Lahey (London: Bloomsbury Press, 2016). The concept is also used in the critical gaming literature as well. 2. A good example of this approach is Robert H. Carlson, Biology Is Technology: The Promise, Peril, and New Business of Engineering Life (Cambridge, Mass.: Harvard University Press, 2011). 3. A good example of this is the textbook, Sean B. Carroll, Jennifer K. Grenier, and Scott D. Weatherbee, From DNA to Diversity: Molecular Genetics and the Evolution of Animal Design (New York: Blackwell Publishing, 2001). 4. The quotations are taken from the website for the Institute for Systems Biology, https://www.systemsbiology.org/. 5. I will talk about these histories at greater length later. Two popular approaches, however, have been Eugene Thacker’s influential Biomedia (Minneapolis: University of Minnesota Press, 2004), and Lily E. Kay, Who Wrote the Book of Life? A History of the Genetic Code (Stanford, Calif.: Stanford University Press, 2000). For a recent theoretical account see Alexander R. Galloway, The Interface Effect (Cambridge, UK: Polity, 2013). It is also important to note how research on the history of cybernetics has transformed the way we think of cybernetics as well. Current researchers emphasize how cybernetics is and has been noteworthy for its relationship to diverse disciplines and research programs and not just for a narrowly instrumental purpose. See for instance, Orit Halpern, Beautiful Data: A History of Vision and Reason since 1945 (Durham, N.C.: Duke University Press, 2014); Rheingold Martin, The Organizational Complex: Architecture, Media, and Corporate Space (Cambridge, Mass.: MIT Press, 2003); and Bruce Clarke, Neo­ cybernetics and Narrative (Minneapolis: University of Minnesota, 2014). The literature reevalu­ating the role of systems theory in the arts is also relevant. See Frances Halsall, Systems of Art: Art, History, and Systems Theory (New York: Peter Lang, 2008); and Edward Shanken, Systems (Cambridge, Mass.: MIT Press, 2015). 6. See Berris Charnley, “Geneticists on the Farm: Agriculture and the All-­ English Loaf,” in Scientific Governance in Britain, 1914–­1979, eds. Charlotte Sleigh 219

220 · NOTES TO INTRODUCTION

and Don Leggett (Manchester, UK: Manchester University Press, 2016), 181–­98; Berris Charnley, “Experiments in Empire-­Building: Mendelian Genetics as a National, Imperial, and Global Agricultural Enterprise,” Studies in the History and Philosophy of Science: Part A 44, no.2 (2013): 292–­300; Dominic Berry, “The Plant Breeding Industry after Pure Line Theory: Lessons from the National Institute of Agricultural Botany,” in Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences 46 (2014): 25–­37; Alain Pottage and Brad Sherman, “Organisms and Manufactures: On the History of Plant Inventions,” Melbourne University Law Review 31, no. 2 (2007): 539–­68; Christophe Bonneuil, “Mendelism, Plant Breeding and Experimental Cultures: Agriculture and the Development of Genetics in France,” Journal of the History of Biology 39, no. 2 (2006): 281–­308; Christophe Bonneuil, “Producing Identity, Industrializing Purity: Elements for a Cultural History of Genetics,” in A Cultural History of Heredity IV: Heredity in the Century of the Gene, Preprint 343, eds. Staffan Müller-­Wille, Hans-­Jörg Rheinberger, and John Dupré (Berlin: Max Planck Institute for the History of Science, 2008); Helen Anne Curry, “Industrial Evolution: Mechanical and Biological Innovation at the General Electric Research Laboratory,” Technology and Culture 54, no. 4 (2013): 746–­81; and Phillip Thurtle, The Emergence of Genetic Rationality: Space, Time, & Information in American Biological Science, 1870–­1920 (Seattle: University of Washington Press, 2008). 7. See Ash Amin, Post-­Fordism: A Reader (Hoboken. N.J.: Blackwell Publishing, 1994) for key essays. Also important is the historical literature on consumption and changes in information practices. In the work that follows two authors have been especially useful: James Beniger, The Control Revolution: Technological and Economic Origins of the Information Society (Cambridge, Mass.: Harvard University Press, 1989); and Susan Strasser, Satisfaction Guaranteed: The Making of the American Mass Market (Washington, D. C.: The Smithsonian, 2004). Also useful were Roberta Sassatelli, Consumer Culture: History, Theory and Politics (London: Sage, 2007); and Jackson Lear, Fables of Abundance: A Cultural History of Advertising in America (New York: BasicBooks, 1994). 8. Guy Debord, The Society of the Spectacle (New York: Zone Books, 1995). 9. See Hannah Higgins’s important monograph on the history of grids in art and design, The Grid Book (Cambridge. Mass.: MIT Press, 2009). 10. See for instance the discussion of modularity in Gerhard Schlosser and Günter P. Wagner’s, “The Modularity Concept in Developmental and Evolutionary Biology,” found in Modularity in Development and Evolution, ed. Gerhard Schlosser and Günter P. Wagner (Chicago: University of Chicago Press, 2004). 11. Michel Foucault, Security, Territory, and Population: Lectures at the Collége de France, 1977–­1978, ed. Michel Senellart, trans. Graham Burchell (New York: Picador, 2004), 73.

NOTES TO INTRODUCTION · 221

12. Foucault, 73. 13. “I do not mean to say that the law fades into the background or that the institutions of justice tend to disappear, but rather that the law operates more and more as a norm, and that the judicial institution is increasingly incorporated into a continuum of apparatuses (medical, administrative, and so on) whose functions are for the most part regulatory.” Michel Foucault, The History of Sexuality Volume 1: An Introduction, trans. Robert Hurley (New York: Random House, 2000), 144. 14. Eric H. Davidson, The Regulatory Genome: Gene Regulatory Networks in Development and Evolution (London: Academic Press, 2006), 2. 15. For a slightly dated compendium of regulatory mechanisms in relationship to human diseases see, Gene Regulatory Sequences and Human Disease, ed. Nadav Ahituv (New York: Springer, 2012). 16. Perhaps this is most straightforwardly articulated by Ron Amundson in The Changing Role of the Embryo in Evolutionary Thought: Roots of Evo-­Devo (Cambridge, Mass.: Cambridge University Press, 2007), especially pages 213–­58. 17. It is also important to recognize that although there are resemblances to older more tailored ways of providing clothing that there are still serious limitations on the development of these mannequins for purchasing clothes. For instance, many models still heavily rely on user-generated measurements (and thus are vulnerable to user-generated errors) and the complexity of some of the systems lead to low adoption rates. More interesting is the subjective notion of fit and how that relates to the experience of wearing clothes. A quick but older overview can be found here: Lisa Wang, “Can Technology Solve the Fit Problem in Fashion E-­Commerce?,” The Business of Fashion, August 12, 2014, https://www.businessoffashion.com/articles/fashion-tech/ can-technology-solve-fit-problem-fashion-e-commerce. 18. Garland Allen’s essay on the dialectical nature of biological development acknowledges this. See Garland E. Allan, “A Century of Evo-­Devo: The Dialectics of Analysis and Synthesis in Twentieth Century Life Science,” in From Embryology to Evo-­Devo: A History of Developmental Evolution, ed. Manfred Laubichler and Jane Maienschein (Cambridge, Mass.: MIT Press), 123–­68. The expanded concept of regulation, however, also helps me see how this one type of synthesis in understanding is related to other syntheses in other fields. 19. A secondary literature on Vilém Flusser has been slow in accretion. Especially important for me was Anke Finger, Rainer Guldin, and Gustavo Bernardo, Vilém Flusser: An Introduction (Minneapolis: University of Minnesota Press, 2011); and Flusseriana: An Intellectual Toolbox, eds. Siegfried Zielinski, Peter Weibel, and Daniel Irrgang (Minneapolis, Minn.: Univocal Publishing, 2015). 20. Vilém Flusser, Writings, ed. Andreas Strohl, trans. Erik Eisel (Minneapolis: University of Minnesota Press, 2002), 22–­23.

222 · NOTES TO INTRODUCTION

21. Flusser, 23. 22. For an excellent review, see Elizabeth Porter, “Taking Images Seriously,” Columbia Law Review 114, no. 7 (2014) 1687–­782. 23. Vilém Flusser, Towards a Philosophy of Photography, trans. Anthony Matthews (London: Reaktion Books, 2000), 14. 24. Flusser, 14. 25. See for instance, Nigel Thrift, Non-­Representational Theory: Space, Politics, Affect (London: Routledge, 2008). 26. Vilém Flusser, Into the Universe of Technical Images, trans. Nancy Ann Roth (Minneapolis: University of Minnesota Press, 2011), 21. 27. For detailed information see the Society for Developmental Biology’s online database, https://www.sdbonline.org/sites/fly/segment/evenskp1.htm. 28. http://www.sdbonline.org/sites/fly/segment/kruppel1.htm. 29. Flusser, Into the Universe of Technical Images, 21. 30. See, for the masterworks by Marshall McLuhan: Marshall McLuhan, Understanding Media: The Extensions of Man (Cambridge, Mass.: MIT Press, 1994); and Marshall McLuhan and Quentin Fiore, The Medium Is the Massage: An Inventory of Effects (Berkeley, Calif.: Gingko Press, 2001). 31. See, for instance, Jean Baudrillard’s writings on the hyperreal. Two references are especially important. The first is the most referenced essay on the hyperreal: “The Precession of the Simulacrum” from Simulacrum and Simulation, trans. Sheila Fraser Glaser (Ann Arbor: University of Michigan Press, 1996), 1–­42. The book where he outlines his ideas on the hyperreal in relationship to his ideas on symbolic exchange is Jean Baudrillard, Symbolic Exchange and Death, trans. Iain Hamilton Grant (London: Sage Publications, 1993), especially the chapter on “The Order of Simulacra” (50–­86). 32. Alexander Galloway, Protocol: How Control Exists after Decentralization (Cambridge, Mass.: MIT Press, 2004). 33. Michel Foucault, Society Must Be Defended: Lectures at the Collège de France, 1975–­1976 (New York: Picador, 2003), 241. 34. Sean Carroll, Endless Forms Most Beautiful: The New Science of Evo Devo (New York: W. W. Norton & Co., 2006), 100. 35. Thomas Lamarre, The Anime Machine: A Media Theory of Animation (Minneapolis: University of Minnesota Press, 2009). 36. See Johann Wolfgang von Goethe and Gordon L. Miller, The Metamorphosis of Plants (Cambridge, Mass.: MIT Press, 2009). A key piece in the history of biology that frames biological research in terms of forms in cell theory and individual development is William Coleman, Biology in the Nineteenth Century: Problems of Form, Function and Transformation (Cambridge: Cambridge University Press, 1978). The important inspiration for Coleman’s work is E. S. Russell, Form and Function: A Contribution to the History of Animal Morphology (London:

NOTES TO INTRODUCTION · 223

John Murray, 1916). There are numerous more recent studies preoccupied with this issue and I will cover these in greater depth in chapter 1, “Life on the Line.” 37. Richard Burian, “Reconceiving Animals and Their Evolution: On Some Consequences of New Research on the Modularity of Development and Evolution,” from The Epistemology of Development, Evolution, and Genetics: Selected Essays (Cambridge: Cambridge University Press, 2005), 258. 38. Gilbert Simondon, L’individuation psychique et collective  (Paris: Aubier, 1989; reprinted in 2007). 39. Gilbert Simondon, “The Genesis of the Individual,” from Incorporations, eds. Jonathan Crary and Sanford Kwinter (New York: Zone Books, 1992): 297–­319. A good secondary source is Muriel Coombs, Gilbert Simondon and the Philosophy of the Transindividual, trans. Thomas Lamarre (Cambridge, Mass.: MIT Press, 2012). See also Tyler Fox’s analysis, “The Genesis of the Individual” from http:// www.tylersfox.com/blog/127. 40. Alireza Iranbakhsh and Seyyed Hassan Seyyedrezaei, “The Impact of Information Technology in Biological Sciences.” Procedia Computer Science 3 (2011): 913–­16. 41. Thacker, Biomedia. 42. Kay, Who Wrote the Book of Life. 43. Robert S. Ledley, “Digital Electronic Computers in Biomedical Science,” Science 130, no. 3384 (1959): 1225–­34. 44. J. C. Kendrew, et al. “Structure of Myoglobin: A Three-­Dimensional Fourier Synthesis at 2 Å Resolution,” Nature 185, no. 4711 (1960): 422–­27. 45. Ledley, “Digital Electronic Computers in Biomedical Science,” 1226. 46. An emphasis on work is found in many branches of sociology, but it is especially prevalent in the fields of ethnographic laboratory analysis and in the Tremont Institute’s emphasis on pragmatics. The key reference for ethnographies of laboratory practice is Bruno Latour’s and Steven Woolgar’s influential Laboratory Life (Beverly Hills, Calif.: Sage, 1979). A key reference for the sociology of work as a form of social pragmatics is Elihu Gerson’s tone-setting 1983 “Scientific Work and Social Worlds,” Knowledge 4, no. 35. 47. Geoffrey C. Bowker and Susan Leigh Star, Sorting Things Out: Classification and Its Consequences (Cambridge, Mass.: MIT Press, 2000). 48. This is an emphasis of Thacker’s, Biomedia. The quotation is from Eugene Thacker, “What is Biomedia?,” Configurations 11.1 (2003): 52. 49. Kay, Who Wrote the Book of Life? 50. See Walter B. Cannon, The Wisdom of the Body (New York: W. W. Norton & Company, 1932). 51. Norbert Weiner, Cybernetics or Control and Communication in the Animal and the Machine, Second Edition (Cambridge, Mass.: MIT Press, 1961), 115. 52. M. W. Nirenberg and H. J. Matthaei, “The Dependence of Cell-­Free Protein

224 · NOTES TO INTRODUCTION

Synthesis in E. coli upon Naturally Occurring or Synthetic Polyribonucleotides,” Proceedings of the National Academy of Sciences of the United States of America 47, no. 10 (1961): 1588–­602. See also, Kay, Who Wrote the Book of Life?, 234 and 235–­56. 53. F. Jacob and J. Monod, “Genetic Regulatory Mechanisms in the Synthesis of Proteins,” Journal of Molecular Biology 3 (1961): 318–­56. 54. See Horace Freeland Judson, The Eighth Day of Creation: The Makers of the Revolution of Biology (New York: Simon and Schuster, 1979), 404–­12. Also see Kay, Who Wrote the Book of Life?, 209–­13. As Evelyn Fox Keller notes, even Thomas Morgan recognized that development required genes to act in different ways. See Evelyn Fox Keller, The Century of the Gene (Cambridge. Mass.: Harvard University Press, 2002), 56. 55. I’m thinking here of Gilles Deleuze’s influential but much too abbreviated essay, “Postscript on the Societies of Control,” October 59 (Winter 1992): 3–­7. Also important is how some have used this article in contemporary critique. See Galloway, The Interface Effect. 56. Lamarre, The Anime Machine, 26–­44.

