507 43 4MB
English Pages 351 Year 2009
ReAction!
This page intentionally left blank
ReAction!
Chemistry in the Movies
Mark Griep and Marjorie Mikasen
1 2009
1 Oxford University Press, Inc., publishes works that further Oxford University’s objective of excellence in research, scholarship, and education. Oxford New York Auckland Cape Town Dar es Salaam Hong Kong Karachi Kuala Lumpur Madrid Melbourne Mexico City Nairobi New Delhi Shanghai Taipei Toronto With offices in Argentina Austria Brazil Chile Czech Republic France Greece Guatemala Hungary Italy Japan Poland Portugal Singapore South Korea Switzerland Thailand Turkey Ukraine Vietnam
Copyright © 2009 by Oxford University Press, Inc. Published by Oxford University Press, Inc. 198 Madison Avenue, New York, New York 10016 www.oup.com Oxford is a registered trademark of Oxford University Press 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 permission of Oxford University Press. Library of Congress Cataloging-in-Publication Data Griep, Mark. Reaction! : chemistry in the movies / Mark Griep and Marjorie Mikasen. p. cm. Includes bibliographical references and index. ISBN 978-0-19-532692-5 1. Science in motion pictures. I. Mikasen, Marjorie. II. Title. PN1995.9.S265G75 2009 791.43⬘66—dc22 2009007872
9
8
7
6
5
4
3
2
1
Printed in the United States of America on acid-free paper
For our parents
This page intentionally left blank
Acknowledgments
We thank the Alfred P. Sloan Foundation’s Public Understanding of Science and Technology program and Doron Weber, its program director, for generously supporting this project. We also thank the University of Nebraska–Lincoln Chemistry Department for believing in the merits of this project from the beginning. The department provided funds to purchase videos, provided feedback about M.G.’s themed lectures, allowed M.G. to teach a course titled “Chemistry in the Movies,” and helped us in the hunt for movies with chemical themes. Special thanks to Jeremy Lewis, Ned Sears, Brian Desmond, and Trish Watson at Oxford University Press. The following archives, librarians, and archivists provided invaluable help (in alphabetical order): Margaret Adamic, Corporate Legal Department, Disney Enterprises, Inc.; Karla Buhlman, Gene Autry Entertainment; Ned Comstock, archivist, University of Southern California (USC) Doheny Library; Michael Duchemin and Marva Felchlin, Autry National Center; David Edgar, British Film Institute, London; Mark Gens, UCLA Library Special Collections; Danny Ladely, director, and Bill Fetch, Mary Riepma Ross Media Arts Center, Lincoln, Nebraska; Sandra Joy Lee, director, Warner Bros. Archives, USC School of Cinematic Arts; Mark Quigley, manager, UCLA Film and Television Archive; Jenny Romero, Special Collections, Margaret Herrick Library; and Dave Smith, director, Walt Disney Archives. We also thank the staff at the British Library, London, and the staff at Love Library, University of Nebraska–Lincoln. We have enjoyed watching movies for as long as we remember. We thank the following people and events for inspiring us to enjoy movies and to think more deeply about their meaning. In 1979, M.M. took a course from Tom Conley titled “Italian Neo-Realist Cinema” at the University of Minnesota. M.G. snuck in to watch two of the fi lms. In 1987, we attended Joshua Hassel’s “Pop into Film” presentation in Denver. It was sponsored by the Metropolitan Denver Arts Alliance in conjunction with a juried art exhibition titled “Communications ’86.” Hassel used short clips from 1960s popular and art fi lms to demonstrate the many uses of art in fi lm. From him, we learned of Sixth Avenue Video, run by Sonja, who rented a great collection of movies on video. In 1998, M.G. fi rst read about the National Institutes of Health’s “Screening Science” series and wondered
viii
Acknowledgments
how to bring something like that to the Chemistry Department at the University of Nebraska–Lincoln. In 2000, we discovered that Elvis was a chemist in Clambake. In 2001, John Fortman presented a wonderful talk in Hastings, Nebraska, titled “Serious and Delirious Uses of Chemistry in the Movies” that was sponsored by the Nebraska Local Section of the American Chemical Society. He paired movie clips with a chemical magic show. In 2007, independent fi lmmaker Jon Jost included us in his fi lm-in-progress Swimming in Nebraska, when he was shooting in Lincoln. We immersed ourselves in the vision of this singular director and learned a bit about making movies at the same time. The quote (page 81) “It seems to me . . . answers could save us,” is from page 42 of PRISONS WE CHOOSE TO LIVE INSIDE by DORIS LESSING. Copyright © 1988 by Doris Lessing. Reprinted by permission of HarperCollins Publishers. Additional territory: Jonathan Clowes, Ltd., Iron Bridge House, Bridge Approach, London NW1 8BD, England.
Contents
Introduction The Dark and Bright Sides of Chemistry in the Movies, 3 Chapter 1 Dr. Jekyll’s Mysterious Transformative Formula, 9 Chapter 2 Invisibility Steals the Seen: Chemistry Creates Criminal Opportunities, 35 Chapter 3 Isomorphs of Paranoia: Chemical Arsenals, 65 Chapter 4 Bad Company: The Business of Toxicity, 103 Chapter 5 A Master/Slave Narrative: Drug Addiction and Psychoactives, 134 Chapter 6 Inventors and Their Often Wacky Chemical Inventions, 155 Chapter 7 Hard Science = Hard Evidence: Forensic Chemistry and Chemical Detectives, 186 Chapter 8 Chem 101: Learning by Doing, 221 Chapter 9 Good News: Research and Medicinal Chemists Making a Difference, 250 Chapter 10 First, Do No Harm: (but Before That, Self-Experiment), 284 Conclusion Chemistry in the Movies, 295 Appendix 1 How to Use This Material in the Classroom, 297 Appendix 2 About the Back Cover Art, 301 References, 303 Movie Index, 323 Subject Index, 331
This page intentionally left blank
ReAction!