1. Life on the Line 1. Although these arguments are quite old, you can find recent examples in Eugene Thacker, After Life (Chicago: University of Chicago Press, 2010): and Stefan Helmreich, “What Was Life: Answers from Three Limit Biologies,” Critical Inquiry 37, no. 4 (Summer 2011). 2. Stan Lee and John Buscema, How to Draw Comics the Marvel Way (New York: Touchstone, 1984). 3. Some of the most important of these for me include: Svetlana Alpers, The Art of Describing: Dutch Art in the Seventeenth Century (Chicago: University of Chicago Press, 1983); Linda Dalrymple Henderson, The Fourth Dimension and Non-­Euclidean Geometry in Modern Art (Princeton, N.J.: Princeton University Press, 1983); Martin Kemp, The Science of Art: Optical Themes in Western Art from Brunelleschi to Seurat, reprint ed. (New Haven, Conn.: Yale University Press, 1992); Barbara Maria Stafford, Voyage into Substance: Art, Science, Nature, and the Illustrated Travel Account, 1760–­1840 (Cambridge, Mass.: MIT Press, 1984). 4. See Lorraine Daston and Peter Galison, Objectivity (Cambridge, Mass.: Zone Books, 2007); Michael Lynch and Steve Woolgar, Representation in Scientific Practice (Cambridge, Mass.: MIT Press, 1990); Brian Rotman, Signifying Nothing: The Semiotics of Zero (Houndmills, UK: Macmillan, 1987); Catelijne Coopmans, Representation in Scientific Practice Revisited (Cambridge, Mass.: MIT Press, 2014); Luc Pauwels, ed., Visual Cultures of Science: Rethinking Representational Practices in Knowledge Building and Science Communication (Hanover, N.H.: Dartmouth College Press, 2005).

NOTES TO CHAPTER 1 · 225

5. See Pamela Kort, Max Hollein, and Schirn Kunsthalle Frankfurt, Darwin: Art and the Search for Origins (Cologne: Wienand, 2009); Robert Brain, The Pulse of Modernism: Physiological Aesthetics in Fin-­de-­Siècle Europe (Seattle: University of Washington Press, 2015); Nancy Anderson and Michael R. Dietrich, eds., The Educated Eye: Visual Culture and Pedagogy in the Life Sciences (Hanover, N.H.: Dartmouth College Press, 2012); Hannah Landecker, “Microcinematography and the History of Science and Film,” ISIS: Journal of the History of Science in Society 97, no. 1 (2006): 121–­32; Martin J. S. Rudwick, Scenes from Deep Time: Early Pictorial Representations of the Prehistoric World (Chicago: University of Chicago Press, 1995); Janice Neri, The Insect and the Image: Visualizing Nature in Early Modern Europe, 1500–­1700 (Minneapolis: University of Minnesota Press, 2011); Sachiko Kusukawa, Picturing the Book of Nature: Image, Text, and Argument in Sixteenth-­ Century Human Anatomy and Medical Botany (Chicago: University of Chicago Press, 2011). 6. See Stephen S. Hall, Mapping the Next Millennium: The Discovery of New Geographies (New York: Random House, 1992); Dorothy Nelkin and M. Susan Lindee, The DNA Mystique: The Gene as a Cultural Icon (New York: Freeman, 1995); Suzanne Anker and Dorothy Nelkin, The Molecular Gaze: Art in the Genetic Age (Cold Spring Harbor, N.Y.: Cold Spring Harbor Laboratory Press, 2003); Robert Mitchell, Bioart and the Vitality of Media (Seattle: University of Washington Press, 2010). 7. This appears not only in biology, it is a basic type of tensions identified with Aristotelian ideas of hylomorphism. See for instance Aristotle’s description of the role of matter, form, and privation in change, Physics, 192a and b. Aristotle, The Works of Aristotle: Physica, Vol. II, trans. R. P. Hardie and R. K. Gaye (Oxford: Clarendon Press, 1930), 192a and b. 8. Goethe’s appreciation of Kant seems to remain limited to the Critique of Judgment. The primary English source for this claim is Karl J. Fink’s, Goethe’s History of Science (Cambridge: Cambridge University Press, 1991), 77–­79. Especially important for Fink’s conclusions are the letters between Goethe and Schiller. 9. Henri Focillon, The Life of Forms in Art, trans. George Kubler (Cambridge, Mass.: Zone Books, 1992), 41. 10. Jean Molino, “Introduction” to Henri Focillon, The Life of Forms in Art, trans. Kubler (Cambridge: Zone Books, 1992), 14. 11. Deleuze and Guattari write about Worringer’s emphasis on abstraction: “It is Worringer who accorded fundamental importance to the abstract line, seeing it as the very beginning of art or the first expression of an artistic will. Art as abstract machine.” Gilles Deleuze and Félix Guattari, A Thousand Plateaus: Capitalism and Schizophrenia, trans. Brian Massumi (Minneapolis: University of Minnesota Press, 1987), 496. The Dictionary of Art Historians goes so far as to state that “Worringer’s contribution to art and art history cannot be underestimated.” Max Dvorák,

226 · NOTES TO CHAPTER 1

“Worringer, Wilhelm,” Dictionary of Art Historians, https://dictionaryofart historians.org/worringerw. 12. Wilhelm Worringer, Abstraction and Empathy, trans. Michael Bullock (Chicago: Ivan R Dee, 1997), 4. 13. Molino, 10. 14. Michael K. Richardson and Gerhard Keuck, “Haeckel’s ABC of Evolution and Development,” Biological Reviews of the Cambridge Philosophical Society 77, no. 4 (2002): 495–­528. 15. Ronald Rainger, An Agenda for Antiquity: Henry Fairfield Osborn and Vertebrate Paleontology at the American Museum of Natural History, 1890–­1935 (Tuscaloosa: University of Alabama Press, 1991), esp. 163–­77. I also comment on this in Phillip Thurtle, The Emergence of Genetic Rationality: Space, Time, and Information in American Biological Science, 1870–­1920 (Seattle: University of Washington Press, 2007), 203–­5. 16. Ernst Haeckel, The Wonders of Life: A Popular Study of Biological Philosophy, trans. Joseph McCabe (London: Watts, 1904), 171. 17. Haeckel, 170. 18. Haeckel, 170. 19. Ernst Haeckel, The History of Creation, 2 vols., trans. E. Ray Lankester (New York: D. Appleton, 1901), 1:83. 20. Gordon L. Miller argues, for instance, that this was one of the main differences between Goethe’s efforts at systematics and Linnaeus, who Goethe in general admired. Goethe sought a dynamic way to conceive of the relationship between forms and thus appealed to dynamic archetypes instead of “fixed forms and species.” See Gordon L. Miller, “Introduction,” from Johann Wolfgang von Goethe and Gordon L. Miller, The Metamorphosis of Plants (Cambridge, Mass.: MIT Press, 2009), xxiii. 21. Johann Wolfgang von Goethe and Gordon L. Miller, The Metamorphosis of Plants (Cambridge, Mass.: MIT Press, 2009), 102. 22. See for instance Marsha Morton’s observation of how the “flowing tentacles” of the medusa reminded Haeckel of the hair “of his first wife Anna Sethe.” Marsha Morton, “From Monera to Man: Ernst Haeckel, Darwinismus, and Nineteenth Century German Art” in Barbara Larson and Fae Brauer, The Art of Evolution: Darwin, Darwinisms, and Visual Culture (Hanover, N. H.: University Press of New England, 2009), 63. 23. Haeckel, The History of Creation, 1:22. 24. Ernst Haeckel, Kristallseelen: Studien über das anorganische Leben (Leipzig: Alfred Kröner Verlag, 1917). 25. Haeckel, The History of Creation, 1:404–­5. 26. Haeckel, 1:409–­10.

NOTES TO CHAPTER 1 · 227

27. Haeckel, 1:323. 28. See Harrington’s discussion on the importance of Kant and Goethe for German Holism, Anne Harrington, Reenchanted Science: Holism in German Culture from Wilhelm II to Hitler (Princeton, N.J.: Princeton University Press, 1996), 4–­7. 29. Thinkers with commitments as different as Gilles Deleuze and Andrew Ward read the third Critique in these terms. This is in opposition to another type of reading of the third Critique that conceives of it as Kant’s inability to think life because of his inability to apply determinative judgment to living things. See Eugene Thacker. “On Ascensionism” Inflexions 7, “Animating Biophilosophy” (March 2014): 1–­7, www.inflexions.org. Important as well has been Robert Mitchell and Jacques Khalip’s discussion of Kant and the romantic image in “Release—­ (Non)Origination—­Concepts” in Jacques Khalip and Robert Mitchell, Releasing the Image: From Literature to New Media (Stanford, Calif.: Stanford University Press, 2011). 30. Immanuel Kant, Critique of the Power of Judgment, ed. Paul Guyer, trans. Paul Guyer and Eric Matthews (Cambridge: Cambridge University Press, 2000), Introduction §IV, 5:179. 31. Kant, A132/B171. 32. Kant, FI 20:211. 33. Kant, 5:220, §10. 34. Kant, 5:220, §10. 35. Kant, §17, 236/84. 36. Robert Wicks, Kant on Judgment (London: Routledge, 2007), 53. 37. Kant, 5:422, §81 38. Steven Shaviro has recently argued for the importance of Kant’s treatment of the “beautiful” in relationship to the “sublime,” which has tended to dominate recent appeals to Kant’s aesthetics in literary studies. Steven Shaviro, Without Criteria: Kant, Whitehead, Deleuze, and Aesthetics (Cambridge, Mass.: MIT Press, 2009). 39. Kant, 5:422, §81. 40. Haeckel was seen as a materialist by many of the neo-­Kantians. Frederick C. Beiser, The Genesis of Neo-­Kantianism, 1796–­1880 (Oxford: Oxford University Press, 2014), 448–­50. 41. Haeckel, The History of Creation, 1:105. 42. See for instance, Frederick C. Beiser’s chapter “Encounter with Darwinism” in The Genesis of Neo-­Kantianism, 1796–­1880 (Oxford: Oxford University Press, 2014). 43. Haeckel, The History of Creation, 1:106. 44. Some of the key references here are Timothy Lenoir, Strategy of Life: Teleology and Mechanics in Nineteenth Century Biology (Chicago: University of Chicago Press, 1982); James G. Lennox, “Darwin Was a Teleologist,” Biology and Philosophy 8

228 · NOTES TO CHAPTER 1

(1993),  409–­21;  Francisco Ayala, “Teleological Explanations in Evolutionary Biology,”  Nature’s Purposes: Analyses of Function and Design in Biology (Cambridge, Mass.: MIT Press, 1998); Michael T. Ghiselin, Darwin’s Language May Seem Teleological but His Thinking Is Another Matter,” Biology and Philosophy  9 (1994): 489–­92; Ernst W. Mayr, “The Idea of Teleology,” Journal of the History of Ideas 53 (1992): 117–­35; Michael Ruse, “Kant and Evolution,” in Justin E. H. Smith, The Problem of Animal Generation in Early Modern Philosophy (Cambridge: Cambridge University Press, 2006), 402–­15; D. M. Walsh, “Organisms as Natural Purposes: The Contemporary Evolutionary Perspective,” Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences, 37 (2006): 771–­91; Marcel Quarfood, “Kant on Biological Teleology: Towards a Two-­ Level Interpretation,” Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences 37 (2006): 735–­47. 45. Robert Brain, “Protoplasmania: Huxley, Haeckel, and the Vibratory Organism in Late Nineteenth-­Century Science and Art,” in Barbara Larson and Fae Brauer, The Art of Evolution: Darwin, Darwinisms, and Visual Culture (Hanover, N. H.: University Press of New England, 2009). 46. Focillion, 34. 47. Olaf Breidbach, Visions of Nature: The Art and Science of Ernst Haeckel (Munich: Prestel, 2006), 25. 48. Focillion, 48. 49. Nick Hopwood, Haeckel’s Embryos: Images, Evolution, and Fraud (Chicago: University of Chicago Press, 2015). 50. The debate has been masterfully summarized in Nick Hopwood’s text cited above. 51. Ernst Haeckel, The Evolution of Man: A Popular Exposition of the Princi­ pal Points of Human Ontogeny and Phylogeny (New York: D. Appleton, 1879), xxxiv–xxxv. 52. Nick Hopwood, “Pictures of Evolution and Charges of Fraud: Ernst Haeckel’s Embryological Illustrations,” Isis 97 (2006): 265. 53. Niles R. Holt, “Monists & Nazis: A Question of Scientific Responsibility,” The Hastings Center Report 5, no. 2 (1975). The connection seems most viable, according to Heiner Fangerau, in the earlier stages of the Monist League. After World War I the values between the two groups “diverged at all levels” as the Monists concentrated more specifically on social environmentalism as opposed to strictly racial theories of social improvement. Also see Heiner Fangerau, “Monism, Racial Hygiene, and National Socialism,” in Monism: Science, Philosophy, Religion, and the History of a Worldview, ed. Todd Weir (New York: Palgrave, 2012), 237. 54. See for instance, Robert Richards, The Tragic Sense of Life (Chicago: University of Chicago Press, 2008), 269–­76.

NOTES TO CHAPTER 2 · 229

55. Anne Harrington, Reenchanted Science: Holism in German Culture from Wilhelm II to Hitler (Princeton, N.J.: Princeton University Press, 1999), 10. 56. Harrington, 188. 57. Sanford Kwinter, “Who’s Afraid of Formalism,” in Far From Equillibrium: Essays on Technology and Design Culture, ed. Cynthia Davidson (New York: Actar, 2008), 146. 58. Luciana Parisi, Contagious Architecture: Computation, Aesthetics, and Space (Cambridge, Mass.: MIT Press, 2013).

2. Envisioning Grids 1. Nick Hopwood, Haeckel’s Embryos: Images, Evolution, and Fraud (Chicago: University of Chicago Press, 2015), 5. See especially pages 73–­88 for the theoretical and practical considerations in changing from an illustration style utilizing a progression to one utilizing a grid that compares progressions. 2. Hopwood, 85. 3. Christiane Nüsslein-­Volhard and Eric Wieschaus, “Mutations Affecting Segment Number and Polarity in Drosophila,” Nature 287, no. 5785 (1980): 799. 4. Timothy Samara, Making and Breaking the Grid (Gloucester, Mass.: Rockport Publishers, 2005), 29. 5. Lewis I. Held Jr., How the Snake Lost Its Legs: Curious Tales from the Frontier of Evo-­Devo (Cambridge: Cambridge University Press, 2014), 2. 6. See for instance, Sean B. Carroll, “Homeotic Genes and the Evolution of Arthropods and Chordates,” Nature 376, no. 6540 (1995): 479–­85; and Hox Genes: Studies from the 20th and the 21st Century, ed. Jean S. Deutch (Austin, Tex.: Springer Science, 2010). 7. Grids in relationship to social control will be covered in greater depth in chapter 3. For now, suffice to say this way of viewing grids has a long genealogy in the literature on governmentality (the late Foucault, for instance) and critical analysis of the biological sciences. See for instance, Nik Rose, The Politics of Life Itself: Biomedicine, Power, and Subjectivity in the Twenty-­First Century (Princeton, N.J.: Princeton University Press, 2007). 8. “The problem is not that of being free, but of finding a way out, or even a way in, another side, a hallway, an adjacency.” Gilles Deleuze and Félix Guattari, Kafka: Toward a Minor Literature (Minneapolis: University of Minnesota Press, 1986), 7–­8. 9. See Caroline Levine, Forms: Whole, Rhythm, Hierarchy, Network (Prince­ ton, N.J.: Princeton University Press, 2015). 10. Vilém Flusser, Vilém Flusser: Writings, ed. Andreas Strohl, trans. Erik Eisel (Minneapolis: University of Minnesota Press, 2002), 22.

230 · NOTES TO CHAPTER 2

11. Flusser, 23. 12. Flusser, 23. 13. Vilém Flusser, Towards a Philosophy of Photography, trans. Anthony Matthews (London: Reaktion Books, 2000), 9. 14. Flusser, Writings, 28. 15. See for instance, Friedrich Kittler, “Perspective and the Book,” Grey Room 5 (2001): 39. 16. See for instance, Martin Kauffman’s interesting account on the proliferation of illustrated texts in a fourteenth-century English publication. M. Kauffmann, “Illustration and Ornament,” in The Cambridge History of the Book in Britain, ed. N. J. Morgan and R. M. Thomson (Cambridge: Cambridge University Press, 2008), 474–­87. 17. Armin Vit and Bryony Gomez-­Palacio, Graphic Design, Referenced: A Visual Guide to the Language, Applications, and History of Graphic Design (Gloucester, Mass.: Rockport Publishers, 2009), 50. 18. One might even be tempted to think about Flusser’s arguments in terms explored in chapter 1 regarding the role of judgment (as opposed to reason) in nature and art. 19. Josef Müller-­Brockmann, Grid Systems in Graphic Design (Zurich: Verlag Niggli, 1968), 7. 20. Massimo Vignelli, Grids: Their Meaning and Use for Federal Designers (Washington, D.C.: U.S. Government Printing Office, 1978), 4. 21. James Beniger, The Control Revolution: Technological and Economic Origins of the Information Society (Cambridge, Mass.: Harvard University Press, 1989); and Susan Strasser, Satisfaction Guaranteed: The Making of the American Mass Market (Washington, D.C.: The Smithsonian, 2004). Also important is Roberta Sassatelli, Consumer Culture: History, Theory and Politics (London: Sage, 2007); and Jackson Lear, Fables of Abundance: A Cultural History of Advertising in America (New York: BasicBooks, 1994). 22. See for instance, Alfred Chandler’s Scale and Scope: The Dynamics of Industrial Capitalism (New York: Belknap Press, 1994). 23. Milton Caniff in conversation with Will Eisner in Will Eisner, Will Eisner’s Shop Talk (Milwaukie, Ore.: Dark Horse Comics, 2001), 115. 24. Mike Benton, The Comic Book in America: An Illustrated History (Dallas, Tex.: Taylor Publishing, 1989), 14. 25. Benton, 15. 26. Benton, 15. 27. C. C. Beck talking to Will Eisner in Eisner, 67. 28. Art Spiegelman and Chip Kidd, Jack Cole and Plastic Man: Forms Stretched to Their Limits! (New York: DC Comics, 2001), 38.

NOTES TO CHAPTER 3 · 231

29. Marshall McLuhan, Understanding Media: The Extensions of Man (New York: McGraw-­Hill, 1964), 169. 30. Lears, Fables of Abundance, 20. 31. Martin Kemp, “Noticing Nature,” Nature 393, no. 25 (1998): 25. 32. Kemp, 26. 33. Daele Wolfle, “Science Changes,” Science 122, no. 3157 (1955): 7. 34. Vignelli, Grids, 15. 35. Bang Wong, “Points of View: Layout,” Nature Methods, 8, no. 10 (2011): 783. 36. Flusser, Towards a Philosophy of Photography, 14. 37. David Reed, The Popular Magazine in Britain and the United States (London: The British Library, 1997). 38. Flusser, Towards a Philosophy of Photography, 30. 39. See Alexander Galloway’s important Protocol: How Control Exists after Decentralization (Cambridge, Mass.: MIT Press, 2006). 40. Flusser, Towards a Philosophy of Photography, 66–­67. 41. Vilém Flusser, Into the Universe of Technical Images, trans. Nancy Ann Roth (Minneapolis: University of Minnesota Press), 38. 42. Flusser, 135. 43. Flusser, 16. 44. Jay David Bolter and Dianne Gromala, Windows and Mirrors: Interaction Design, Digital Art and the Myth of Transparency (Cambridge, Mass.: MIT Press, 2005). 45. Flusser, Towards a Philosophy of Photography, 15. 46. How mechanical images can contribute to a form of objective that removes an interpretive subject has famously been argued by Loraine Daston and Peter Galison in Objectivity (New York: Zone Books, 2010). 47. Erik Loyer, “Designer’s Statement” from “Totality for Kids,” Vectors 4, no. 1 (2013), http://vectors.usc.edu/projects/index.php?project=99&thread=Authors Statement. 48. “Understanding Comics,” http://scottmccloud.com/2-print/1-uc/.

3. Warped Grids 1. Steve Biodrowski, “The Fly (1958)—A Retrospective,” Cinefantastique 16, no. 3 and 4 (1986): 85. 2. On the changing role of movie advertising in the first half of the twentieth century, Garth S. Jowett, “Giving Them What They Want: Movie Audience Research before 1950,” in Current Research in Film: Audiences, Economics, and Law, Volume 1, ed. Bruce A. Austin (Norwood, N.J.: Ablex Publishing Corporation, 1985), 19. 3. George Langelaan, “The Fly,” Playboy, June 1957.

232 · NOTES TO CHAPTER 3

4. Luther S. West, The Housefly: Its Natural History, Medical Importance, and Control (Ithaca, N.Y.: Comstock Publishing Company, 1951), 424. 5. “Chapter 14: 1440: The Smooth and the Striated,” especially 474–­76, in Gilles Deleuze and Félix Guattari, A Thousand Plateaus: Capitalism and Schizophrenia, trans. Brian Massumi (Minneapolis: University of Minnesota Press, 1987). 6. Deleuze and Guattari, 478. 7. Helpful recent overviews on the issue of biopower include: Cary Wolfe, Before the Law: Humans and Other Animals in a Biopolitical Frame (Chicago: University of Chicago Press, 2012); Beyond Biopolitics: Essays in the Governance of Life and Truth, ed. Patricia Ticento Clough and Craig Willse (Durham, N.C.: Duke University Press, 2011); Bruce Braun, “Biopolitics and the Molecularization of Life,” Cultural Geographies 14 (2007): 6–­28; Thomas Lemke, Biopolitics an Advanced Introduction (New York: New York University, 2011); Tim Dean, “The Biopolitics of Pleasure”, The South Atlantic Quarterly 111, no. 13 (Summer 2012): 477–­96; Timothy Campbell and Adam Sitze, eds., Biopolitics: A Reader (Durham, N.C.: Duke University Press, 2013); Timothy C. Campbell, Improper Life: Biopolitics from Heidegger to Agamben (Minneapolis: University of Minnesota Press, 2011); Jeffrey T. Nealon, Plant Theory: Biopower and Vegetable Life (Stanford, Calif.: Stanford University Press, 2016); and Vanessa Lemm and Miguel Vatter, eds., The Government of Life: Foucault, Biopolitics, and Neoliberlism (New York: Fordham University Press, 2014). 8. Cyndy Hendershot offers a wonderful reading of gender and the Cold War fears found in The Fly (1958) in “The Cold War Horror Film: Taboo and Transgression in The Bad Seed, The Fly, and Psycho,” Journal of Popular Film and Television 29, no. 1 (2001): 20–­31, doi:10.1080/01956050109601006. 9. Lizabeth Cohen, A Consumer’s Republic: The Politics of Mass Consumption in Postwar America (New York: Vintage, 2003), 11. In the following analysis, I stick most closely to Lizabeth Cohen’s account of the rise of consumer culture for greater clarity. This was a fast-­growing field after the turn of the twenty-­first century and there are many insightful accounts, although few have the scope that Cohen’s book possesses. Besides the texts listed in chapter 2, note 21, other useful texts have included Antonella Caru and Bernard Cova, Consuming Experience (London: Routledge, 2013); Carolyn M. Goldstein, Creating Consumers: Home Economists in Twentieth-­Century America (Durham: University of North Carolina Press, 2012); Joanne Hollows, Domestic Cultures (London: McGraw-­Hill Education, 2008); Claude S. Fischer, Made in America: A Social History of American Culture and Character (Chicago: University of Chicago Press, 2010); Jenny Shaw, Shopping (Cambridge: Polity, 2010); and Bruce Pietrykowski, The Political Economy of Consumer Behavior: Contesting Consumption. (London: Routledge, 2009). 10. See chapter 5: “Residence: Inequality in Mass Suburbia” in Cohen, A Consumer’s Republic.

NOTES TO CHAPTER 3 · 233

11. See chapter 7: “Culture: Segmenting the Mass” in Cohen. 12. The Fly, directed by Kurt Neumann (1958; Los Angeles, Calif.: 20th Century Fox Home Video, 2007), DVD. 13. Olen J. Earnest, “Star Wars: A Case Study of Motion Picture Marketing,” in Current Research in Film: Audience, Economics and Law, Volume 1, ed. Bruce A. Austin (Norwood, N.J.: Ablex Publishing Corporation, 1985), 4. 14. Michel Foucault, Discipline and Punish: The Birth of the Prison, trans. Alan Sheridan (New York: Random House, 1977), 145. 15. There is very deep literature on the relationship between industrialization and bodies. Some of the most useful for me include: Anson Rabinbach, The Human Motor: Energy, Fatigue, and the Origins of Modernity (Berkeley: University of California Press, 1990); Linda Nochlin, The Body in Pieces: The Fragment as a Metaphor of Modernity (London: Thames and Hudson, 2001); and David A. Mindell, Between Human and Machine: Feedback, Control, and Computing Before Cybernetics (Baltimore, Md.: Johns Hopkins University Press, 2001). 16. Chapter 11 from Michel Foucault, Society Must Be Defended: Lectures at the Collége de France, 1975–­1976, ed. Mauro Bertani and Alessandro Fontana, trans. David Macey (New York: Picador, 2003). 17. Foucault, 249. 18. Michel Foucault, Security, Territory, and Population: Lectures at the Collége de France, 1977–­1978, ed. Michel Senellart, trans. Graham Burchell (New York: Picador, 2009), 45. 19. Foucault, 47. 20. Foucault, 48–­49. 21. Foucault, 72. 22. Foucault, 73. 23. See, for instance, Brian Massumi, The Power at the End of the Economy (Durham, N.C.: Duke University Press, 2015); and John Protevi, Political Affect: Connecting the Social and the Somatic (Minneapolis: University of Minnesota Press, 2009). 24. See Robert Mitchell’s insightful, “Biopolitics and Population Aesthetics” South Atlantic Quarterly 115, no. 2 (2016): 367–­98 25. Michel Foucault, Discipline and Punish: The Birth of the Prison, trans. Alan Sheridan (New York: Random House, 1977), 55–­60. 26. Bernhard Siegert, for instance, astutely points out how grids lend spaces an address. See Bernhard Siegert, Cultural Techniques: Grids, Filters, Doors, and Other Articulations of the Real, trans. Geoffrey Winthrop-­Young (New York: Fordham University Press, 2015), especially 97–­120. The reference for the discussion of Foucault is the second chapter of Security, Territory, and Population, especially important is his discussion of grain scarcity in the lecture of “18 January 1978.”

234 · NOTES TO CHAPTER 3

27. I thank Adam Nocek for suggesting these sources. 28. Kisho Kurokawa, Each One a Hero: The Philosophy of Symbiosis (New York: Kodansha America, Inc., 1997), 123. 29. Salvator-­John Liotta and Matteo Belfiore, Patterns and Layering: Japanese Spatial Culture, Nature, and Architecture (Berlin: Gestalten, 2012), 14. My goal in discussing Japanese architecture is not to make an essentialist distinction between Western and Eastern architecture. Rather, I want to point to different ways of imagining and engaging with space. The spatial dynamics are a convenient point as they are often theorized in opposition to a Western organicist idea of space. I concentrate on architecture in this section but investigate animated surfaces in the same way later on. Blaine Brownwell’s Matter in the Floating World: Conversations with Leading Japanese Architects and Designers (New York: Princeton Architectural Press, 2011) has also been useful. Kisho Kurokawa’s Each One a Hero is especially useful for his distinctions between Western and Japanese conceptions of space. 30. Salvator-­John Liotta, “Patterns, Japanese Spatial Culture, Nature, and Generative Design” in Liotta and Belfiore, 12. 31. Steven Connor, Fly (London: Reaktion Books, 2006), 54. 32. Robert Kohler, Lords of the Fly: Drosophila Genetics and the Experimental Life (Chicago: University of Chicago Press, 2004). 33. This passage was inspired by the wonderful book by Vilém Flusser and Louis Bec, Vampyroteuthis Infernalis: A Treatise, with a Report by the Institut Scientifique de Recherche Paranaturaliste, trans. Valentine A. Pakis (Minneapolis, University of Minnesota Press, 2012). 34. The House Fly, Encyclopedia Britannica Films, 1958, YouTube video, 15:40, http://www.youtube.com/watch?v=tZCvFmh6VxE. 35. West, The Housefly: Its Natural History, Medical Importance, and Control, 92. 36. West, 353. 37. West, x. 38. Curt Stern, “Two or Three Bristles,” American Scientist 42, no. 2 (1954): 213. 39. Stern, 211. 40. Stern, 213–­4. 41. Bryan Shorrocks, Drosophila (Invertebrate Types), (Cambridge, Mass.: Ginn & Company, 1972), 11. 42. Stern, “Two or Three Bristles,” 214. 43. The Fly, directed by Kurt Neumann (1958; Los Angeles, Calif.: 20th Century Fox Home Video, 2007), DVD. 44. For a wonderful reading comparing the 1958 and 1986 versions in terms of communications and cybernetics see Bruce Clarke, “Mediating The Fly: Post­ human Metamorphosis in the 1950s,” Configurations 10, no. 1 (2002): 169–­91. 45. Steve Biodrowski, “The Fly (1958)—­A Retrospective,” Cinefantastique  16,

NOTES TO CHAPTER 3 · 235

no. 3 and 4 (1986): 85. See also Anthony Timpone, Men, Makeup, and Monsters: Hollywood’s Masters of Illusion and FX (New York: St. Martin’s Press, 1996). 46. Anthony Timpone, “David Cronenberg: Lord of the Fly, Part One,” Fangoria 56 (August 1986): 21–­23. 47. J. Craig Venter et al., “The Sequence of the Human Genome,” Science 291, no. 5507 (2001): 1304–­51. 48. Venter et al., 1346. 49. Venter et al., 1346. 50. National Human Genome Research Institute, “Background of Comparative Genomic Analysis,” http://www.genome.gov/10005835. 51. Bateson’s term for this type of variation was actually, “homoeosis.” William Bateson, Materials for the Study of Variation (London: MacMillan and Company, 1894), 85. 52. Bateson, 85. 53. For a detailed account of the debates between Richard Goldschmidt and Sewall Wright on the role of discontinuity in evolution, see Michael R. Dietrich, “From Hopeful Monsters to Homeotic Effects: Richard Goldschmidt’s Integration of Development, Evolution, and Genetics,” American Zoologist 40, no. 5 (2000): 738–­47. 54. Clapperton Chakanetsa Mavhunga, “Vermin Beings: On Pestiferous Animals and Human Game,” Social Text 106 (Spring 2011): 163. 55. This extrapolation is purely my own. It should be noted that in this article Mavhunga is interested in thinking about the regulation of bodies in an explicitly post-­colonial as opposed to purely biopolitical framework. 56. Roberto Esposito, Bíos: Biopolitics and Philosophy (Minneapolis: University of Minnesota Press, 2008), especially the discussion in chapter 2. All three volumes of Esposito’s work on biopolitics are now available as English translations. See also, Communitas: The Origin and Destiny of Community (Stanford, Calif.: Stanford University Press, 2009), and Immunitas: The Protection and Negation of Life (London: Polity Press, 2011). 57. Esposito, Bíos, 52. 58. Giorgio Agamben, Homo Sacer: Sovereign Power and Bare Life, trans. Daniel Heller-­Roazen (Stanford, Calif.: Stanford University Press, 1998), 1. 59. Agamben, 6. 60. Agamben, 8. 61. See Thomas Wall’s, Radical Passivity: Levinas, Blanchot, and Agamben (Albany: State University of New York Press, 1999), for an extended discussion of how Agamben’s philosophy has been influenced by negative dialectics. 62. This is one of the points of emphasis for the treatment of biopower in