This page intentionally left blank
Introduction The Dark and Bright Sides of Chemistry in the Movies
Almost a decade ago we sat down to watch some pure home video entertainment: Clambake, starring Elvis Presley. Halfway through the movie, to our complete surprise, Elvis turned into a chemist! You might say this book began at that moment. In the following pages, we examine the presence of chemistry in one of the most accessible of all cultural products: movies. It has been noted repeatedly in recent years that chemistry is one of the least popularized areas of the hard sciences. When the term “science” is used in popular culture and the media, it often refers to medicine, physics, and biology. Monsters and superheroes, mad scientists and geniuses, aliens and mutants—all spring with ease from these realms. The discipline of chemistry seems by comparison to be underrepresented in cultural depictions, with an appearance harder to trace and an impact less openly acknowledged. Is this perception truly accurate? One need only name names—Dr. Jekyll and Mr. Hyde—to see that this can hardly be the case. From the silent era through today, it is clear that chemistry has always been in the movies. In fact, chemical themes and characters have been capturing the imaginations of audiences for more than a century. They appear in surprising and significant ways and have generated some of the most enduring fictions and motifs in movie history, and thus in the culture at large. In its starring role, chemistry, the transformative science, has been moving us as it changes with the times. More than 1,200 motion pictures were compiled, analyzed, and categorized for this project. The sheer quantity of movies containing some aspect of chemistry was eye-opening. What started out as an exercise in curiosity, casually noting the appearance of chemistry themes through random movie viewing, quickly turned into serious study as the numbers kept rising. Inspired by the National Institutes of Health (NIH) “Science in Cinema” fi lm series, we started making a list. Every summer since 1998 (Zurer 1998), the NIH has screened six fi lms, each dealing with a different medical theme, followed by a short scientific analysis from an NIH researcher working in a relevant field. As our own list grew larger, patterns were noted, a web of associations emerged, and trends became apparent. 3
4
ReAction! Chemistry in the Movies
The two-part structure of this book is driven by what the movies say about themselves. It is clear from the examples on our list that chemistry is portrayed nearly equally as positive or negative in movie story lines. The visual image of a monad comes to mind, a circle with two equal, intertwined parts. The halves of this whole are entangled and interdependent. Each half has its identity amplified and clarified by its relationship with the other. It stands to reason that bad actions and outcomes would not seem so bad if not contrasted with good actions and outcomes, and vice versa. Thus, this book has two parts: the dark side and bright side of chemistry in the movies. Our book can also be seen to embody the duplex chemical character that is Dr. Jekyll and Mr. Hyde and has taken its template from him. Recently, an advocate for the discipline of chemistry called for more positive, accurate images of chemistry, decrying the misrepresentation of science shown in the Jekyll and Hyde story (Emsley 2006). While one may sympathize with this sentiment, and it may be persuasive on a public relations level, to adopt this viewpoint would gloss over the difficult aspects of human nature. Would it not impose a constrained view on a scientific discipline founded on the principles of free inquiry? In truth, the negative and positive aspects of human needs and desires have always played off of one another, and we ignore this at our peril. There is also a long history of acknowledgment and educational focus in the sciences that connects ethical action to individual responsibility, much of which has evolved from grappling with the enormously destructive impacts of such scientific discoveries as the use of poison gas in WWI or the detonation of the atomic bomb in WWII. Warfare aside, science has made many benevolent but also many insidious inroads into our daily lives, causing us to wonder if our very sense of what it means to be human is at stake. Indeed, as noted in the 1995 National Academy of Sciences publication On Being a Scientist, “Science and technology have become such integral parts of society that scientists can no longer isolate themselves from societal concerns.” Nevertheless, scientists do not have concentrated training in the areas of ethics and morals. What scientists do well is derive information from the scientific process that can influence the moral values of individuals in society in the areas of religion, politics, and economics. The fi ndings of science are ethically neutral, but the activity of science is not (Bronowski 1965). Chemistry is the only science that shares its name with an industry, and as such, it carries the heavy associative burden of some negative uses of its discoveries. Since their beginning, motion pictures have mined the human condition for dramatic and comedic material. Anything human beings are capable of, including chemical capabilities, becomes fair game for storytelling. The somewhat innocuous subject matter of movies might seem an unlikely place to explore the complexity of moral issues and personal responsibility related to the science of chemistry. Taken as a whole, however, the movies described in this book show a wide range
Introduction
5
of possibilities of action and agency in the moral dimension. They give us examples of inner resources or abilities that are developed or neglected by the chemist-actors as they maneuver in situations that present them with competing loyalties. The dialectical arrangement of themes in the dark and bright sides also emphasizes that these themes stand in a relational way to each other and can illuminate each other. Indeed, this may lead to new insights and awareness on the part of the reader. We are not, however, endorsing a relativist position. In fact, we have deliberately reversed the expected coupling of bright and dark—we have intentionally given the bright side the last word. Chemistry is powerful: It has the power to damage, possibly beyond habitation, the planet that is our home. It can take life indiscriminately and ruin health and property. Could one ask for a more potent force? Its responsible use is necessary for our survival. We will need all of our value systems working together to create a better why, how, when, and where for the use of chemistry and its products in the future. Early in this project we asked friends and acquaintances to name some movies they had seen containing chemistry or references to chemistry, and most commonly they would say Frankenstein. The science featured in the 1931 Universal Studios movie production of Frankenstein is biology and physics. We wondered why it is so hard to for people to see the chemistry in a movie, why it is so easily overlooked. Maybe it is because chemistry is a practical science, or because chemistry uses symbolic language to convey specialized knowledge, and many people do not have this educational background. Yet the practice of chemistry has its basis fi rmly in reality. Since chemistry is sometimes misidentified with other disciplines such as physics, we also wondered why chemistry is, by implication, so frequently associated with the horror genre. The answer may be that it is science and one must keep in mind that the physics of Frankenstein and the chemistry of Dr. Jekyll and Mr. Hyde both reflect the “terror and promise of science” (Tudor 1989). The two-part character of Jekyll and Hyde is foreshadowed by dualities evident in the Dr. Frankenstein character, namely, his divorce of rational self from emotional self. The character of Dr. Jekyll is more psychologically developed, however, and he owes his separation of personality and compartmentalization of action to his chemical expertise. What both movies also share is that they are based on works of literature. The 1886 novella Strange Case of Dr. Jekyll and Mr. Hyde by Robert Louis Stevenson is the story of a physician/chemist bent on separating the good and bad aspects of human nature. There are many movie versions of this story, down to the present day. One of the most acclaimed is the 1931 version of Dr. Jekyll and Mr. Hyde featuring Fredric March, and it solidifies chemistry’s central role in cinema history. But the scope of this book goes beyond cinema history. We feel the Jekyll and Hyde duality of dark and bright holds a key to understanding
6
ReAction! Chemistry in the Movies
how chemistry is perceived in Western culture. We are using this duality to guide the structure and presentation of themes in our book. The fi rst five chapters have dark chemical themes, whereas the last five chapters have bright chemical themes. Each chapter begins with a chemical analysis of that theme and ends with chemical and narrative summaries of the movies contained within that chapter. The chemical analysis of each theme has five sections. The fi rst introduces the theme shared by the movies and chemistry. The second expands on the chemistry behind the source material, usually a book or personal biography, which spawned the majority of movie narratives in that theme. The third section deals with the real chemistry behind the movie narratives. Viewers may well ask: Can the chemistry in the movie really work? The fourth section discusses the psychological responses to the theme, broadening the context and positioning it within a continuum that includes societal, historical, and political dimensions. The fifth section describes the archetypal movie for that theme. The archetypal movies are selected from a group of about 12 movies that are summarized in each chapter. An “archetype,” a term that became widely disseminated through the writings of psychologist C. G. Jung, is the “possibility of representation” (Jung 1980). Traditionally depicted through mythology, religion, storytelling, and art, archetypes fi nd their primary contemporary expression through the mass medium of movies (Indick 2004). The contents that make up archetypes are motifs with typical images and associations, or archetypal ideas, and they have the power to influence and fascinate (Jung 1980). Certain movies are distinguished from others; they very clearly represent the core value and significance of the chemical theme, often exuding a vivid emotional tone. Using chemical imagery in his writings to describe how the framework of an archetype functions, Jung compares its form to the “axial system of a crystal” and notes that the ions and molecules aggregate in a “specific way” (Jung 1970). He suggests that there is something a priori, or a “preform,” that exists before the crystal has its own material existence. Each of our chapters can be thought of as the matrix on which a striking movie example crystallizes in an archetypal way. The movies in each chapter were selected by ranking them according to the following criteria: contemporary (meaning released after 1970), available on VHS or DVD, included women or other underrepresented groups in significant roles, or especially favored by one or both of the authors. The idea of an archetype fi lm came out of this ranking exercise after we discovered concurrence between the especially favored categories and some older fi lms. Even though featuring women protagonists was one of our selection criteria, there are only about a dozen movies in this book featuring women chemists as characters, and another dozen featuring women nonchemists in dominant lead roles. It is not surprising to any contemporary fi lm
Introduction
7
Figure 1. The percentage of physical science Ph.D.s awarded to women (solid circles) and the number of women appearing as chemists in the movies (open circles). The Ph.D. data are from “Doctorates in the 20th Century” (National Science Foundation 2006).
viewer that most movies are about men. It may be surprising to learn, however, that there were more women chemists on the screen in the 1930s through 1950s than there have been since (figure 0.1, open circles). This recent paucity contrasts very poorly with the documented rise in the percentage of physical science Ph.D.s earned by women in U.S. universities (figure 1, solid circles). Chemists earn about two-thirds of the physical sciences degrees. After the movies were selected for each chapter, we sought to describe the fundamental dark/bright duality of the fi lms in the book. It seemed that chemistry in the movies went beyond horror versus comedy, but it was not clear how this played out in the language of fi lm criticism. An examination of the fi lm genres found in the various chapters (table 1) revealed that the most common genre in the dark side is thriller, not horror. In fact, horror is less common than drama. The analysis confi rmed that the most common genre on the bright side was comedy, but additionally that the second most common was drama. Given that drama is found on both sides of our duality, the book’s dominant genre duality seems to be thrillers versus comedies. Either someone is going to be killed, or the chemistry provides a fi rm foundation from which to launch amusing or satirical comment. If instead we take into account that, among all movies listed on the Internet Movie Database (www.imbd.com), there are many
8
ReAction! Chemistry in the Movies
Table 1. Genre analysis of the movies in each chaptera Chapter
Chemical Theme
Movie Genre Signature
Jekyll and Hyde (9) Invisible man (7) Chemical weapons (15) Bad companies (12) Addiction (11)
9 horror, 5 sci-fi, 4 drama 6 sci-fi, 5 horror, 4 [black] comedy, 4 thriller 8 thriller, 6 drama 8 drama, 4 thriller 8 drama
Inventors (11) Forensics (13) In the classroom (11) Good researchers (12) Drug discovery (15)
8 comedy 7 drama, 7 mystery, 7 thriller, 5 action, 5 crime 7 comedy, 4 sci-fi, 4 romance 5 drama, 4 comedy 8 drama, 8 sci-fi, 6 horror, 6 comedy
The Dark Side 1 2 3 4 5 The Bright Side 6 7 8 9 10
a The genres were taken from the Internet Movie Database (www.imdb.com) in 2008. A “genre signature” is defi ned here as genres found in more than one-third of the movies in that chapter.
more dramas (121,000) and comedies (97,000) than romance (22,000), thrillers (18,000), horror (12,500), sci-fi (9,000), mystery (8,000), and biography (4,000), we learned that our “dark side” movies were rich in thrillers, horror, and sci-fi, whereas our “bright side” movies were rich in biography, mystery, and romance. This genre duality feels right in that there is an inherent danger to every discovery but it is accompanied by an element of excitement.
1 Dr. Jekyll’s Mysterious Transformative Formula
MIRRORS AND CHIRALITY “Jekyll and Hyde” is a phrase known to many, though few have read the short novella published in 1886. It is far more likely that people have encountered the phrase during conversation or in one of its numerous adaptations. In fact, the Strange Case of Dr. Jekyll and Mr. Hyde by Robert Louis Stevenson is the most adapted story of all time, even exceeding such texts as Mary Shelley’s Frankenstein, Charles Dickens’s A Christmas Carol, and Shakespeare’s The Tempest (Rose 1996). The idiom “Jekyll and Hyde” usually refers to someone or something that manifests its opposite tendency in different contexts. Colloquially, it does not always carry an explicit chemical connotation. But, in the more than 100 stage, movie, television, and cartoon adaptations (for a continually updated list, see Dury 2006), Jekyll is nearly always transformed into Hyde after ingesting or injecting a chemical formula of his own manufacture. For this reason, it is the single most important example of chemical self-experimentation in the movies. Nearly all of the dramatic Jekyll and Hyde adaptations have important scenes in which the mirror is used as a research tool (table 1.1). After Jekyll transforms into Hyde for the fi rst time, he determines that the experiment was a success by looking into a mirror. He sees the monstrous Hyde in the reflection and knows that he, Jekyll, no longer looks like himself. It is very likely he no longer even feels viscerally like himself. The transformation scene, preceding the mirror scene, often shows Jekyll painfully grimacing, shaking, or groaning. The mirror scene is the point of full realization. We can conclude that Jekyll’s mind, though now contained in the persona of Hyde for the fi rst time, is still able to internalize this realization with a scientist’s thinking process. As the story progresses, however, Hyde becomes increasingly more powerful and uses the mirror for self-satisfied confi rmation that he has trumped Jekyll yet again. The mirror is the axis on which the status of the Jekyll and Hyde character fl ips. The mirror scene initiates an understanding that Jekyll and Hyde function as a paired unit. Together they make a whole, yet each of them is a discrete physical entity in space and time, with competing interests. What is Dr. Jekyll’s original intention when he designs 9
10
ReAction! Chemistry in the Movies
Table 1.1. The first mirror scenes in the Jekyll and Hyde films discussed in this chapter Title (Year)
First Mirror Scene
Mary Reilly (1996)
13:00: Jekyll looks backward into bedroom cheval glass while asking Reilly to have Poole take it to the laboratory 23:30: Pride looks into bedroom wall mirror after fi rst self-injection 26:00: Blake admires his reflection as he carries the mirror from the drawing room to his laboratory 24:45: Jekyll struggles to drawing room after fi rst self-injection only to fi nd the reflection of Sister Hyde 1:14:30: Jekyll speaks to Hyde in small laboratory mirror as they each struggle for dominance 36:00: Jekyll drinks potion, hallucinates, and then awakens to admire his reflection in a laboratory mirror 5:00: As Jekyll leaves to give his speech at the medical college, he stops at hall mirror to straighten up 25:00: After transformation, Hyde enters drawing room to look at himself in a mirror we do not see 1:00: After fi rst transformation, Hyde peers into small laboratory mirror
Dr. Black, Mr. Hyde (1976) I, Monster (1971)
Dr. Jekyll and Sister Hyde (1971)
The Two Faces of Dr. Jekyll (1960) Dr. Jekyll and Mr. Hyde (1941)
Dr. Jekyll and Mr. Hyde (1931)
Dr. Jekyll and Mr. Hyde (1920) Dr. Jekyll and Mr. Hyde (1912)
his experiment? In the 1931 Paramount Pictures version, Fredric March as Dr. Jekyll tells us in a lecture, “Man is not truly one, but truly two.” He categorizes one side as noble and the other as basely impulsive. He further describes the sides as good and bad, respectively, and sees them as “chained together” in an ongoing struggle. He asks rhetorically how much better humans would be if the two selves were separated and the good self were allowed to be truly good, unencumbered by its evil counterpart. Dr. Jekyll seeks to separate the two selves that comprise one human being by utilizing the power of chemicals to affect a split. Thus, on its most primary level, the Jekyll and Hyde story illuminates the functioning dynamic of the pair and the concepts of symmetry and dominance. There are pairs found in nature. In one of the “space arrangement riddles” of chemistry, this molecular phenomenon is called chirality (Brown 2003). Two chiral molecules form a pair called enantiomers, Greek for opposite parts (enantios + mere). The mirror is the central metaphor for understanding the structure and conformation of chiral molecules in three-dimensional space. What a mirror really does is interchange front and back (Close 2000). When the axis perpendicular to the mirror reverses, an imaginary mirror-image space comes into being. The form
Dr. Jekyll’s Transformative Formula
4
*
11
1
1
3
3
2
*
4
2
Figure 1.1. A chiral compound has a central atom (asterisk) attached to four different atoms or groups of atoms. Two compounds that are mirror images of one another are called enantiomers.