236 · NOTES TO CHAPTER 3

Michael Hardt and Antonio Negri’s, Empire (Cambridge, Mass.: Harvard University Press, 2001). 63. The most important case study here is, of course, Darwin’s reading of political economy and his adoption of the ideas of Adam Smith and Thomas Malthus. There are many important articles here. The classics are Silvan S. Schweber, “The Origin of the Origin Revisited,” Journal of the History of Biology 10, no. 2 (September 1977): 229–­316. G. M. Hodgson covers the application of evolutionary theory to economics in Economics and Evolution: Bringing Life Back into Economics (Ann Arbor: University of Michigan Press, 1993). Key for this type of thinking is Adrian Desmond’s work, especially, The Politics of Evolution: Morphology, Medicine, and Reform in Radical London (Chicago: University of Chicago Press, 1992). Recently, other scholars have developed theories on how deeply intertwined ideas of heredity are with economic thought. See for instance, Berris Charmley, “Experiments in Empire-­Building: Mendelian Genetics as a National, Imperial, and Global Agricultural Enterprise,” Studies in History and Philosophy of Science Part A 44, no. 2 (June 2013): 292–­300; and Christophe Bonneuil, “Mendelism, Plant Breeding and Experimental Cultures: Agriculture and the Development of Genetics in France,” Journal of the History of Biology 39, no. 2 (July 2006): 281–­308. For a background on the history of political economic thought in relationship to heredity see Staffan Müller-­Wille and Hans-­Jörg Rheinberger, A Cultural History of Heredity (Chicago: University of Chicago Press, 2012); and Phillip Thurtle, The Emergence of Genetic Rationality: Space, Time, and Information in American Biological Science, 1870–­1920 (Seattle: University of Washington Press, 2007). 64. See Phillip Thurtle and A. J. Nocek on “Vitalizing Thought” for the importance of “an exact” thinking for understanding biology, “Animating Biophilosophy,” Inflexions 7 (March 2014): i–­xi. www.inflexions.org. 65. Georges Canguilhem, “Introduction: The Role of Epistemology in Contemporary History of Science” in Ideology and Rationality in the History of the Life Sciences, trans. Arthur Goldhammer (Cambridge, Mass.: MIT Press, 1988), 9. 66. Henning Schmidgen, “Concepts Have a Life of Their Own: Biophilosophy, History and Structure” in Georges Canguilhem, “Animating Biophilosophy,” Inflexions 7 (March 2014): 62–­97, www.inflexions.org. Another recent scholar who has used Canguilhem to great effect in studying the treatment of substance abuse is Todd Meyers, The Clinic and Elsewhere: Addiction, Adolescents, and the Afterlife of Therapy (Seattle: University of Washington Press, 2013). 67. Interestingly, one of the most important concepts of biology is the concept of death. Understandably, there is a deep fear of comprehending the role of death in conceptions of life and vitality. The fear is that the vast scale of modern deaths would become understood as a bodily and hence natural process. This is, I think, a problem of dialectics and not death, per se. Not only is it important to recognize the role of death in life, it is also important to avoid holding death up as a negation

NOTES TO CHAPTER 3 · 237

of life. There are many types of deaths. The problem isn’t with death, per se, but with how death is thought to order society. 68. Schmidgen,“Concepts Have a Life of Their Own,” Inflexions 7. 69. In fact, this is how Canguilhem treats the concept of homeostasis, as a historical period in the tradition of thinking of regulation as a concept. See Georges Canguilhem, “The Development of the Concept of Biological Regulation in the Eighteenth and Nineteenth Centuries,” in Ideology and Rationality in the History of the Life Sciences, trans. Arthur Goldhammer (Cambridge, Mass.: MIT Press, 1988), 9. 70. Georges Canguilhem, The Normal and the Pathological, trans. Carolyn R. Fawcett (New York: Zone Books, 1991), see especially 115–­23. 71. Valerie Ahl and T. F. H. Allen, Hierarchy Theory: A Vision, Vocabulary, and Epistemology (New York: Columbia University Press), 148. 72. Jeffrey Nealon, Plant Theory: Biopower and Vegetable Life (Stanford, Calif.: Stanford University Press, 2016), xv. 73. Michel Serres, The Parasite, trans. Lawrence Schehr (Minneapolis: University of Minnesota Press, 2007). 74. Timothy Samara, Making and Breaking the Grid (Gloucester, Mass.: Rockport Publishers, 2005), 11. 75. See Gilles Deleuze and Félix Guattari, A Thousand Plateaus: Capitalism and Schizophrenia, trans. Brian Massumi (Minneapolis: University of Minnesota Press, 1987). See especially “Chapter 3: 10,000 B.C.: The Geology of Morals (Who Does the Earth Think It Is?)” as well as “Chapter 11: 1937: Of the Refrain.” Levi Bryant also has a useful blog entry that incorporates an interview of Deleuze on his use of assemblages. See “Deleuze on Assemblages” at https://larvalsubjects.wordpress. com/2009/10/08/deleuze-on-assemblages/. Manuel Delanda’s, A New Philosophy of Society: Assemblage Theory and Social Complexity (London: Bloomsbury Academic Press, 2006) is another popular application of assemblage theory to social collectives. 76. Alexander G. Weheliye, Habeas Viscous: Racializing Assemblages, Biopolitics, and Black Feminist Theories of the Human (Durham, N.C.: Duke University Press, 2014), 13. The problem of assemblages, however, are that they lack a way to think about the regularities of form. 77. Alessandra Raengo, Liquid Blackness: A Research Project on Blackness and Aesthetics, http://liquidblackness.com/about/. 78. Raengo. 79. Marty Roth focuses on the idea of doubling in relationship to the other in his article “Twice Two: ‘The Fly’ and ‘Invasion of the Body Snatchers,’ ” Discourse 22, no. 1 (2000): 103–­16. 80. Vilém Flusser, Into the Universe of Technical Images, trans. Nancy Ann Roth (Minneapolis: University of Minnesota Press), 31.

238 · NOTES TO CHAPTER 4

4. Modulations 1. Jessica A. Bolker, “Modularity in Development and Why It Matters to Evo-­ Devo,” American Zoologist 40, no. 5 (October 1, 2000): 775. 2. Manfred Laubichler, “Form and Function in Evo Devo: Historical and Conceptual Reflections,” in Form and Function in Developmental Evolution, ed. Manfred D. Laubichler and Jane Maienschein (Cambridge: Cambridge University Press, 2010), 30. 3. It is interesting to note that even much of the philosophical discourse suggesting that the use of modules has been overrated, still use a grid theory of development. The difference is that the grid is a product of a larger infrastructure and not just the product of an organism’s internal development. See, for instance, how the concept of “scaffolding” is deployed in Linda R. Caporael, James R. Griesemer, and William C. Wimsatt, eds. Developing Scaffolds in Evolution, Culture, and Cognition (Cambridge. Mass.: MIT Press, 2014). 4. Andrew L. Russell, “Modularity: An Interdisciplinary History of an Ordering Concept,” Information & Culture 47, no. 3 (2012): 259. 5. Gerhard Schlosser and Günter P. Wagner, “The Modularity Concept in Developmental and Evolutionary Biology,” in Modularity in Development and Evolution, ed. Gerhard Schlosser and Günter P. Wagner (Chicago: University of Chicago Press, 2004), 5. 6. Granted, most comic books tend to subsume many of the associative qualities to very linear types of narratives, but this is not always the case. Even in the most action driven types of comics, there are wonderful moments where the association between panels relies less on action-­to-­action transitions and more on other forms of associations between panels. See for instance, Scott McCloud’s useful discussion of panel transitions in Understanding Comics: The Invisible Art (Northampton, Mass.: Kitchen Sink Press, 1993), 60–­89. 7. Carliss Y. Baldwin and Kim B. Clark, Design Rules: The Power of Modularity: Volume 1 (Cambridge, Mass.: MIT Press, 2000). 8. Andrew L. Russell, “Modularity: An Interdisciplinary History of an Ordering Concept,” Information & Culture 47, no. 3 (2012): 257. 9. The key article here is François Jacob, “Evolution and Tinkering,” Science 196 (1977): 1161–6­6. 10. Schlosser and Wagner, “The Modularity Concept in Developmental and Evolutionary Biology,” 6. 11. Sanford Kwinter, “Who’s Afraid of Formalism,” in Far from Equilibrium: Essays on Technology and Design Culture, ed. Cynthia Davidson (New York: Actar, 2008, 146. 12. In fact, one final amendment to the example should be added and that is to suggest that the variations of these series happen because of the process of creating

NOTES TO CHAPTER 4 · 239

variance and not from a principle of design applied outside of the manufacture and play of the game. 13. Sean B. Carroll, Endless Forms Most Beautiful: The New Science of Evo Devo (New York: W. W. Norton & Company, 2006), 8. 14. Joseph Needham, “On the Dissociability of the Fundamental Processes in Ontogenesis,” Biological Reviews 8, no. 2 (1933): 180–­223. 15. Carroll, Endless Forms Most Beautiful, 100. 16. Some of the early criticisms are described in Beatrice Bateson, William Bateson, F.R.S., Naturalist: His Essays & Addresses, Together with a Short Account of His Life (Cambridge: Cambridge University Press, 1928). For a collection of different criticisms of Bateson’s work see the webpage created by Donald Forsdyke, “Opposition to Bateson,” http://post.queensu.ca/~forsdyke/bateson3.htm. 17. Donald Mackenzie, “Sociobiologies in Competition: The Biometrician-­ Mendelian Debate,” Biology, Medicine, and Society, 1840–­1940 (Cambridge: Cambridge University Press, 1981), 243–­88; and Patrick Bateson, “William Bateson: A Biologist Ahead of His Time,” Journal of Genetics 81, no. 2 (2002): 49–­58 (see the discussion on 51–­52). For a critique from a leading statistician see R. A. Fisher, The Genetical Theory of Natural Selection: A Complete Variorum Edition (Oxford: Oxford University Press, 1930), ix. 18. The key article here is William Coleman, “Bateson and Chromosomes: Conservative Thought in Science,” Centaurus 15, no. 3 (1971): 228–­314. But make sure to read the reevaluation by A. G. Cock, “William Bateson’s Rejection and Eventual Acceptance of Chromosome Theory,” Annals of Science 40, no. 1 (1983): 19–­59. 19. Ernst Mayr, “The Recent Historiography of Genetics,” Journal of the History of Biology 6, no. 1 (1973): 125–­54. 20. Mark Samuel Blumberg, Freaks of Nature: What Anomalies Tell Us about Development and Evolution (Oxford: Oxford University Press, 2009), 30–­37. 21. Carroll, Endless Forms Most Beautiful, 27, 45–­48. Sean B. Carroll, Jennifer K. Grenier, and Scott D. Weatherbee, From DNA to Diversity: Molecular Genetics and the Evolution of Animal Design (Malden, Mass.: Wiley-­Blackwell, 2004), 10. 22. Rudolf A. Raff, “Evo-­Devo: The Evolution of a New Discipline,” Nature Reviews Genetics 1, no. 1 (2000): 74–­79. 23. Lewis I. Held Jr., How the Snake Lost Its Legs: Curious Tales from the Frontier of Evo-­Devo (Cambridge: Cambridge University Press, 2014), 71. 24. Other important pieces on Bateson are the biography by Alan Cock and Donald R. Forsdyke, Treasure Your Exceptions: The Science and Life of William Bateson (New York: Springer Publishing, 2008); and Stuart A. Newman, “William Bateson’s Physicalist Ideas,” in From Embryology to Evo-­Devo: A History of Developmental Evolution, ed. Manfred D. Laubichler and Jane Maienschein (Cambridge, Mass.: MIT Press, 2007), 83–­108. Also, Bateson figures prominently in

240 · NOTES TO CHAPTER 4

Berris Charnley’s dissertation, Agricultural Science, Plant Breeding and the Emergence of a Mendelian System in Britain, 1880–­1930 (Leeds, UK: The University of Leeds, 2011). 25. William Bateson, Materials for the Study of Variation Treated with Especial Regard to Discontinuity in the Origin of Species (London: Macmillan and Co., 1894), vi. 26. W. Bateson, 16. 27. W. Bateson, 23. 28. W. Bateson, 20. Emphasis in the original. 29. W. Bateson, 35. 30. The list here is long, but two important references are Aristotle, The History of Animals Book IV, and Georges Cuvier, Leonhard Stejneger, and M. Valen­ ciennes, Le Règne Animal Distribué d’Après son Organisation, pour Servir de Base a l’Histoire Naturelle des Animaux et d’Introduction a l’Anatomie Comparée (Paris: Chez Déterville, 1817). Also important is the work of Étienne Geoffroy Saint-­ Hilaire, and as we have seen, Goethe’s studies in morphology. 31. William Bateson, Materials for the Study of Variation, 19. 32. W. Bateson, 17. 33. W. Bateson, 14. 34. Gregory Bateson, Steps to an Ecology of Mind: Collected Essays in Anthropology, Psychiatry, Evolution, and Epistemology (Chicago: University of Chicago Press, 2000), 380. 35. G. Bateson, 380. 36. This was part of what I wanted to address in my book, The Emergence of Genetic Rationality, when I tried to anchor the record-­keeping practices in turn-ofthe-twentieth-­century American biology in specific experiences of space and time. 37. The list for this is long: for a historical account that uses this framework see Lily E. Kay, Who Wrote the Book of Life?: A History of the Genetic Code (Stanford, Calif.: Stanford University Press, 2000). For a recent theoretical account see Alexander R. Galloway, The Interface Effect (Cambridge: Polity, 2013). 38. W. Bateson, Materials for the Study of Variation, 379. 39. William Bateson, Problems of Genetics: With Illustrations (New Haven, Conn.: Yale University Press, 1913). See the discussion on pages 6–­79. Also, see the discussion of harmonics in relationship to layers in chapter 5 of this book. 40. Stephan S. Hall, Mapping the Next Millennium (Random House: New York, 1992), 194. 41. Ian Duncan and Geoffrey Montgomery, “E. B. Lewis and the Bithorax Complex: Part II. From Cis-­Trans Test to the Genetic Control of Development,” Genetics 161, no. 1 (2002): 4. 42. E. B. Lewis, “A Gene Complex Controlling Segmentation in Drosophila,” Nature 276 (1978): 565–­70.

NOTES TO CHAPTER 4 · 241

43. Howard D. Lipshitz, Genes, Development, and Cancer: The Life and Work of Edward B. Lewis (Boston: Kluwer Academic, 2004), 167–­68. 44. Terrell Tannen, “Edward B. Lewis,” The Lancet 364, no. 9435 (2004): 658. 45. See, for instance, Paul Berg’s 1980 Nobel lecture published as “Dissections and Reconstructions of Genes and Chromosomes,” Science 213, no. 4505 (1981): 296–­303. 46. Hall, Mapping the Next Millennium, 194. 47. Calvin B. Bridges and Thomas Hunt Morgan, The Third-­Chromosome Group of Mutant Characters of Drosophila Melanogaster (Washington, D.C.: Carne­gie Institution of Washington, 1923). 48. The precise genetics of the transformations were worked out by Ed Lewis in a paper published in 1949. E. B. Lewis, “A Study of Adjacent Genes,” Heredity 3 (1949): 130. 49. Ian Duncan and Geoffrey Montgomery, “E. B. Lewis and the Bithorax Complex,” Genetics 160, no. 4 (2002): 1269. 50. See for instance, G. W. Beadle and E. L. Tatum, “Genetic Control of Biochemical Reactions in Neurospora,” Proceedings of the National Academy of the Sciences, U.S. 27 (1941): 499–­506. 51. Barbara McClintock, “The Relation of Homozygous Deficiencies to Mutations and Allelic Series in Maize,” Genetics 29, no. 5 (1944): 478–­502. 52. E. B. Lewis, “Pseudoallelism and Gene Evolution,” Cold Spring Harbor Symposia on Quantitative Biology 16 (1951): 159–­74, doi:10.1101/SQB.1951.016.01.014. 53. Ian Duncan and Geoffrey Montgomery, “E. B. Lewis and the Bithorax Complex,” Genetics 160, no. 4 (2002): 1267. 54. Lewis, “Pseudoallelism and Gene Evolution, 159. 55. Peter A. Lawrence and Michael Locke, “A Man for Our Season,” Nature 386, no. 6627 (1997): 757–­58. 56. James F. Crow and Welcome Bender, “Edward B. Lewis, 1918–­2004,” Genetics 168, no. 4 (2004): 1778. 57. Vilém Flusser, Towards a Philosophy of Photography, trans. Anthony Matthews (London: Reaktion Books, 2000), 9. 58. For key reviews see Gerd B. Müller and Günter P. Wagner, “Homology, Hox Genes, and Developmental Integration,” American Zoologist 36, no. 1 (1996): 4–­13; and Gunte P. Wagner, Chris Amemiya, and Frank Ruddle, “Hox Cluster Duplications and the Opportunity for Evolutionary Novelties,” Proceedings of the National Academy of Sciences of the United States of America 100, no. 25 (2003): 14603–­6. For the important article that identified key sequences in the homeotic genes and thus allowed for studies comparing sequences across species see William McGinnis et al., “A Conserved DNA Sequence in Homoeotic Genes of the Drosophila Antenna­ pedia and Bithorax Complexes,” Nature 308, no. 5958 (1984): 428–­33. 59. Although this is the visual effect, it is important to note that Lewis did not

242 · NOTES TO CHAPTER 4

think of the segments as modules (or, more precisely, as genetically discrete elements). Instead, each of the genes in the complex defined the modules. This is important for bithorax as the original mutant bx only affected half the development consequence for the segment as only half of the haltere was replaced with a wing. 60. Howard D. Lipshitz, “From Fruit Flies to Fallout: Ed Lewis and His Science,” Developmental Dynamics 232, no. 3 (2005): 529–­46. 61. See note 58. 62. Lewis, “A Gene Complex Controlling Segmentation in Drosophila,” 570. 63. Lipshitz, “From Fruit Flies to Fallout,” see especially 534–­35. 64. Vilém Flusser, Into the Universe of Technical Images, trans. Nancy Ann Roth (Minneapolis: University of Minnesota Press, 2011), 21. 65. The history of genetics would not be possible without the study of mutants. The list of papers supporting this claim is long and includes almost all the literature on the history of genetics. A recent and comprehensive entry point into this literature would be Staffan Müller-­Wille and Hans-­Jörg Rheinberger, A Cultural History of Heredity (Chicago: University of Chicago Press, 2014). 66. See for instance the email communication from Ed Lewis to H. Allen Orr from 30 December 1996. Lewis writes: “I was impressed by Bateson’s incredible insight [on homeosis]” and that he thinks the low regard for Bateson by many geneticists was due to his initial reluctance to adopt the chromosome theory of inheritance. Edward B. Lewis Papers, California Institute of Technology, Box 5. 67. W. Bateson, Materials for the Study of Variation, 76. 68. Lewis, “A Gene Complex Controlling Segmentation in Drosophila,” 569. 69. An interesting analog to the comparison between Bateson and morphology on form in the social sciences is the comparison between the emphasis on imitation and repetition of Gabriel Tarde and the organicism of Emile Durkheim. See Bruno Latour, “Gabriel Tarde and the End of the Social,” in The Social in Question. New Bearings in History and the Social Sciences, ed. Patrick Joyce (London: Routledge, 2014), 117–­32. Also important are the articles found in the “Gabriele Tarde” special edition of Economy and Society, ed. Andrew Barry and Nigel Thrift, 36, no. 4 (2007). 70. Christiane Nüsslein-­Volhard and Eric Wieschaus, “Mutations Affecting Segment Number and Polarity in Drosophila,” Nature 287, no. 5785 (1980): 795–­801. 71. See for instance, the page on Ed Lewis on the Nobel Prize’s website, ­Edward B. Lewis—­ Facts, http://www.nobelprize.org/nobel_prizes/medicine/ laureates/1995/lewis-facts.html. 72. McGinnis, et al., “A Conserved DNA Sequence in Homoeotic Genes of the Drosophila Antennapedia and Bithorax Complexes,” 428–­33. 73. Carroll, Endless Forms Most Beautiful, 100.

NOTES TO CHAPTER 5 · 243

5. Drawing Together 1. “Edward B. Lewis—­ Facts,” https://www.nobelprize.org/nobel_prizes/ medicine/laureates/1995/lewis-facts.html. 2. On Ed Lewis and animation, see Matthew P. Scott and Peter A. Lawrence, “Obituary: Edward B. Lewis (1918–­2004),” Nature 431, no. 7005 (2004): 143; Terrell Tannen, “Edward B. Lewis,” Lancet 364, no. 9435 (2004): 658; and Ian Duncan, and Geoffrey Montgomery. “E. B. Lewis and the Bithorax Complex: Part II. From Cis-­ Trans Test to the Genetic Control of Development,” Genetics 161, no. 1 (2002): 1–­10. 3. Computer Graphics World 27, no. 12 (December 2004): 12–­14. Also, famously, Lev Manovich had a chapter on compositing in his influential The Language of New Media (Cambridge, Mass.: MIT Press, 2002). Tellingly, Manovich compares the technique to film editing. As I will argue below, despite some of their similarities, compositing originally has much more to do with the graphic arts than cinema. The fact that it is used to create animation is a very interesting exploration into how layering an image, in addition to grids, can create dynamic constructions needed for world building. 4. François Jacob, “Time and the Invention of the Future,” in The Possible and the Actual (Seattle: University of Washington Press, 1994), 53. 5. Eric Hazen, “Morphing Confocal Images and Digital Movie Production Methods in Molecular Biology,” Confocal Microscopy Methods and Protocols, Vol. 122 (Totowa, N. J.: Humana Press, 1998), 421. 6. There are several recent studies that point to the link between animation, vitality, and life. David Willis, Inanimation: Theories of Inorganic Life (Minneapolis: University of Minnesota Press, 2016); Inga Pollman, Cinematic Vitalism: Film Theory and the Question of Life (Amsterdam: Amsterdam University Press, 2017); Mel Chen, Animacies: Biopolitics, Racial Mattering, and Queer Affect (Durham, N.C.: Duke University Press, 2012); and Phillip Thurtle, “Animation and Vitality,” in “Animating Biophilosphy,” Inflexions 7 (March 2014): 98–­117, www.inflexions .org. Especially important has been Robert Mitchell, Bioart and the Vitality of Media (Seattle: University of Washington Press, 2010). Jane Bennet’s Vibrant Matter: A Political Ecology of Things (Durham, N.C.: Duke University Press, 2010) has also been especially influential. 7. See, for instance, the book by Ollie Johnston and Frank Thomas, The Illusion of Life: Disney Animation (New York: Disney Editions, 1981). 8. Thomas Lamarre, The Anime Machine: A Media Theory of Animation (Minneapolis: University of Minnesota Press, 2009), xxv. 9. Tom Sito, Moving Innovation: A History of Computer Animation (Cambridge, Mass.: MIT Press, 2013), 5–­7. 10. See Lisa Cartwright, Screening the Body: Tracing Medicine’s Visual Culture

244 · NOTES TO CHAPTER 5

(Minneapolis: University of Minnesota Press, 1995); and José Van Dijck, The Transparent Body: A Cultural Analysis of Medical Imaging (Seattle: University of Washington Press, 2005). 11. Janet H. Iwasa, “Bringing Macromolecular Machinery to Life Using 3D Animation,” Current Opinion in Structural Biology 31 (2015): 85. 12. Lamarre, The Anime Machine, xxiv. 13. Such as that found in Laura Mulvey’s important essay “Visual Pleasure and Narrative Cinema,” in Visual and Other Pleasures (New York: Palgrave Macmillan, 2009). 14. I’m thinking here of Sean Carroll’s analysis of the molecule distal-­less in the expression of the wing spots in butterfly wings. Sean B. Carroll et al., “Pattern Formation and Eyespot Determination in Butterfly Wings,” Science 265, no. 5168 (1994): 109–­14. 15. This is the goal of a collaborative project undertaken with Adam Nocek and Tyler Fox provisionally entitled, “super-­naturalisms.” 16. Please see Science Is Fiction: The Films of Jean Painlevé, ed. Andy Masaki Bellows and Marina McDougall (Cambridge, Mass.: MIT Press, 2001). For an interesting discussion of the history of Painlevé’s documentaries in relationship to images in the history of science, see Adam J. Nocek, Animate Biology: Data, Visualization, and Life’s Moving Image, dissertation, University of Washington, 2015. 17. Scott and Lawrence, “Obituary: Edward B. Lewis (1918–­2004),” 143 18. André Bazin, “The Ontology of the Photographic Image,” trans. Hugh Gray, Film Quarterly 13, no. 4 (1960): 7–­8. 19. Bernard Stiegler, Technics and Time, 2: Disorientation, trans. Stephen Barker (Stanford, Calif.: Stanford University Press, 2008), 16. 20. Vilém Flusser, Into the Universe of Technical Images, trans. Nancy Ann Roth (Minneapolis: University of Minnesota Press, 2011), 48. 21. Bazin, “The Ontology of the Photographic Image,” 8. 22. Bazin, 9. 23. It is useful to note that one other important area of research for Ed Lewis was researching how best to describe the effect of radioactive fallout for Japanese survivors of the nuclear bomb. See Howard D. Lipshitz, Genes, Development, and Cancer: The Life and Work of Edward B. Lewis (Boston: Kluwer Academic, 2004), for an overview. 24. There is a vibrant history of heredity in relationship to industrialism and empire. See Berris Charnley, “Geneticists on the Farm: Agriculture and the All-­ English Loaf,” Scientific Governance in Britain, 1914–­1979, ed. Charlotte Sleigh and Don Leggett (Manchester, UK: Manchester University Press, 2016), 181–­98; Berris Charnley, “Experiments in Empire-­Building: Mendelian Genetics as a National, Imperial, and Global Agricultural Enterprise,” in Studies in the History and Phi-

NOTES TO CHAPTER 5 · 245

losophy of Science: Part A 44, no.2 (2013): 292–­300; Dominic Berry, “The Plant Breeding Industry after Pure Line Theory: Lessons from the National Institute of Agricultural Botany,” Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences 46 (2014): 25–­37; Alain Pottage and Brad Sherman, “Organisms and Manufactures: On the History of Plant Inventions,” Melbourne University Law Review 31, no. 2 (2007): 539–­68; Christophe Bonneuil, “Mendelism, Plant Breeding, and Experimental Cultures: Agriculture and the Development of Genetics in France,” Journal of the History of Biology 39, no. 2 (2006): 281–­308; Christophe Bonneuil, “Producing Identity, Industrializing Purity: Elements for a Cultural History of Genetics” in A Cultural History of Heredity IV: Heredity in the Century of the Gene, Preprint 343, ed. Staffan Müller-­Wille, Hans-­Jörg Rheinberger, and John Dupré (Berlin: Max Planck Institute for the History of Science, 2008); Helen Anne Curry, “Industrial Evolution: Mechanical and Biological Innovation at the General Electric Research Laboratory,” Technology and Culture 54, no. 4 (2013): 746–­81; and Phillip Thurtle, The Emergence of Genetic Rationality: Space, Time & Information in American Biological Science, 1870–­1920 (Seattle: University of Washington Press, 2008). 25. Vilém Flusser, Towards a Philosophy of Photography, trans. Anthony Matthews (London: Reaktion Books, 2000), 66–­67. 26. Flusser, 36. 27. The most important reference regarding the relationship of film to perceptions of time and space are undoubtedly the cinema studies of Gilles Deleuze. Gilles Deleuze, Cinema 1: The Movement-­Image, trans. Hugh Tomlinson and Barbara Habberjam (Minneapolis: University of Minnesota Press, 1986); and Gilles Deleuze, Cinema 2: The Time-­Image, trans. Hugh Tomlinson and Robert Galeta (Minneapolis: University of Minnesota Press, 1989). 28. Walter Murch, In the Blink of an Eye: A Perspective of Film Editing, 2/e (Los Angeles: Silman-­James Press, 2001), 3–­4. I’ve argued elsewhere that Deleuze’s focus on screen, shot, and cut is a bit misleading for thinking about the vitality of animation. See Thurtle, “Animation and Vitality.” 29. Verina Gfader, “Nervous Light Planes,”  Animation  3, no. 2 (July 2008): 147–­67. 30. See Kit Laybourne, The Animation Book: A Complete Guide to Animated Filmmaking—­From Flip-­Books to Sound Cartoons to 3-­D Animation (New York: Three Rivers Press, 1998). 31. Tina Sotiriadi, William Kentridge, Third Text 13, no. 48 (1999): 106–­8. 32. See the interview of William Kentridge discussing his film Tide  Table in “Beach Life: William Kentridge and Tide Table” (San Francisco Museum of Modern Art), https://www.sfmoma.org/watch/beach-life-william-kentridge-and-tide -table/.