reflected in this space is geometrically equal but not congruent with its “real” counterpart (figure 1.1). Chirality means “handedness”; chiral is from the Greek for hand (kheir). This term neatly conjures up our own body as a reference point for understanding. By invoking the bilateral, or two-part, asymmetry that is expressed in the external form of our own body, we see how fundamentally ingrained is left and right in the natural world. Consider our hands (or just as easily our two feet). Our pair of hands can be thought of as two three-dimensional solids of relative equality (Shubnikov and Koptsik 1974), but they are not identical because the left hand cannot occupy the same three-dimensional space as the right hand. One hand is not superimposable on the other; they do not have coincidence. It is only when the left hand is reflected on a mirror plane that it appears to have the same spatial orientation as the right hand. Our hands are similar yet also different in this important respect. Chirality holds life on Earth in a profound handclasp of parity and privilege. Though the origins are not fully understood, biological life shows a defi nite bias in favor of one molecule over the other in a mirror-image pair. There is the maxim, however, that if it is possible for two opposite forms of a molecule to exist, these are rarely used simultaneously; organic life will make a single choice (Whyte 1975). The difference between left- and right-handed molecules is fundamental to the efficient functioning of biological life. The bias of a single handed-form over the other is present in the building blocks of protein, DNA, and polysaccharides (Bernal et al. 1972). In general, life uses only left-handed amino acids and right-handed sugars. A significant property of chiral molecules is optical activity, which is displayed in relation to the structural asymmetry of the individual molecule. Is there an analogy from chirality that we can take to the movies? Hyde is conceived by Jekyll and then brought into physical being through experimental research that leads to the ingestion of a chemical substance. The mirror scene links Jekyll and Hyde together, giving us a
12
ReAction! Chemistry in the Movies
striking mental image of a similar pair possessing an oppositional undertone. Constrained by the properties that comprise our three-dimensional world, the opposing forms of Jekyll and Hyde cannot exist simultaneously in physical space. One occupies real space, and the other imaginary space. We can further extend the analogy to say that since Jekyll and Hyde is an active, living character with an enantiomer form, only one side of his handed state will be expressed at one time. Which bias will get the “upper hand”? The resulting movie story lines utilize this dynamic tension to create screen versions that twist and turn the location, duration, and conformation of Jekyll and Hyde. We don’t have to limit our discussion of chirality to analogies, however. We’ll see in the next section how a chiral molecule may hold surprising significance for interpreting the Jekyll and Hyde tale. This will be a book of correspondences, oppositions, and stresses on inflection. One depiction of chemistry will come face-to-face with another depiction of chemistry. In a broad sense, the movie screen is the mirror plane that reflects the constructs society brings to it. When Jekyll looks into the mirror and sees Hyde, he sees not only a transformation of his own making, but also a dilation on the scope of real and imaginary.
STEVENSON, HIS NOVELLA, AND THE CHEMISTRY OF THE HYDE FORMULA Even though this fi rst chapter is about the most referenced chemist in the movies, the dark Jekyll and Hyde movies have the least chemistry in them. Stevenson’s 1886 novella is a mystery/detective story in which the reader doesn’t know until the last chapter that Jekyll and Hyde are the same person. In his 1998 book about mad scientists, horror analyst David Skal says that the Jekyll and Hyde story “reflects the Victorian Era’s dilemma vis-à-vis evolutionary theory: how to reconcile the human personality’s higher and lower natures.” He notes that this meshed well with Victorian interest in balancing scientific reductionism with the craving for spiritual transcendence. Most classically horrific Dr. Jekylls are physicians carrying out pharmacological research with an unnamed compound that physically transforms them into someone who feels no remorse for the terrible acts he performs. We have a natural fear of the unknown that horror movies exploit. This is probably why most comedy Dr. Jekylls spend more fi lm time naming and explaining the nature of their Hyde formulas than do the dramatic versions (see chapter 8 for two examples). Because the compound is not named in the original or in the dramatic adaptations, it becomes possible to transfer our anxiety about any new psychopharmaceutical to the Hyde compound. If Stevenson had named the active ingredient in the Hyde formula, the fi lm adaptations might have told the tale of the chemical and not of the transformation. The work’s lack of defi nitiveness on this point
Dr. Jekyll’s Transformative Formula
13
gives power to the many re-adaptations. The book and movies are tales of science anxiety, not science. Dr. Jekyll rashly decides to self-experiment with a new chemical, and the experiment goes awry. On the other hand, some chemical and literary forensics allows us to identify the most likely candidates for the Hyde formula ingredients and even to speculate about the Hyde molecule itself. There is enough information provided in the novella to identify the “blood-red liquor” with near certainty. An examination of Stevenson’s life allows us to speculate about the “unknown impurity.” Together, this vastly enriches the scientific reading of this classic text by bringing its alchemical and gothic roots to the fore.
The Chemistry of the “Blood-Red Liquor” From “Dr. Lanyon’s Narrative,” the penultimate chapter of Stevenson’s novella (emphasis added): The powders were neatly enough made up, but not with the nicety of the dispensing chemist; so that it was plain they were of Jekyll’s private manufacture; and when I opened one of the wrappers I found what seemed to me a simple crystalline salt of a white color. The vial, to which I next turned my attention, might have been about half full of a blood-red liquor, which was highly pungent to the sense of smell, and seemed to me to contain phosphorus and some volatile ether. At the other ingredients, I could make no guess.
From this list it appears that there are three components to the Hyde formula: the white crystalline salt, the blood-red liquor, and “other ingredients.” The white crystalline powder is described in the next section, the blood-red liquor is identified in this section, and it is not necessary to describe the “other ingredients” because they are not used to make the Hyde formula. The color and smell of the “blood-red liquor” along with the specific mention of the element phosphorus suggests it is white phosphorus and gold chloride dissolved in carbon disulfide. Such a mixture produces a clear ruby red suspension of gold colloid particles. Carbon disulfide liquid was commonly used to dissolve compounds that weren’t particularly soluble in water. Today, it has been replaced by solvents such as dimethyl sulfoxide (DMSO). Phosphorus was one of the elements discovered by an alchemist, the German apothecary Henning Brand, in 1669. He concentrated gallons of urine and separated and reduced its components until he had a material that glowed while it combusted. There are two pure forms of phosphorus, white and red. The two forms differ in the way the phosphorus atoms pack together, but both must be kept from the atmosphere by storage in a solvent such as carbon disulfide. The two forms are nearly identical in their chemical reactivities; they are very strong reducing agents that react to form phosphate. The solvent carbon disulfide (S=C=S) gains the
14
ReAction! Chemistry in the Movies
pungent odor of hydrogen sulfide (H2S) as it decomposes to carbonylsulfide (O=C=S) in humid environments: CS2 + H 2O o COS + H 2S
It is possible to make an even stronger alchemical connection. In 1857, Michael Faraday used phosphorus dissolved in carbon disulfide to reduce a yellow-colored gold chloride solution to make a ruby-colored suspension of gold nanoparticles: Au+ + e – o Au0 (Faraday 1857). To achieve this alchemy, Faraday dissolved white phosphorus in “two or three times its bulk” of carbon disulfide, added a gold chloride solution that he “spread about by a glass stirrer, so as to form a flowing layer on the surface,” collected the thin fi lm onto a glass slide, and examined it under a microscope. He determined that it was metallic gold even though it was tinted pink, or rose, or ruby. The solution’s color changed as the gold transformed from an ion to small metal particles. Faraday was so interested in the solubility of his colloidal gold in water that he did not provide an explanation for the change in color except to make the analogy with gold leaf (gold metal pounded into very thin layers), which has veins of red, blue, green, and violet. Today, we know that the different colors result from different-sized clusters of gold metal atoms, so Faraday’s experiment is considered to be the birth of nanoscience. In 2006, chemical historian Ryan Tweney reproduced Faraday’s experiments and found that there was considerable skill involved in each step of the procedure (Tweney 2006). It has been established by other researchers that the gold nanoparticles are colored due to surface plasmon resonance (SPR) (Hu et al. 2006). SPR occurs because the atoms in the gold nanoparticles absorb only certain frequencies of light. The gold nanoparticles absorb and reflect green light, which means that these solutions allow ruby-colored light to pass through. Today we know that Faraday’s gold nanospheres must have been about 50 nm in diameter. When that information is combined with gold’s known molar mass (197.0 g/mol) and density (19.3 g/cm3), it is possible to calculate that each 50-nmdiameter gold nanosphere contains about 4 million atoms of gold metal, n Au (equation 1.1): n Au ⫽
6.022 ⫻1023 atoms mole 19.3g 4 (25 ⫻10⫺7 cm)3 ⫻ ⫻ ⫻ 3 mole 197.0 g cm3
The Chemistry of the “Unknown Impurity” Also from “Dr. Lanyon’s Narrative” (emphasis added): “Have you a graduated glass?” he asked. I rose from my place with something of an effort and gave him what he asked. He thanked me with a smiling nod, measured out a few minims of the red tincture and added one
Dr. Jekyll’s Transformative Formula
15
of the powders. The mixture, which was at first of a reddish hue, began, in proportion as the crystals melted, to brighten in color, to effervesce audibly, and to throw off small fumes of vapor. Suddenly and at the same moment, the ebullition ceased and the compound changed to a dark purple, which faded again more slowly to a watery green. My visitor, who had watched these metamorphoses with a keen eye, smiled, set down the glass upon the table, and then turned and looked upon me with an air of scrutiny.