246 · NOTES TO CHAPTER 5

33. See the discussion of Charles Darwin’s writing on the ocean in Gillian Beer, Darwin’s Plots: Evolutionary Narrative in Darwin, George Eliot and Nineteenth-­ Century Fiction (Cambridge, Mass.: Cambridge University Press, 1983), 114–­36. 34. Beatrice Bateson, William Bateson, Naturalist: His Essays and Addresses Together with a Short Account of His Life (Cambridge, Mass.: Cambridge University Press, 1928), 46. 35. B. Bateson, 35. 36. B. Bateson, 45. 37. B. Bateson, 36. 38. B. Bateson, 38. 39. E. G. Lutz, Animated Cartoons: How They Are Made, Their Origin and Development (New York: Charles Scribner and Sons, 1920). 40. A good general book on wave dynamics, albeit with math is, Fredric Raichlen, Waves (Cambridge, Mass.: MIT Press, 2012). An excellent popular exploration in the dynamics of waves in many media is Gavin Pretor-­Pinney, The Wavewatcher’s Companion (London: Bloomsbury, 2010). 41. Serving as inspiration here is Aden Evans’s analysis of Fourier Transform analysis in Sound Ideas: Music, Machine, and Experience (Minneapolis: University of Minnesota Press, 2005), 2–­8. 42. The importance of compositing is one of the most important lessons from Thomas Lamarre, The Anime Machine (Minneapolis: University of Minnesota Press, 2009). 43. This is the implicit argument for the structure of most histories of film, where a section on drawn moving images precedes photographed moving images. See for instance, Charles Musser, The Emergence of Cinema: The American Screen to 1907 (Berkeley: University of California Press, 1994). 44. Most of the known examples of these films have been in structural biology, where scientists are eager to see how molecular forms and functions interact in the interior of a cell. I will cover two especially important animators, Drew Berry and Janet Iwasa. It is important to remember that structural biology’s use of animation is most definitely not the only use biologists found for animation. I cover Ed Lewis’s animated movie in detail, but labs in evo devo also use animation to great effect. Sean Carroll’s lab is one that has used computerized animation technologies to make points about evolution and development, http://carroll.molbio .wisc.edu/movies.html. 45. Ria Misra, “Animator Drew Berry Explains How to Get a Job in Science Art,” io9, August 14, 2016, www.io9.com. 46. Stephen Curry, Drew Berry quoted in “The Art and Science of Animating Life,” The Guardian, June 9, 2015. 47. Curry. 48. “On Beginning and Ending with Apoptosis: Cell Death and Biomedicine,”

NOTES TO EPILOGUE · 247

in Remaking Life and Death: Toward an Anthropology of the Life Sciences, ed. Sarah Franklin and Margaret Lock (Santa Fe, N.M.: School of American Research Press, 2003). 49. Stephen Curry, Drew Berry quoted in “The Art and Science of Animating Life.” 50. Lamarre, The Anime Machine, 30. 51. Janet H. Iwasa, “Animating the Model Figure,” Trends in Cell Biology 20, no. 12 (2010): 699–­704. 52. Iwasa, 699. 53. B. Bateson, William Bateson, Naturalist, 45. 54. Janet Iwasa, “Crafting a Career in Molecular Animation,” Molecular Biology of the Cell 25, no. 19 (2014): 2891–­93. 55. Please see Adam Nocek’s upcoming work on visualization in the biological sciences for a theoretically robust discussion on how animations can be conceived as a form of experimentation. His manuscript is provisionally named, Animating Capital: Molecules, Labor and the Cultural Production of Science. 56. Phillip Thurtle and A. J. Nocek, “Vitalizing Thought,” in “Animating Bio­ philosphy,” Inflexions 7 (March 2014): i–­xi. 57. For an important discussion on the chiasm and the flesh of the world see Maurice Merleau-­Ponty, “The Intertwining—­The Chiasm,” The Visible and the Invisible, trans. Alphonso Lingis (Evanston, Ill.: Northwestern University Press, 1968). For more contemporary uses of the concept of chiasm see Michael Fortune, Promising Genomics: Iceland and deCODE Genetics in a World of Speculation (Berkeley: University of California Press, 2008); and Elizabeth Wilson, Psychosomatic: Feminism and the Neurological Body (Durham, N.C.: Duke University Press, 2004), 58–­59.

Epilogue 1. Andrew Russell and Lee Vinsel, “Hail the Maintainers,”, Aeon, April 7, 2016, accessed 20 August 2017. https://aeon.co/essays/innovation-is-overvalued -maintenance-often-matters-more. 2. María Puig de la Bellacasa, Matters of Care: Speculative Ethics in More Than Human Worlds (Minneapolis: University of Minnesota Press, 2017). 3. William Bateson, Materials for the Study of Variation Treated with Especial Regard to Discontinuity in the Origin of Species (London: Macmillan and Co., 1894), 76. Emphasis mine.

This page intentionally left blank

Index Adam (Rodin), 26 (fig.) Adams, Neal, 73 advertising, 3, 4, 5, 6, 15, 55, 65, 69, 70, 72; comics and, 73, 74, 78–79, 90; development of, 102; mass market, 102; revenue from, 80 aesthetics, 18, 29, 56, 124, 125; biology and, 30, 31; fluid, 134; ideas on, 30; life and, 134; politics and, 54, 124; types of, 54 affects, regulating, 106–9 After Effects, 194 Agamben, Giorgio, 94, 116, 117, 118, 235n61 Agenda for Antiquity (Rainger), 32 Alh, Valerie, 120 Alias Systems: Maya and, 170 Allen, Garland, 221n18 Allen, T. F. H., 120 American Association for the Advancement of Science, 87 American Museum of Natural History, 32 analysis, 10, 14, 16, 18, 36, 66, 126; aesthetic, 15, 18–19; political, 17; proliferation of, 28 anatomy, 129, 176; biology and, 208 animal forms: creation of, 137; diversity in, 166 Animated Cartoons (Lutz), 189 “Animating the Model Figure” (Iwasa), 203

animation, 2, 3, 167, 168, 174, 177, 183, 196–98, 199–200, 203, 206; background image for, 189 (fig.); biology and, 120–21, 207, 208, 246n44; biomedical, 196; cel, 171, 184–85, 188–89, 190, 191 (fig.), 194, 201; compositing and, 192, 208; computerized, 171; control by, 176, 181; epistemic value of, 204; filmed, 171, 184; images and, 172, 184, 208; movement in, 192; ontology of, 204; photorealism and, 177; politics and, 173, 208; programs, 171, 194, 201, 205; scientific research and, 203–4; stop-motion, 18, 20, 184, 204; technical, 204–5; 3D, 171; twentiethcentury, 171, 182, 188; understanding, 182, 194–95; using, 169–70, 172 animators, work of, 190, 195–96 Anime Machine, The (Lamarre), 170 Anthropogenie (Haeckel), 47, 50, 169; plate from, 48 (fig.), 51 (fig.) Apoptosis (Berry), 196, 199 (fig.); scene capture from, 197 (fig.) Aristotle, 225n7, 240n30 art, 2, 30, 225n11; biology and, 31, 213; fine, 188; industrial, 188; science and, 28–29, 31 associations, 69, 238n6; creating, 124; dissolution of, 39 associative logic, 71, 74, 82, 83, 85 associative meanings, 69, 70

249

250 · INDEX

associative power, 14, 15, 68, 74, 102, 168 associative thinking, 14, 132 backgrounds, 168, 170, 189, 190, 193, 195; cross-disciplinary, 20 Baldwin, Carliss, 133 Bateson, Beatrice, 186 Bateson, Gregory (father), 16 Bateson, Gregory (son), 143 Bateson, William, 18, 129, 130, 179, 242n69; evo devo and, 137; homeosis and, 113–14, 235n51, 242n66; illustrations by, 17, 141 (fig.), 142 (fig.); Lewis and, 145, 164; literature on, 239n24; modularity and, 138; on normal forms, 213; organisms and, 138–39; patterns and, 143, 204; segments and, 187; selection and, 138; variations and, 17, 137–45; waves and, 186; work of, 16, 144 Baudrillard, Jean, 14, 222n31 Bazin, André, 177, 178, 179, 181 Beadle, G. W., 151 beauty, 30, 33, 107; reflective judgment of, 42; sublime and, 227n38 Beck, C. C., 73 Beethoven, Ludwig van, 165 Beniger, James, 70 Benton, Mike, 73 Berg, Paul, 146, 241n45 Berkeley Drosophilia Transcription Network (BDTN), 13 Berry, Drew, 195–96, 246n44; animation and, 196–98, 199–200; Apoptosis of, 199 (fig.); scene capture of, 197 (fig.); visualization software and, 205 biochemistry, 5, 22, 131, 150, 153 bioengineering, 2, 8, 213 bioinformatics, 2, 21

biological processes, 20, 21, 22, 25, 101, 107–8 biological sciences, 3, 56, 68, 107, 172, 208, 229n7, 247n55 biological thought, 64, 169, 208; change in, 171; understanding, 115–19; using, 119–22 biology: cell, 8; consumer, 17; cultural basis of, 3; discipline in, 129; history of, 29, 32, 118, 129, 133, 146, 169; molecular, 1, 2; nature of, 31, 118; principles of, 19; techniques/­ insights for, 208; types of, 164 biomedia, goal of, 21 biomolecules, 1, 2, 205 biopolitics, 115, 121–22, 125, 235nn55–56; conception of, 116; politics end of, 122 biopower, 94, 100, 115, 116, 232n7, 235n62; homeostasis and, 118; ­oppositional forms of, 117; populations and, 118; prevalence of, 94; study of, 118; theories of, 118, 119; understanding, 116 Birth of Biopolitics and Security, Territory, and Population (Foucault), 16, 119 bithorax complex, 114, 137, 148, 149, 150 (fig.), 155, 156, 158, 162, 164, 165, 200, 201–2; development of, 152; mutant of, 149 (fig.), 152 “Bithorax Complex, The” (Lewis), 167 bithoraxoid, 148 Blake, William, 25; title page of, 27 (fig.) Blender, 205 Blumberg, Mark S., 138 bodies: behavior of, 21; industrialization of, 99, 233n15; misregulation of, 104; modularity and, 135–36; procreative power of, 94; regulation

INDEX · 251

of, 16, 115–19; theories of, 121; ­understanding, 109–15 Bolker, Jessica, 129 Brain, Robert, 44 brands, 70; consumers and, 4 Breidbach, Olaf, 46 Bridges, Calvin, 148; illustration by, 150 (fig.); mutations and, 149, 151–52 Broad Institute of the Massachusetts Institute of Technology, 81 Brundle, Seth, 111, 127 Burian, Richard, 19 Buscema, John, 25; illustration from, 28 (fig.) butterflies, 141 (fig.) Caltech genetics group, 145, 167 Canguilhem, Georges, 118, 119, 237n69 Caniff, Milton, 73 Cannon, Walter, 22 Captain Marvel, 73 carbon, 29, 38; associative properties of, 39; binding of, 39; philos and, 39–40 Carroll, Sean, 17, 138, 156, 166, 244n14, 246n44; development and, 136–37; modularity and, 144 Carus, Carl Gustav, 33 causalities, 10, 151, 159 Cell (journal), 13 cells, 1, 2, 188, 189, 212; dynamics of, 208–9 cellular biology, 208 cellular dynamics, 18, 172, 198 cellular membranes, 94, 120, 121 change: biological, 169, 173, 206; evolutionary, 195; metaphysical, 184; molecular, 195; organismal, 195; political economic, 184; regulating, 184, 188–90, 192–95; structure and, 29–30; waves of, 181–88

Chaplin, Charlie, 99 chemistry, 117; biology and, 208 Chimera, 171 Chiroptera, 44, 45, 45 (fig.) chitin, 7, 107, 108, 132 chromosomes, 150, 151, 158, 165 circulation, desire and, 96–98 Clark, Kim, 133 clothes: buying, 221n17; virtual trying on, 9 codes, 15; computer, 21; genetic, 8, 21–22; organisms and, 21 Cohen, Lizbeth, 95, 232n9 Cole, Jack, 74 comics, 88–89, 91–92, 93, 132; advertising and, 73, 74, 78, 90; content management and, 88; grids and, 72–74, 78–79 communications, 144; cybernetics, 234n44; electronic, 91; industrialized, 55; scientific, 56; visual, 88–89 community, 115–16; formation of, 115, 119; organizations, 116 complexity, 1, 5, 7, 33, 36, 37, 39, 57, 66, 69, 79, 95, 98, 110, 113; architectural, 44, 45; biological, 2, 6, 117; material, 94 composite motion, 195–96, 198–200 composite sensibilities, value of, 22–23 compositing, 18, 168–69, 172, 208, 209; animation and, 192, 208; different types of, 172, 212 composition, 28, 32, 36, 42, 49, 50, 78, 103, 139, 194, 195, 200; dynamics, 194; elements of, 184, 190; internal, 78 Computer Graphics World (Sauer), 168 computers, 89; biology and, 2, 21; databases and, 19; enthusiastic use of, 168 computing, 2, 19–22; biology and, 144, 171; rise of, 3, 144

252 · INDEX

concepts, 16, 20, 21, 68, 118; biological, 3, 29, 236n67 Connor, Steven, 104 constructivism, 16, 94, 183 consumerism, 4, 70, 71–72, 91, 93, 96 consumers, 79, 98; envisioning, 3–4; regulating, 95–96 consumption, 7, 96, 125, 126, 220n7 control, 144, 181; politics of, 23 Cootie game, 134–35, 135 (fig.), 148 critical theory, 123, 124, 169–70 Critique of Judgment, The (Kant), 30, 40, 41, 225n8, 227n29; reading, 41, 227n29 Critique of Practical Reason (Kant), 41 Critique of Pure Reason (Kant), 41 Cronenberg, David, 111, 112, 113, 127 culture: aesthetics of grids in, 55; consumer, 232n9; mass, 95, 96; popular, 144; scientific, 83; visual, 32, 56 Curry, Stephen, 196 Curse of the Fly, The (film), 127 cybernetics, 21, 143, 144, 219n5, 234n44 cytology, 151 Darwin, Charles, 44, 46, 51, 245n33; political economy and, 236n63 data: collections of, 16; crystallo­ graphic, 20; evaluating, 19; points, 11, 13, 204; scientific, 207; standardized, 21; strands of, 199 Davidson, Eric H., 7–8 Davis, Geena, 127 Davis, Jack, 73 death: cell, 196; concept of, 236n67; social order and, 237n67 death ligand, 198 death receptors, 198 Delambre, André, 96, 97, 101, 109–10 Delambre, François, 96

Delambre, Helene, 91, 92; poster of, 92 (fig.) Deleuze, Gilles, 93, 224n55, 225n11, 227n29, 245nn27–28; assem­ blages and, 123; interview of, 237n75 Design Rules (Baldwin and Clark), 133 desire, 3, 4, 6, 7, 10, 16, 39, 72, 100, 102, 124, 125, 133, 143; circulation and, 96–98 Desmond, Adrian, 236n63 development, 8, 9, 31–32, 90, 129, 135, 160, 169, 195; animal, 154; biological theorizing, 132; complexities of, 56; comprehensive view of, 161; concept of, 17, 90, 132; embryonic, 57, 58 (fig.), 165; envisioning, 164; grids and, 56–57, 59–62, 64–66, 238n3; levels of, 152, 155, 158–59; model of, 171; modular vision of, 165; molecular, 203; narrative of, 60, 65–66; normal, 59, 60; regulating, 109–15; sequential, 48–49; studies of, 136–37; technical, 189; understanding, 106–9 developmental biology, 2, 8, 18, 57, 60, 61, 62, 95, 129, 136, 138, 148; development of, 154; discipline of, 17; grids and, 64; image from, 11; morphology and, 53 developmental series, 46–47, 49 Diptera, 104 discipline: constructivist moment of, 98–100; depiction of, 98 Discipline and Punish (Foucault), 98–100 Discomedusa, 35, 36 Disney, 182, 184 Disney, Walt, 188, 189 distribution, 4, 70, 72, 90, 93, 96, 97, 98, 101, 106, 125, 170, 188; mass, 55

INDEX · 253

diversity: chemical/morphological, 211–12; ontological, 212 DNA, 111, 112, 165, 201 Drosophila, 108, 166, 167, 173; adult, 147 (fig.); bithorax, 136; embryo, 11, 13, 60, 86; larva, 13, 147 (fig.), 201; life cycle of, 174, 175 (fig.) Drosophila melanogaster, 104; antenna­ pedia mutant, 114 (fig.) Durkheim, Emile, 242n69 Each One a Hero (Kurokawa), 234n29 “Early Effects of the Bithorax Gene Complex” (Lewis), 176 Earnest, Olen, 97 ecology, 68, 134, 169 embryology, 32, 47, 48, 49, 55 embryos, 12, 47, 49, 62, 161; development of, 57, 165; fruit fly, 12 (fig.); images of, 146 Emergence of Genetic Rationality, The (Thurtle), 3, 50 Encyclopedia Britannica Film, 105 Endless Forms Most Beautiful (Carroll), 136, 138 entertainment industry, 4, 6, 79, 87, 90, 168, 171, 194 envisioners, 11, 13, 163, 209, 211, 212 envisioning, 14–18, 126, 162–63, 171, 172, 204, 211; concept of, 13; layers for, 18; role of, 3; visualization and, 9–14 epistemology, 18, 66, 85, 200, 208 Esposito, Roberto, 94, 117, 119; biopolitics and, 115, 116, 235n56; biopower and, 116 eugenics, 50 Evans, Aden, 246n41 evolutionary biology, 2, 8, 18, 36, 40, 50, 60, 61, 95, 129, 132, 138, 208; development of, 154; discipline of,