And from “Henry Jekyll’s Full Statement of the Case,” the fi nal chapter (emphasis added): My provision of the salt, which had never been renewed since the date of the fi rst experiment, began to run low. I sent out for a fresh supply, and mixed the draught; the ebullition followed, and the first change of color, not the second; I drank it and it was without efficiency. You will learn from Poole how I have had London ransacked; it was in vain; and I am now persuaded that my first supply was impure, and that it was that unknown impurity which lent effi cacy to the draught.
When a few minims (i.e., a few drops) of a strong acid, base, oxidant, or reductant are added to a dry powdered compound, it will bubble and give off gases in a reaction. Ebullition is the term for a bubbling solid. The other key to solving Stevenson’s chemical riddle is that very few pharmaceutical compounds change color when they effervesce. Tellingly, he wrote that when the unknown impurity is present, the solution begins red, brightens, becomes dark purple, and then fades to a watery green. In September 1885, one month before Stevenson wrote his famous story, his physician Thomas Bodley Scott treated him with ergotine (Stevenson 1885; Goodwin 2005). Keeping this in mind, the next paragraphs reveal what might be the “unknown impurity.” From age 9 until he died at age 44, Stevenson suffered from a disease that no doctor could correctly diagnose or cure. Since age 29 in 1880, he had the symptom of bleeding lungs and would often cough up blood (Booth and Mehew 1994–1995). This was so acute and recurrent that by 1884 his Edinburgh physician recommended he move to sunnier and drier southern England. This is how Stevenson, his wife Fanny, and her son Lloyd from a previous marriage came to live in Bournemouth from September 1884 to August 1887. It is during these remarkable three years and despite his now chronic coughing that he wrote some of his most memorable works: The Child’s Garden of Verses, Kidnapped, Prince Otto, and Strange Case of Dr. Jekyll and Mr. Hyde. Many contemporaries and biographers have logically concluded that Stevenson had tuberculosis. In 2000, however, Guttmacher and Callahan (2000) speculated that he actually suffered from a rare genetic disease called hereditary hemorrhagic telangiectasia (HHT). This disease is consistent with Stevenson’s respiratory complaints since childhood, recurrent episodes of pulmonary hemorrhage beginning as an adult, and death at 44 from probable cerebral hemorrhage. They note that his mother died at 38 from an apparent stroke, that her father suffered from “a weak
16
ReAction! Chemistry in the Movies
chest” even though he lived a long life, and that several of her siblings died young. The evidence against a diagnosis of tuberculosis is that Stevenson lived 15 years after the onset of the disease, and no one else in his household ever developed it. The evidence against HHT is that he is not reported to have had many nosebleeds, a classic trait of that disease. On the other hand, some descriptions of his bouts of spitting up blood could be interpreted as nosebleeds. His Bournemouth physician Dr. Scott must have decided to give him ergotine because of its ability to constrict blood vessels. It had been known for centuries that when a woman ingested a small amount of ergot fungus during childbirth, she would begin contractions and subsequently suffer little postpartum bleeding. In 1842, the French physician Bonjean was hoping to discover the uterine-contracting compound by soaking ergot in water (Bonjean 1842). While his water extract was less active than the ergot itself, physicians and midwives used it right away as its taste was much less bitter. Bonjean established that the compound in his water extract contained nitrogen, so he named it “ergotine” (ergot + amine). Almost immediately, other chemists and physicians showed that Bonjean’s ergotine was not a pure compound, but none of them were able to separate the mixture into its pure components. Their only separation tools were solvent extraction, crystallization, and distillation. In the following 100 years, better separation methods were developed, and all of the active ergot compounds were identified (figure 1.2). During this same time period, other researchers established that Bonjean’s ergotine slowed the heart, contracted arteries, lowered body temperature, diminished reflexes, and caused short-term madness (i.e., was hallucinogenic)
Dried Ergot
water extract
chloroform extract
Ergotine, 1843
Clavines
nerve stimulation hallucinogenic
Ergotinine, 1875
Ergotamine, 1918
migraine relief vasoconstriction
1948 Ergocorninine Ergocristinine Ergocryptinine none
Figure 1.2. Purification outline for the ergot alkaloids.
ether extract Ergotoxine, 1906
1948 Ergocornine Ergocristine Ergocryptine uterine contractions vasoconstriction nerve stimulation
Dr. Jekyll’s Transformative Formula
17
at high doses. It is obvious that ergot is a “pharmacological toolbox” with many different compounds making up to 2% of its dry mass (Tudzynski et al. 2001). In a letter written only weeks before Stevenson penned Jekyll and Hyde, his wife Fanny described him acting like a mad man after receiving treatment with ergotine to control his bleeding (Stevenson 1885). In 1948, ergotine was shown to be a mixture of water-soluble clavines and ergotamine. Clavines are the precursors of lysergic acid, and the most abundant examples are elymoclavine (figure 1.3), agroclavine, and lysergol. They are probably the major causative agent of convulsive ergotism, a type of ergot poisoning that causes repetitive and painful flexing of all the body’s muscles that is sometimes accompanied by temporary madness (Schardl et al. 2006). A minor component of ergotine is the slightly water-soluble molecule named ergotamine. It was the fi rst ergot compound to be isolated in pure form, a feat achieved in 1918 by Arthur Stoll at Sandoz Pharmaceutical in Basel, Switzerland (Stoll 1920). Ergotamine is widely used for its antimigraine properties even today. Therefore, our chemical sleuthing so far indicates that if Jekyll’s “white crystalline powder” is the mixture called ergotine, then the “unknown impurity” is ergotamine. The color changes provide further evidence.
Cyclized Tripeptide (Alanyl-Prolyl-Phenylalanine) HO
*
CH3 O
* Amide Bond
HN
*
O
OH
N
N H
*
O
O
C
*
*
Indole
N
*
H
CH3
N
*
H
CH3
Lysergic Acid HN
HN
Elymoclavine
Ergotamine
Figure 1.3. Elymoclavine and ergotamine are water-soluble ergot alkaloids that were likely components of Jekyll’s “white crystalline powder.” In these skeleton formulas, the lines represent the bonds between atoms, and carbon atoms are assumed to be at all unlabeled junctions between lines. Asterisks have been added to signify the chiral carbons.
18
ReAction! Chemistry in the Movies
In 1875, Charles Tanret prepared a chloroform extract of ergot that he called ergotinine but that was later shown to be a mixture of ergotaminelike molecules. More important, he also reported a simple colorful test for his ergot alkaloids (Tanret 1875). When he dissolved a few grams of his white crystalline powder in concentrated sulfuric acid, the solution was yellowish orange. The purer the alkaloid, the brighter the orange color. After a few hours, this solution gradually became violet blue. For at least the next twenty years, this chromogenic (color-producing) assay was used to determine the approximate amount of ergotamine-like compounds in ergot extractions. It would certainly have been the assay used by knowledgeable physicians in 1885. Did Dr. Scott show this assay to Stevenson before he treated him? There is no historical evidence on this point, but the description in Stevenson’s novella suggests he did, although Stevenson masterfully embellished the color changes.
Unknown Impurity + Blood-Red Liquor o Hyde Formula No real chemist has ever mixed ergotamine, the “unknown impurity,” with phosphorus dissolved in carbon disulfide, the “blood-red liquor,” so the identity of the Hyde formula is open to speculation. Although phosphorus is a strong reducing agent, it may not have much effect on ergotamine. It may simply react with the acidic solution to create hydrogen gas bubbles, or effervescence. The chemical basis for Tanret’s assay has never been determined, but it seems most likely that the strong sulfuric acid promotes dimerization at carbon 2 of the lysergic acid part (figure 1.4). Atmospheric oxygen that is
O
O
R
O
H H
5'
O
8
CH3
O
H
HN
NH
NH
acid-promoted dimerization
H
H CH3
H3C
N
Ergotamine
N
C
C R
2
CH3
HN
HN
H3C
N
H
oxidation
N N
C
2
HO
O
HN
N
N
CH3 H3C
R C
C
R
O
Reduced Form
Ergotamine 2,2' Dimer
Figure 1.4. Proposed chemical reaction to create the Hyde formula.
O
Oxidized Form
Dr. Jekyll’s Transformative Formula
19
dissolved in the solution is likely to cause subsequent oxidation. Both of these are reasonable assumptions because ergotamine contains an indole group (see figure 1.3). The indole group was fi rst found in indigo dye, which is a dimer of indoles connected with a bond between their carbon 2’s. The name “indigo” means Indian dye and was so named because it was originally cultivated in India by a method that was kept secret for centuries. Indigo has the interesting property that it is clear when reduced and dark blue as it slowly oxidizes by reaction with atmospheric oxygen. It is still used to dye blue jeans, except that the dye is now created synthetically. Ergotamine dimerization in acid would be accompanied by a color change because it would increase the number of conjugated double bonds. The oxidation step may be slow, however. Since neither Jekyll nor Hyde ever waited hours for the color to develop, both of them must have ingested the reduced form. Even so, both the reduced and oxidized forms would be fluorescent and would give off a greenish blue color that Stevenson may have perceived as “watery green” (Stachel et al. 1996).
FROM ALCHEMISTS TO VICTORIAN GENTLEMEN SCIENTISTS According to Rosslynn Haynes (1994), there are six recurrent types of scientists in literature and fi lm: (1) the alchemist, representing the obsessed or maniacal; (2) the stupid virtuoso, representing the out-oftouch with the real world; (3) the unfeeling scientist who has reneged on human relationships and has suppressed human feeling; (4) the heroic adventurer who explores new territories and is a mental giant; (5) the helpless scientist who loses control over his own discovery; and (6) the idealist, who is the unambiguously acceptable stereotype. These literary scientists emerged coincidentally with real scientists. Modern analytical science was set in motion by Isaac Newton’s theories about the nature of gravity and light in the early 1700s. Newton’s special knack was to model the ideal case mathematically and then to determine how closely the real situation resembled the theoretical. Using this method, he repeatedly demonstrated that there were natural laws. The immediate result was that Newton’s mathematical relationships were put to practical use by engineers and caused many scientists to adopt his methods. Between 1830 and 1833, Charles Lyell created our modern understanding of evolution as he published his three-volume geological treatise that was fi lled with evidence. By 1859, the English public was so well read scientifically that the fi rst edition of Charles Darwin’s evidence-based Origin of Species sold out in one day. By the 1870s, professors at the universities of Oxford and Cambridge began teaching courses in science, and the scientific focus moved from the gentleman/scientist engaged in philosophical experiments to the academic scientist engaged in experimental empirical research. This is also when scientists began to
20
ReAction! Chemistry in the Movies
specialize in various branches of science, becoming chemists, physicists, and the like. The fi rst major scientific medical discoveries were also made during the 1800s. The birth of chemical pharmacology began in 1817 when German pharmacist Friedrich Sertürner isolated morphine from opium (Sertürner 1817; Huxtable and Schwarz 2001). First, he soaked the opium in hot water to extract the morphine, and then he added ammonia to precipitate a white crystalline solid. It is also significant he performed tests on dogs and humans to show that the crystalline substance had the same pharmacological properties as the opium but that the nonextracted remainder did not. By 1827, Heinrich Merck of Darmstadt was selling morphine isolated using this method. Morphine was the fi rst natural product, and it set off a wave of similar isolations: strychnine in 1818, caffeine in 1820, cocaine in 1860, and many, many others. Additional medical breakthroughs included vaccination in 1796 by English physician Edward Jenner, the stethoscope in 1816 by French physician Rene Leannec, general anesthesia in 1847 by American dentist William Morton, antisepsis in 1865 by English physician Joseph Lister, the germ theory of disease by German physician Robert Koch in 1877, and the hypodermic needle in 1853 by French physician Charles Pravaz.
PSYCHOLOGY OF OPPOSITIONS There is a transformation scene in the fi nal minutes of I, Monster (1971). The Hyde character Mr. Blake is about to inject his formula. We see his shadow looming darkly in silhouette against the paper-patterned wall of Lanyon’s rooms, one arm greedily outstretched to receive the needle. The fi ngers of Blake’s hand, elongated in the distortion of cast shadow, splay exaggeratedly as the formula begins to take effect. Soon afterward his movements cease and the shadow figure reanimates in calm repose as the Jekyll character, Dr. Marlowe. I, Monster is one of the most psychologically overt movie interpretations of the Jekyll and Hyde story. Dr. Marlowe is an analytical psychologist of the Freudian school. No longer a tale from the Victorian era, this story is set in 1906, fi rmly in the new twentieth century. It is informed by the ideas and methods used in the nascent science of modern psychology, which originated in continental Europe in the late 1800s and spread in influence throughout the world. With it came the promise of new understandings about the human mind and the celebrated names of Freud and Jung. Jungian psychotherapist Barbara Hannah said that Robert Louis Stevenson had “extraordinary insight into the unconscious—far beyond that of his time.” He would have been singularly gifted to practice this new analytical psychology, had its methods been known during his lifetime (Hannah 1971). In devising the bifurcated Jekyll and Hyde character,
Dr. Jekyll’s Transformative Formula
21
Stevenson aptly illustrates the problem of the intimate relationship of the shadow to the self. It is a characterization that anticipates the work of Jung, Freud’s talented student, on the problem of integrating the personal shadow into the larger whole of the self. In Jung’s psychology, the shadow represents the unrecognized, disowned, animal-like personality rejected by the ego. It appears personified in a figure of the same sex as a dark mate who accompanies and clings to its lighter half (Avens 1977). Mythological and cultural counterparts to Hyde–Jekyll are Cain–Abel, Set–Osiris, and Mephisto–Faust. In Jung’s scheme, fi nding the right relationship to the shadow includes acknowledgment and acceptance of it without identifying with it. Failure to do this puts the person at risk of falling under the shadow’s spell, with disastrous results—the very thing that symbolically happens to Dr. Jekyll. While Jung’s psychoanalysis is concerned with the present and how it gives rise to future development in an individual, Freud’s can be characterized as archeological (Salman 1997). Referred to as depth psychology, Freud’s analysis probes into the personal narrative of an individual’s past to uncover repressed authentic desires, often sexual in nature. There is a species level of survival that requires an individual to oppressively conform to societal constraints, causing one to come into conflict on a daily basis with the selfi sh individual need to feel pleasure and avoid pain. Since this confl ict is taking place at the unconscious level, it is revealed in “symbolic disguise” in dreams (Robinson 2007). The dream sequence in Dr. Jekyll and Mr. Hyde (1941) represents Hyde’s driving instinct of self-gratification as one of surreal sexual perversity. He is the whip-happy coachman being pulled by the two women in his life. Let’s return to the wallpaper on which Blake’s shadow is projected in I, Monster. What is the color and texture of the background psychological scene in the Victorian era? Stevenson’s novella was written in 1886, before analytical psychology emerged as an influential practice in Britain. The word “psychology” is Greek for “soul discourse,” and one of the main threads of psychological argument in 1850–1880s Britain concerned itself just with this “psychology of the soul” (Rylance 2000). Proceeding from a view of creation that was pyramidal in scheme, this discourse privileged civilized man with “special dignity.” Although human, women and so-called primitives occupied a lower place in the hierarchy. This creation-view, framed in higher/lower pairings, segregated mind/body functions. They were separate and not equal; each had their place. The “faculty psychology” in currency at the time compartmentalized and categorized human abilities. Higher faculties, such as reasoning and moral ability, resided in the brain. Lower instinctual, animalistic faculties were to be found in the body. This is the cultural framework that formed our Dr. Jekyll. We fi nd him in the 1931 movie version lecturing on man’s “double being,” describing man as “truly two.” Dr. Jekyll’s intellectual and scientific exploration into human nature and how to improve it corresponds to the progressive temper of the age.