17; grids and, 62, 64; image from, 11; morphology and, 53 “Evolution of the Fly” (Lewis), 174; stills from, 175 (fig.) exclusions, 126, 127; politics of, 121, 123 experimentation, 20, 23, 122 expressions, development of, 157 (fig.) Fables of Abundance (Lears), 78 Factor, Max, 110 Fangerau, Heiner, 228n53 Fantastic Four #94, The: advertisement page from, 75 (fig.) fantasy, 87, 110, 111, 181; appeal to, 178; truth and, 86 filmmaking, 168, 176, 183, 184 Fine, Lou, 73 flies, 105–6, 110, 115; evolution of, 175 (fig.); grids/screens and, 93; humans and, 107–8; mutant, 146, 151; physiological needs of, 106; physiological structures of, 160; realism of, 155 Flipbook, 205, 207 (fig.) fluidity, 99, 103, 118, 188; forms of, 204 Flusser, Vilém, 66–68, 126, 154, 171, 184, 221n19; analysis by, 16, 22; apparatus and, 14; camera and, 183; codes and, 15; envisioning and, 13, 14, 162–63, 178, 204, 209, 212; historical thought and, 10–11; hyper-real and, 14; images and, 14, 68, 72, 83–84, 182; judgment/ reason and, 230n18; medium and, 13–14; molecular movement and, 198; phenomenological qualities and, 88; political economic and, 10; technical images and, 18, 87, 88–89, 132; technology and, 15; visual magic and, 15

254 · INDEX

Fly, The (film), 16, 92, 109, 125, 127, 180, 232n8; misregulation and, 96, 106; opening credits to, 105 (fig.); pests and, 104; publicity poster for, 91, 92 (fig.), 95, 97; regulation and, 107; scene from, 111 (fig.), 112 (fig.), 126 (fig.) Fly II, The (film), 127 fly/human hybrid, 112 (fig.), 125, 126; human-sized, 110; imagining, 111 Focillon, Henri, 30, 31, 45, 46 form, 28, 29, 225n7; biology and, 19, 31; function and, 18–19; normal, 155, 213; organic, 65, 100; racial politics of, 49–54; relationships of, 84; science and, 31; sequence and, 32 formalism: organic, 53, 134; scientific/ political implications of, 54; true, 53, 134 Foucault, Michel, 6, 16, 17, 99, 100, 206, 229n7, 233n26; analysis by, 102; biopower and, 94, 118; biopower/ biopolitics and, 115; neoliberalism and, 119; partitioning and, 98; production and, 102; regulation and, 102, 106; usefulness and, 182 Fourier Transform analysis, 20, 193, 193 (fig.), 246n41 Fox, Tyler: super-naturalisms and, 244n15 Freaks of Nature (Blumberg), 138 From DNA to Diversity (Bateson), 138 fruit flies, 104, 113, 145, 154, 174, 181; adult, 155, 156; genetics of, 114 Funnies on Parade, 73 Gaines, Max C., 73 Galison, Peter: mechanical images and, 231n46 Galloway, Alexander: protocol and, 14 “Gene Complex Controlling Seg­

mentation in Drosophila, A” (Lewis), 145 gene expression, 13, 153–54, 155, 165 gene products, 12, 160; autonomy/ interdependence of, 154; expression of, 153 genes, 60, 158, 163; behavior of, 130; BX-C, 161, 164; cluster of, 150; code of, 8; eukaryotic, 146; eve, 12–13, 12 (fig.); gap, 12–13; homeotic, 241n58; kruppel, 12–13, 12 (fig.); master control, 160; mixing, 113; mutations and, 152; proteins and, 153; regulation of, 61; regulatory, 8, 22, 166; structural, 22 “Genetic Control of Segmental Levels of Development in Thoracic and Abdominal Segments,” 156 (fig.) genetics, 8, 17, 22, 65, 112, 137, 138, 143, 154, 157, 163; add-back, 162; bacterial, 146; biology and, 208; criticisms of, 130; history of, 155, 242n65; research in, 104, 178; transmission, 130 genomes, 7, 112, 113 Gerson, Elihu: social pragmatics and, 223n46 Glass, Philip, 165 Glitsch, Adolph, 33 Goethe, Johann Wolfgang von, 44, 165, 225n8, 226n20, 227n28; morphology of, 30, 33, 240n30; reasoning of, 35 Goldblum, Jeff, 127 Goldschmidt, Richard, 235n53 Gomez-Palacio, Bryony: on grids, 69 goods: display of, 70–72; regulation of, 9 graphic design, 2, 4, 5, 15, 72, 81, 98, 213 grids: abstract, 159–60; affective power of, 72–74, 78–79; associative power

INDEX · 255

of, 102; controlling, 66, 70; as design strategy, 69–70; hierarchical, 60; imperial, 55; life in, 6, 95, 211–12; modular aesthetics of, 79; ontology of, 98–100; problematic assumptions and, 55; regulatory power of, 16; role of, 18, 98; spectacle of, 4–6; thinking in terms of, 131; types of, 125; using, 6, 89 (fig.), 98–99, 119, 194; as visualization tool, 55 Grids (Vignelli), 80 Grid Systems in Graphic Design (Müller-Brockmann), 69–70 Guardian, 196 Guattari, Félix, 93, 123, 225n11 Habeas Viscus (Weheliye), 123 Haeckel, Ernst, 18, 33, 45, 120, 134, 138, 169; biogenetic law and, 37; biology and, 29, 39; carbon and, 39; on chemico-physical properties, 39; cosmology of, 40; cover by, 34 (fig.); curved lines and, 35–36; developmental series and, 46–47, 49, 55; development and, 31–32, 164, 165; embryological images and, 55; evolution and, 37, 40; form and, 32, 165–66; illustration by, 48 (fig.); images by, 36, 68; monism and, 37, 38, 49; morphology and, 33, 142; narrative logic of, 50; organisms and, 164; protoplasm and, 44, 186; sequence and, 32; strategies of, 38, 46; vitalism and, 38; white exceptionalism and, 50–51; work of, 15, 44 “Haeckel’s ABC of Evolution and Development” (Richardson and Keuck), 32 Haeckel’s Embryos (Hopwood), 55

Hall, Stephen S., 145 halteres, 148, 151, 179, 242n59; transformation of, 150 (fig.) Hardy, G. H., 186 Harrington, Anne, 51 Hazen, Eric, 169 Hedison, Al, 110 Held, Lewis, Jr., 64, 138; illustration by, 60, 61 (fig.), 62, 63 (fig.), 65 Hendershot, Cyndy, 232n8 heredity, 169, 244n24; chromosomal theory of, 145 Hertzfeld, Andy, 89 Hierarchy Theory (Ahl and Allen), 120 HIV integration, 204, 205, 206 (fig.) “HIV Integration,” 204 Hodgson, G. M., 236n63 holism, 40, 51, 52, 227n28; organic, 8, 66, 173 Holland, W. F., 148 homeobox protein, 165, 207 (fig.) homeomutants, 174, 179 homeostasis, 22, 94, 113, 114, 237n69; biopower and, 118; politics and, 119; security and, 100–102 Homo Sacer (Agamben), 115 Homo sapiens, 49 Hopwood, Nick, 55, 56, 228n50; developmental series and, 47, 48; fraud and, 47 House Fly, The (film), 105 “How Bilaterians Use Hox Genes to Specify Area Codes along Their Anterior-Posterior (A-P) Axis” (Held), 6 How the Snake Lost Its Legs (Held), 60, 138; illustration from, 61 (fig.) How to Draw Comics the Marvel Way (Lee and Buscema), 25; illustration from, 28 (fig.) HOX-81, model of, 207 (fig.)

256 · INDEX

Hox genes, 60, 61 (fig.), 62, 64, 65, 161; grid system for, 63 (fig.) Hox organization, model of, 166 (fig.) human/fly hybrid, 112, 125, 216; human-sized, 110, 111 (fig.) Human Genome Project, rough draft of, 112 illustrations, 5, 17, 65, 200, 230n16; advertising, 73; biological, 32; medical, 32 images, 3, 14, 25, 28, 56, 146, 173; action, 18; aesthetics of, 15; animated, 172, 184, 194, 208; associative logics of, 82; background, 189 (fig.); biology and, 56, 120–21, 169; composite, 168, 172, 194; depth/movement in, 190; digital, 11; dynamics of, 66, 68; envisioning, 178; industrial economies and, 9–10; lines and, 154; magic of, 72; making, 10–11, 16, 69; meanings from, 67; multiplanar, 172; photographic, 85; political economic and, 10; potential, 84; power of, 14, 29, 68; proliferation of, 15; realism of, 178; regulating change through, 188–90, 192–95; scientific, 29, 170–71; space/time and, 67–68; subplanar, 172; surfaces of, 17; technical, 11, 14, 15–16, 18, 23, 83–87, 88–89, 90, 163, 177–81, 182, 207; traditional, 10, 11, 13, 163 immunitas, 115, 116, 119, 120, 121 individuation psychique et collective, L’ (Simondon), 19 industrial economies, images and, 9–10 industrialization, 3, 9–10, 15, 55, 70, 74, 85, 90, 126; bodies and, 233n15; criticism of, 99; heredity and, 244n24; principles of, 188 information: biology and, 144; changes

in, 220n7; communicating, 67; particles of, 86; regulation of, 9; spatial/relational, 68 information theory, 21, 22, 144 innovations, 4, 8–9, 55, 74, 96, 213; conceptual, 2, 131, 137, 171; technological, 2 In the Blink of an Eye (Murch), 183 Into the Universe of Technical Images (Flusser), 126 Iranbakhsh, Alireza, 19 Iwasa, Janet, 200, 246n44; on animation, 171; animation/scientific research and, 203, 204; fluidity and, 204; molecular movement and, 206; software program by, 205 Jacob, François, 22 jellyfish, 37; reproduction cycle of, 37 (fig.) Jenkins’ Groceteria, 71 (fig.) judgment, 30, 44, 208; aesthetic, 29, 41, 43; determinative, 41; reason and, 230n18; reflective, 41–42; teleological, 40, 41–43 Kant, Immanuel, 30, 40, 165, 225n8, 227nn28–29; analytical scheme of, 43–44; beautiful/sublime and, 227n38; beauty and, 42; determinative judgments and, 41; Haeckel and, 43; Newton and, 44; on purposiveness, 42; teleological judgment and, 43 Kay, Lily: information theory and, 21–22 Keller, Evelyn Fox: Morgan and, 224n54 Kemp, Martin, 79 Kendrew, John, 20 Kentridge, William, 190, 245n32;

INDEX · 257

aesthetics and, 188; animation by, 184–85; film by, 188; still form, 185 (fig.) Keuck, Gerhard, 32 Khalip, Jacques, 227n29 Kidd, Chip, 74, 81 Kinney Shoes, 73 Kraft products, 71 (fig.) Kunst-Formen der Natur (Art Forms of Nature) (Haeckel), 15, 32, 33, 35, 37, 44, 49, 50, 53, 169; cover of, 34 (fig.) Kurokawa, Kisho, 103, 234n29 Kwinter, Sanford, 53, 134 labor: clerical, 21; conceptual, 21; hier­ archies of, 189; partitioning, 190; systems of, 23 Lamarre, Thomas, 18, 169–70, 172, 194 larvae, 13, 147 (fig.), 155, 174, 201, 202; fly, 59, 156, 161; hatching of, 105–6 Latour, Bruno, 223n46 Laubichler, Manfred, 129 law of specialization, 48 (fig.) “Layout” (Wong), 81 Lears, Jackson, 78 Ledley, Robert, 20 Lee, Stan, 25; illustration from, 28 (fig.) Legos, 5 Lewis, Edward B., 20, 129, 130, 136, 160, 166, 174, 181, 205, 241n48, 242n66, 244n23; abstract grids and, 159; add-back genetics and, 162; animated film by, 202 (fig.); animation and, 18, 167, 169, 170, 176, 177, 195, 200, 203; Bateson and, 145, 164; bithorax complex and, 137, 148, 152, 201–2; death of, 145; development and, 155, 158, 165; diagrammatic rendering by, 203; envisioning by, 163; eukaryotic organisms and, 146;

experimental strategy of, 161; film by, 171, 179, 180, 180 (fig.), 200; gene expression and, 153–54; genetic regulation and, 176; Hox genes and, 161; Hox organization working model of, 166 (fig.); illustrations and, 17; images of, 179–80; lecture by, 167, 173; modularity and, 154; modules and, 166; Nobel Prize for, 146–47; on normal forms, 213; organisms and, 155, 164; phenotypic traits and, 17, 155; proteins and, 153; segmentation and, 156, 241n59; variations and, 145, 148; vision of, 165 life, 29, 93; aesthetics and, 134; cycles, 212; diversity of, 118; durations, 212; illustrating, 25; metaphysics of, 187; politics and, 116, 117, 119; power and, 94; regulation of, 184; unified theory of, 117; visions of, 18 “Life Cycle of Drosophila, The,” 174; stills from, 175 (fig.) Life of Forms in Art, The (Focillon), 30 lines, 28; curling, 40, 46; curved, 35–36, 46, 148, 169; images and, 154; swirling, 46 Linnaeus, Carl, 226n20 Liotta, John, 103 Loyer, Erik, 89 Lutz, E. G., 189, 190; animation by, 189 (fig.), 191 (fig.), 193 magic, 15, 66, 67, 68, 72, 109, 154, 184 Making and Unmaking the Grid (Samara), 123 Malthus, Thomas, 236n63 Manovich, Lev, 243n3 marketing, 3, 4, 70, 72, 96 Marriage of Heaven and Hell, The (Blake), title page of, 27 (fig.)