22
ReAction! Chemistry in the Movies
Historian Walter Houghton defi nes the Victorian Era as an “age of transition” with the change inherent in this transition possessing a dual aspect, that of destruction and reconstruction (Houghton 1957). The old, rigid, order of doctrines and institutions was being replaced by a new order, one fueled by an expansion of knowledge of all kinds, especially scientific. The optimistic positivism of this new order had its own nagging shadow, however. A pessimistic specter of doubt began to grow out of the sense that beliefs were not on solid footing. And why? The hierarchical scheme of creation had a foil of its own that would come directly out of the Darwinian theory of natural selection through random variations. As Stevenson scholar Julia Reid points out, there was no consensus in the late Victorian era that evolutionary progress was inevitable; unpredictability and extinction were part of natural selection, too (Reid 2006). The question became one of direction: If humans could evolve, could they devolve, as well? The Victorians had a “violent fear of contamination” (Rylance 2000), and the understanding that human nature had prehuman origins was unsettling. Stevenson’s creative genius exploits the “friction and dissention” in the evolutionary thought of the time (Reid 2006). His simian Mr. Hyde is locked in a Darwinian struggle for survival of the fittest with the refi ned gentleman Dr. Jekyll. Who will win?
THE ARCHETYPE MOVIE: DR. JEKYLL AND MR. HYDE (1931) Distribution company: Paramount Pictures Director: Rouben Mamoulian Screenwriters: Percy Heath and Samuel Hoffenstein, adapted from the screenplay for Dr. Jekyll and Mr. Hyde (1920) Short summary: Physician Dr. Henry Jekyll develops a drug to separate the good and evil sides of man and then self-experiments Plot description: Dr. Henry Jekyll (pronounced Jee-kill in this version and played by Fredric March, who shared the Academy Award for Best Actor for his performance) is a brilliant physician who spends his days giving free medical care to the poor and his nights working impetuously in his dungeon-like laboratory. He is engaged to Muriel Carew (Rose Hobart), the daughter of Brigadier General Danvers Carew (Holliwell Hobbes). Carew greatly disapproves of Jekyll’s contempt for the social conventions of the upper classes, so he decides to cool his daughter’s relationship by taking her on a several-months-long trip to Europe. In her absence, Jekyll decides to self-experiment with his formula, which leads him to an encounter with “Champagne Ivy” Pearson (Miriam Hopkins), a singer at the local tavern. Dr. Lanyon (Holmes Herbert) is Jekyll’s best friend and the voice of the scientific establishment that Jekyll ignores but upon whom he relies.
Dr. Jekyll’s Transformative Formula
23
The movie opens from Jekyll’s point of view and continues from this perspective until he begins his lecture at St. Simon’s University. As he leaves his home, he stops to adjust his tie in a mirror and we see him for the fi rst time. After he begins speaking, we see him from the point of view of his audience, and the viewer becomes typically omniscient thereafter. The opening line of his lecture is: “Gentlemen, London is so full of fog that it has penetrated our minds, set boundaries for our vision.” He then says he has analyzed the human psyche and found that “man is not truly one, but truly two.” One is noble and good, while the other is animal-like and bad. He ends his lecture by saying, “I have found that certain chemicals have the power . . . ” which fades into the voices of young men delighted by the beard-pulling lecture and an older physician voicing disgust with Jekyll’s priorities. In the next scene, we see bubbling apparatus that is worthy of a well-funded alchemist (figure 1.5). Jekyll rapidly and carelessly makes
Figure 1.5. Dr. Jekyll (Fredric March) is overshadowed by his chemical apparatus. DR. JEKYLL & MR. HYDE © Turner Entertainment Co. A Warner Bros. Entertainment Company. All Rights Reserved. Photo courtesy of the Academy of Motion Picture Arts and Sciences.
24
ReAction! Chemistry in the Movies
a solution while keeping notes in a book. After he adds a drop of something to the solution in his graduated cylinder, he determines that it works by examining a drop of it under a microscope. Before he brings the formula to his lips, he locks the door and writes a quick note to Muriel to tell her how much he loves her. He stands in front of the mirror as he drinks the clear formula. He clutches his throat and his skin darkens before he falls out of view of the mirror. When he rises, the room swirls around him and scenes from the previous confl icts are replayed. He groans, the image blurs, and he breathes like a monkey. When it is over, he is pleased with his appearance in the mirror and exclaims, “Free, free at last” while he stretches as though he were aching to go somewhere. His future father-in-law objects to Jekyll’s lack of propriety and all the time he spends with patients who can’t pay for medical services. To cool their relationship, Carew takes his daughter on a long trip to Europe. This is the trigger for Jekyll to take the formula for the second time. As Hyde, he begins a relationship with a Music Hall girl named Champagne Ivy to whom Hyde is very cruel. After he kills her, the police almost catch him so he vows never to take the formula again. While on his way to Muriel’s house as a reformed man, he passes through the park and sees a cat stalking a songbird. The off-screen violence causes him to transform spontaneously for the fi rst time. He runs home, but the butler Poole won’t let the stranger Hyde into the house. So, Hyde writes a note to his friend Lanyon asking him to retrieve some items from Jekyll’s lab and to bring them to Lanyon’s house. The bottom of the note indicates to take vials marked A.H.S.T.R.M. [Note from the authors: The meaning of the letters is obscure. If you, dear reader, can solve this mystery, please send us a note.] After observing the transformation and hearing Jekyll’s confession, Lanyon convinces Jekyll he must do the right thing. Jekyll breaks off his engagement with Muriel, becomes agitated, transforms into Hyde, and kills her father. The police chase him to Jekyll’s laboratory, where they shoot him, and the movie ends as he reverts in a series of dissolves from Hyde into Jekyll. Commentary: This is considered to be the best Jekyll and Hyde adaptation and ranks among the best fi lms in this book. It is unusual in many ways. It was produced by Paramount rather than Universal, which produced both Dracula (1931) and Frankenstein (1931) in addition to most of the other classic horror fi lms. It was directed by Rouben Mamoulian, who had directed only a few fi lms before this one but who had been a very successful opera director. It stars Fredric March, who had played only light romantic comedy prior to this fi lm. March shared the Best Actor Award at the fifth annual Academy Awards for his performance. He remained the only actor to win for performance in a horror fi lm until 1982, when Anthony Hopkins won for Silence of the Lambs. A number of the visual puns in Jerry Lewis’s The Nutty Professor (1963) (see chapter 8) can be fully appreciated only by watching the 1931 version.
Dr. Jekyll’s Transformative Formula
25
When MGM decided to produce its 1941 version, it purchased all of the rights to the 1931 version and placed the master copy in its vault. Archivist Raymond Rohauer discovered those prints in 1967 and presented a copy to director Rouben Mamoulian. Thus began its second life in revival theaters and on television. In 1989, MGM/UA Video restored 6 minutes of scenes that had been cut from the movie and released the fi lm on video. You’ll know you are watching the restored version if the movie begins with a subjective camera shot of Jekyll playing an organ. An important note: What strikes one in viewing the Jekyll and Hyde movies is their misogynistic bent. It is difficult to stomach such violence against women and disturbing that this story theme has been gratuitously adapted so many times without scrutiny. An early exception was the 1913 King Baggot silent version, in which Hyde’s violence was solely directed toward boys and men. This version was thought to be “lost” until recently and has now been accepted into the National Film Registry of the U.S. Library of Congress because of its social and historical significance.