258 · INDEX

Materials for the Study of Variation (Bateson), 113–14, 139, 141, 142, 164; illustration from, 141 (fig.), 142 (fig.) mathematics, biology and, 187 Matthaei, Johann, 22 Mavhunga, Clapperton, 115, 235n55 Maya, 170, 194, 196, 198–99, 201, 205 McClintock, Barbara, 151 McCloud, Scott, 89 McGinnis, William, 165 McLuhan, Marshall, 14, 78 media theory, 10, 13, 23, 56 medusa, 36, 37, 46, 186, 226n22 Mendel, Gregor, 150 metaphors, 2, 52, 209; computational/ informational, 19–20 metaphysics, 38, 184, 187 microscopy, confocal, 12, 13 Miller, Gordon L., 226n20 misregulation, 16, 96, 104, 106, 110, 125, 174 Mitchell, Robert, 227n29 modeling, 207; dynamic scientific, 200–203; hypothetic, 200; molecular, 20 Modern Times (film), 99 modularity, 154, 155, 161, 166; biological science of, 56; bodies and, 135–36; coherent theory of, 144; development of, 17, 131, 137; historiographic paradox of, 134; importance of, 176; nonlinear nature of, 133; theories of, 5–6, 129 modules, 5, 57, 59, 131–37, 165, 166; aspects of, 131, 132, 137; making, 18 molecular dynamics, 18, 172, 198, 208–9 Molino, Jean, 30, 31 monism, philosophy of, 37, 38 Monist League, 228n53 Monod, Jacques, 22, 169

Morgan, Thomas Hunt, 145, 224n54 morphology, 30, 33, 142, 169, 240n30; biology and, 53; German, 115; teleological judgment and, 40 Morton, Marsha, 226n22 movement, 190; in animation, 192; complex, 193; molecular, 198, 206; periodicities of, 193; scientific, 52; sources of, 168 Müller-Brockmann, Josef, 69–70 Mulvey, Laura, 244n13 Murch, Walter: filmmaking and, 183 Musca domestica, 104, 106 Muscoidea, 106 “Music in 12 Parts” (Glass), 165 mutants, 142, 152, 160, 174, 242n65; bithorax, 163, 174, 179; drosophila, 153; homeotic, 202 “Mutants of the Bithorax Gene Complex” (Lewis), 174, 179; stills from, 180 (fig.) mutations, 113, 114, 149, 163, 174, 176, 179; genes and, 152; phenotypic effects of, 177; pseudoallelism and, 151 “Mutations Affecting Segment Number and Polarity in Drosophilia” (Nüsslein-Volhard and Wieschaus), 59, 165; illustration from, 59 (fig.) narratives, 20, 55, 56, 57, 59, 60, 65–66, 67, 74, 78, 88, 90, 174, 179, 183, 238n6 National Science Foundation, 37, 196 natural history, biology and, 208 natural selection, 44, 139 nature, artforms of, 33, 35–49 Nature (journal), 13, 79, 136, 145, 148, 196 Nature Methods (journal), 81 Natürliche Schöpfungsgeschichte (The

INDEX · 259

History of Creation), 32, 52; frontispiece from, 53 (fig.) Nealon, Jeffrey, 122 neoliberalism, 16, 119, 125 Neumann, Kurt, 91 Newton, Isaac, 44 New York Botanical Garden, 80 Nirenberg, Marshall, 22 Nocek, Adam, 234n27, 244n15, 247n55 nucleic acids, 7, 8, 38, 146, 152 Nüsslein-Volhard, Christiane, 136, 165, 167; embryos and, 57, 60; grids and, 57, 59; illustration by, 59 (fig.) Nye, Ben, Sr., 110 Oberkampf manufactory, 98 objects, motion of, 192 (fig.) Oken, Lorenz: morphology and, 33 Ontology of the Photographic Image, The (Bazin), 177 organisms, 1, 15, 25, 28, 39, 40, 44, 53, 62, 94, 99, 110, 112, 113, 120, 136, 137, 138–39, 152, 158, 174, 176; building, 7, 131, 155, 164; changes in, 134, 139; codes and, 21; conceptions of regulation of, 114–15; development of, 8, 9, 90, 115, 129, 133–34, 160, 195, 213; eukaryotic, 146; form of, 18–19, 46; function of, 18–19, 46; grids and, 5, 64; modular, 133, 161; organically integrated, 161; phenotype of, 150; sectioning off, 140; well-ordered, 173 Orr, H. Allen, 242n66 Osborn, Henry Fairfield, 32 Painleve, Jean, 174, 244n16 panels, 5, 17, 52, 56, 57, 62, 72, 74, 81, 88–89, 90, 132, 133; phenomenological power of, 88

Pardee, Arthur, 22 particles, 11, 13, 22, 83, 85, 86, 87, 109, 110, 126, 131, 155, 162, 163, 164, 170, 181, 182; material, 186; regulating, 84 Pasteur Institute, 22 patterns, 140; creating, 181; expression, 157 (fig.); mechanism of, 204 pests, 104–6, 127, 173 Phaeodarea, 44, 45 (fig.) phenomenology, 10, 16, 66, 90, 207 phenotypes, 17, 150, 154, 155, 160, 161; changes in, 151; stylized, 162 (fig.) photography, 84, 183; aesthetic qualities of, 177; live-action, 20; mechanical act of, 178–79; realism and, 177, 178; stop-motion, 182 Photoshop, 194 phylogeny, 32, 37 physics, 43; biology and, 169, 208 Plastic Man, 74, 76 (fig.), 77 (fig.), 78 Plastic Man (Cole), 74; page from, 76 (fig.) Playboy, 91 political economy, 11, 16, 22, 67, 78, 83, 90, 101, 105, 144, 184, 187, 190, 208; animation and, 208; history of, 236n63; images and, 10 political engagement, 116, 121, 122 politics, 50–51, 66, 94, 95, 116, 118, 126–27; aesthetics and, 124; animation and, 173; forms of, 31, 123, 126; homeostasis and, 119; life and, 117, 119; racial, 52; theories of, 121 power: hierarchies of, 54; life and, 94; sovereign, 116; visualizing, 126 Procter & Gamble, 73 production, 79, 126; film, 183; industrial, 71; regulation of, 102 product promotion, 70, 71–72, 71 (fig.) projection, 177–81

260 · INDEX

proteins, 11, 38, 152; clusters of, 121; genes and, 153; isolating, 1; modification of, 160; mutation of, 202 protoplasm, 38, 44, 186 pseudoallelism, mutations and, 151 purity, conception of, 133 Pymol, 171 Quaife, Veronica, 127 race, 52, 123 race studies, critical, 23 racial discourse, aesthetics of, 124 racial oppression, 123–24 racism, 50, 52 Raengo, Alessandra, 124 Rainger, Ronald, 32 realism, 155, 177, 179; ontological, 178; photography and, 178, 184; scientific, 203 Reenchanted Science (Harrington), 51 regulation, 100, 104, 107, 109–15, 124, 130, 144, 184, 206, 207, 221n15; concept of, 7, 221n18; dynamics of, 173; genetic, 7, 8, 62, 176; neoliberal form of, 102; politics of, 121; role of, 91, 94, 122; shift in idea of, 8; strategies of, 31; subtle forms of, 6–7; varieties in, 6–9 “Regulation of the Bithorax Gene Complex (a model)” (Lewis), 176, 200, 202 (fig.) relationships, 4, 10, 11, 16, 23, 29, 31, 32, 33, 35, 40, 43, 44, 45, 62, 65, 66, 70, 73, 84, 85, 86, 89, 91, 93, 94, 95, 102, 103, 107, 108, 111, 115, 118, 120, 121, 123, 130, 131, 138; aesthetic, 30; arranging, 113; biological, 3, 30, 53, 54, 144; compositional, 52; dialectic, 134, 164; formal, 50, 52, 154; grid, 19, 124–27; Kantian, 132;

logical, 144; narrative, 78; political, 117, 119, 208; racial, 52; teleological, 114 repetition, 67, 72, 130, 132, 137, 142, 242n69; law of, 139; as organizing principle, 140 Repetition of Parts, 140, 142 representation, 86, 177–81, 200 “Reproductive Cycle of the Jellyfish,” 37 (fig.) Return of the Fly, The (film), 127, 173–74, 176 Richards, Robert, 47 Richardson, Michael, 32 ripples, 130, 186, 187 RNA, 1, 8, 206 Robinson, Miranda, 145 Rodin, Auguste, 25; statue by, 26 (fig.) Rosenbleuth, Arturo, 22 Roux, Wilhelm, 51 Russell, Andrew, 131, 133 Russell, Bertrand, 160 Saint-Hilaire, Etienne Geoffroy, 240n30 Samara, Timothy, 58, 123 Satyrus hyperantus, 141, 141 (fig.) Sauer, Jeff, 168 Schaper, William, 134 Schiller, Friedrich, 225n8 Schlosser, Gerhard, 131 Schmidgen, Henning, 118 science: art and, 31; as constructivist practice, 16; sociologies of, 20; surfaces of, 79–82; technology and, 86; visual of, 25 Science (journal), 13, 148; advertising revenue for, 80; makeover for, 79–80; online, 87–90; screen capture for, 89 (fig.) Scientific Enterprise, The (Russell), 160

INDEX · 261

scientific publications, 13, 16, 19, 82; grids and, 79, 80, 82 (fig.) screens, 68, 91, 92–93, 104, 105, 109, 115; mobile, 103; use of, 93–94 security: flexible spaces of, 102–3; homeostasis and, 100–102; systems of, 206 Security, Territory, and Population (Foucault), 16, 101, 102 segmentation, 64, 65, 138, 155, 157 (fig.), 158, 187, 187 (fig.) segments, 147, 156, 187; abdominal, 160 (fig.); anterior, 159; changes in, 201; depiction of, 159; development of, 158–59, 159 (fig.); expression of, 160 (fig.); thoracic, 148, 157 (fig.), 160 (fig.) sequences, 62; developmental, 18, 48, 55, 137; form and, 32; genomic, 113; nucleotide, 13 Serres, Michel, 122 Seyyedrezaei, Seyyed Hassan, 19 Shaviro, Steven, 227n38 Simon, Joe, 73 Simondon, Gilbert, 19 Siphonophorae, 36 Smith, Adam, 236n63 social control, 119, 229n7 Social Darwinism, 49 social operations, effective, 100 social organization, rules/laws in, 6 social science, research in, 96 sociology, 20, 21, 223n46 software: animation, 196, 205; mastering, 205; 3D, 198–99; visualization, 171, 205 space, 78; material orders of, 17; organic conceptions of, 120, 234n29; perceptions of, 245n27; Western/ Japanese conceptions of, 234n29 speciation, 114, 139, 169

specimens, sequential development of, 48–49 spiders, 94, 120, 125, 180 Spiegelman, Art, 74, 81 stacking, 71–72, 71 (fig.) standardization, 2, 3, 6, 9, 21, 133 Stern, Curt, 107–8, 109, 121 Stiegler, Bernard, 178 Stoltz, Eric, 127 Strasser, Susan, 70 structural biology, 8, 208, 246n44 structure: change and, 29–30; segmental, 164; unsegmented, 187 (fig.) “Stylised Phenotypes Detectable in First Interstar Larvae or Mature Embryos,” 162 (fig.) subjectivity, 94–95, 221n17 surfaces: concentration across, 122 (fig.); difference and, 122 (fig.); magic of, 66–68; multiple, 121; politics of, 127 symmetry, law of, 139 synthetic biology, 2, 8, 213 systems theory, 2, 144 Tarde, Gabriel, 242n69 Tatum, E. L., 151 T Cell, death ligand from, 198 technical images, 11, 16, 18, 88, 132, 207; codes of, 15; described, 83; imperatives of, 89–90; surfaces of, 83–87; understanding, 14; value of, 84 technology, 11, 14, 15, 91, 100, 101, 109–10, 207; biology as, 2; communication, 95; entertainment, 95; imagining, 126; science and, 86; systems of labor and, 23; visualization, 4, 7, 10, 146, 169 teleological judgment, 40, 43, 44, 54 teleportation, 16, 91, 97, 101 Thacker, Eugene, 21

262 · INDEX

Them! (film), 180 Theory of Natural Selection, 44, 139 thought, fluid nature of, 203–7 Thousand Plateaus, A (Deleuze and Guattari), 93 Tide Table (film), 245n32 “Tide Table” (Kentridge), scene from, 185, 185 (fig.), 187–88 transformations, 110, 184, 186, 241n48 Tremont Institute, pragmatics and, 223n46 tusks, segmentation and, 187 (fig.) tweeners, 190 tween frames, 194 20th Century Fox, 16, 91 “Two or Three Bristles” (Stern), 107–8 ultrabithorax, 148 understanding, 3, 5, 6; making and, 2 Understanding Comics (McCloud), 89 urbilaterans, 62, 63, 64, 66 usefulness, 65, 69, 89, 98, 99, 182 variations, 137–45, 238n12; homeotic, 148; mod 70s, 145–56, 158–66; study of, 140–41; types of, 17 Vashavsky, Alexander, 145 Vectors (Loyer), 89 Venter, J. Craig, 112 Versuch de Metamorphose der Pflanzen zu erklären (The Metamorphosis of Plants) (Goethe), 35 Vignelli, Massimo, 70, 80, 81 Village of the Giants (film), 180 virtual engagement, 9, 122

visual elements, 57, 74, 194 visualization, 3, 4, 8, 181, 208, 247n55; envisioning and, 9–14 visual worlds, dynamic/complex, 194 Vit, Armin, 69 vitalism, 37, 38, 43 Von Baer, law of specialization and, 47, 48 (fig.) Wagner, Günter P., 131 Walas, Chris, 127 Walter and Eliza Hall Institute of Biomedical Research, 196 Ward, Andrew, 227n29 warped grids, 119–22, 123–24 wave form, complex, 192–93, 193 (fig.), 194, 246n40 Weheliye, Alexander G., 23 Weiner, Norbert, 22 West, Luther S., 93, 106 Who Wrote the Book of Life (Kay), 22 Wicks, Robert, 42 Wieschaus, Eric, 59, 60, 136, 165, 167; illustration by, 59 (fig.) Wisdom of the Body, The (Cannon), 22 Wolfle, Daele, 79 Wong, Bang, 81; depiction by, 82 (fig.) Woolgar, Steven, 223n46 world building, 2, 3, 8, 10, 19, 243n3 Worldwide Protein Databank, 205 Worringer, Wilhelm, 30, 225n11 Wright, Sewall, 235n53 zebrafish, embryonic development of, 57, 58 (fig.)

(continued from page ii)



23 Vampyroteuthis Infernalis: A Treatise, with a Report by the Institut Scientifique de Recherche Paranaturaliste Vilém Flusser and Louis Bec



22 Body Drift: Butler, Hayles, Haraway Arthur Kroker



21 HumAnimal: Race, Law, Language Kalpana Rahita Seshadri



20 Alien Phenomenology, or What It’s Like to Be a Thing Ian Bogost



19 CIFERAE: A Bestiary in Five Fingers Tom Tyler



18 Improper Life: Technology and Biopolitics from Heidegger to Agamben Timothy C. Campbell



17 Surface Encounters: Thinking with Animals and Art Ron Broglio



16 Against Ecological Sovereignty: Ethics, Biopolitics, and Saving the Natural World Mick Smith



15 Animal Stories: Narrating across Species Lines Susan McHugh



14 Human Error: Species-­Being and Media Machines Dominic Pettman



13 Junkware Thierry Bardini



12 A Foray into the Worlds of Animals and Humans, with A Theory of Meaning Jakob von Uexküll



11 Insect Media: An Archaeology of Animals and Technology Jussi Parikka



10 Cosmopolitics II Isabelle Stengers



9 Cosmopolitics I Isabelle Stengers



8 What Is Posthumanism? Cary Wolfe



7 Political Affect: Connecting the Social and the Somatic John Protevi



6 Animal Capital: Rendering Life in Biopolitical Times Nicole Shukin



5 Dorsality: Thinking Back through Technology and Politics David Wills



4 Bíos: Biopolitics and Philosophy Roberto Esposito



3 When Species Meet Donna J. Haraway



2 The Poetics of DNA Judith Roof



1 The Parasite Michel Serres

PHILLIP THURTLE is associate professor in the Comparative History of Ideas program and in the history department at the University of Washington.
[Grid] Biology in the Grid [Phillip Thurtle]

Related documents

273 Pages • 94,683 Words • PDF • 12.5 MB

146 Pages • 60,378 Words • PDF • 6.8 MB

462 Pages • 197,484 Words • PDF • 5.8 MB

8 Pages • 5,110 Words • PDF • 1 MB

330 Pages • 105,680 Words • PDF • 16.8 MB

1,465 Pages • 883,178 Words • PDF • 69.8 MB

895 Pages • 455,245 Words • PDF • 67.7 MB

48 Pages • 26,163 Words • PDF • 2.3 MB

6 Pages • 2,990 Words • PDF • 1.5 MB

1,442 Pages • 978,819 Words • PDF • 114 MB

1,253 Pages • 580,073 Words • PDF • 188 MB