THE DRAMATIC JEKYLL AND HYDE ADAPTATIONS Mary Reilly (1996) Distribution company: TriStar Pictures Director: Stephen Frears Screenwriter: Christopher Hampton, adapted from the same-titled 1990 novel by Valerie Martin Short summary: Devoted housemaid Mary Reilly is capable of loving the retiring Dr. Henry Jekyll and the evil Mr. Edward Hyde MPAA rating: R Plot description: Housemaid Mary Reilly (Julia Roberts) spends her long workweeks scrubbing floors and carrying food trays in the home of the physician Dr. Henry Jekyll (John Malkovich). Just as Reilly and Jekyll fi nd they are attracted to one another despite their different social classes, the younger and more vibrant Mr. Edward Hyde (Malkovich again) appears in the home as Jekyll’s assistant. Reilly is initially repelled by Hyde’s aggressive behavior, but her continued presence seems to calm him, and she grows to like him. At one point, Jekyll asks whether she would want to be able to act without consequences, and she says that she doesn’t believe in actions without consequences. In the 14-minute scene beginning at 1:34:30, Hyde explains to Reilly that Jekyll tested two drugs but the cure “took an unexpected form—me.” Hyde was able to come back without injection because he has the stronger personality. Just then, Jekyll emerges after a struggle and sends Reilly to his laboratory for some chemicals. She takes a case from a cabinet to Jekyll/Hyde off-screen and he grabs it. Jekyll emerges and tells butler Poole (George Sheen) that he must go to Finlay’s: “Wait while they
26
ReAction! Chemistry in the Movies
analyze this older sample to fi nd the impurity and then make it for me.” Soon, we see many broken vials on the laboratory floor, and we know this approach was not successful. Commentary: Even though this adaptation is unusual in that Jekyll never uses a mirror to confi rm his transformation, the audience peers behind him and around the corners of his life to see him in a new way. The story concerns the actions and conversations of the house staff in response to Jekyll’s actions and is the only adaptation to mention that the Hyde formula was the action of a minor impurity. It is also the only adaptation in which one good woman is able to love both Jekyll and Hyde. This is the opposite of The Two Faces of Dr. Jekyll (1960), in which one bad woman rebuffs both Jekyll and Hyde. Valerie Martin wrote her novel in response to the nearly complete absence of female characters in Stevenson’s tale to elevate Jekyll’s unnamed maid to the central character. Dr. Black, Mr. Hyde (1976) Distribution company: Dimension Pictures Director: William Crane Screenwriter: Lawrence Woolner, loosely adapted from Stevenson’s novella Short summary: Medical researcher and physician Dr. Henry Pride selfexperiments with a liver regeneration serum that he and Dr. Billie Worth created MPAA rating: R Plot description: Dr. Henry Pride (Bernie Casey, a former Los Angeles Rams running back) is a medical researcher and physician who donates his time to treat patients at the “Watts Free Clinic and Thrift Shop.” In the scene beginning at 11:00, Pride speaks into a tape recorder about his experiments to reverse cirrhosis of the liver. As he does so, he pours a red solution into a blue one and injects the mixture into a brown rat. The rat’s fur turns white due to “reduction of pigmentation,” and it dominates the other rats by driving them into a corner. The next morning, his colleague Dr. Billie Worth (Rosalind Cash) arrives at the lab to fi nd Pride asleep at the bench. They discover that the white rat has killed the other rats. A montage shows Pride’s 1970s-era research lab equipment, including a scintillation counter with teletype printer. He tells his tape recorder, “Tests on animals inconclusive. Need human test. Chemistry’s not right.” When a woman is wheeled in on a gurney by a nurse who says that her liver is shot, Pride tells Worth that this is a great opportunity, but Worth says it is not right to test it. Pride responds that they can’t just let her die in the same way as his mother, so he injects her with the solution. During the night, the elderly lady emerges from her bed white-faced and wild-eyed to kill a nurse. She dies and then reverts back to her original self.
Dr. Jekyll’s Transformative Formula
27
In the scene beginning at 22:00, Pride injects himself at home in his bedroom. This is confusing because he knows that his test on the patient didn’t work. Or, did it work the way he intended? The answer is revealed much later. He looks at himself in the mirror and begins to experience pain. He emerges with a white face, gray-white hair, and white irises. He stares at himself in the mirror before driving his Rolls Royce to the Moonlight Lounge, where he kills four prostitutes and one pimp in the next 40 minutes of screen time. As Pride, he dates Linda Monte (Marie O’Henry), one of his patients, and explains that his mother worked as a cleaning woman in a bordello. He resents prostitutes because his mother died when they wouldn’t help her after she fell ill. Linda is touched by this story and agrees to go to his house. Things turn ugly when he tries to force her to inject the cirrhosis cure. Commentary: Even though Dr. Black, Mr. Hyde (1976) was the third blaxploitation horror fi lm after Blacula (1972) and Blackenstein (1973), it does contain an interesting scientific basis that relates to skin color, which is consistent with the intended audience for the fi lm. Cirrhosis of the liver occurs during the third stage of chronic high alcohol consumption (Berg et al. 2007). The fi rst stage results in a fatty liver, and the second stage in alcoholic hepatitis that is characterized by an inflamed liver and increasing number of dead liver cells. Cirrhosis occurs when dense scar tissue replaces the usual spongy tissue. By the time cirrhosis sets in, the blood is no longer fi ltered very well and certain compounds begin to accumulate. Jaundice is an important symptom of the disease, and it causes the skin and whites of the eyes to become yellow due to the buildup of bilirubin (figure 1.6), a yellow-colored metabolic product from the heme in red blood cells. As cirrhosis continues, the neurotoxic ammonia ion accumulates in the blood, fi rst causing coma and fi nally death.
O COO-
-OOC NH
O NH N H
N H
Bilirubin Figure 1.6. Bilirubin is the yellow-colored compound that builds up in the blood and body tissues of chronic alcoholics.
28
ReAction! Chemistry in the Movies
I, Monster (1971) Production company: Amicus Productions, UK Distribution company: British Lion Films in UK, Cannon Group in USA Director: Stephen Weeks Screenwriter: Milton Subotsky, loosely adapted from Stevenson’s novella Short summary: Psychoanalyst Dr. Charles Marlowe self-experiments with a drug that suppresses a person’s dominant psychological drive Plot description: At his men’s club in November 1906, psychoanalyst Dr. Charles Marlowe (Christopher Lee) discusses the theory that “some men are born evil and some good” with his colleague Dr. Lanyon (Mike Raven) and lawyer friend Frederick Utterson (Peter Cushing). Marlowe believes that everyone has two sides and that they can be separated. In the next scene starting at 6:30, Marlowe pours a test tube of clear solution into a blood-red solution to form a neutral white precipitate. As we learn later, the blood-red solution relieves repression and the clear solution reverses the effect. First, he experiments on his friendly cat, which goes wild after being injected. Then, he experiments on two of his patients, with two very different outcomes. It causes a depressed young woman to become sexually aggressive, and an aggressive male patient to cower like a child in fear of being beaten. Marlowe can’t understand how the drug can work differently on different people, so he self-experiments for the fi rst time on November 4, 1906. In his transformed state, he smiles broadly while a clarinet trills cheerfully, plays with the fi re from his Bunsen burners, and then smiles at his reflection as he brings the mirror from another room into the lab. He teases the guinea pigs and is about to slice one of his mice with a razor when the church bells peal. He uses the reversion formula and is somber again. Back at the club, Marlowe tells Lanyon that Freud believes some diseases are caused by repression of one’s thoughts. Later, he describes Freud’s theory that the psyche consists of the id, ego, and superego. His drug must repress whichever drive is dominant. The id is the unconscious and amoral primitive drive. If it is repressed, the superego becomes dominant and the person becomes a cowering fearful child. Ego is the conscious, logical, and realistic drive. The superego is the moral, critical, and guilt-forming drive. If it is repressed, the person becomes amoral. He continues self-experimenting, and it leads to no good. Commentary: This is the fi rst version to use different solutions for the Hyde transformation and the Jekyll reversion. He becomes the animalistic Hyde when using the blood-red solution and the logical Jekyll when using the clear solution.
Dr. Jekyll’s Transformative Formula
29
Dr. Jekyll and Sister Hyde (1971) Production company: Hammer Productions, UK Director: Roy Ward Baker Screenwriter: Brian Clemens, loosely based on Stevenson’s novella Short summary: Physician Dr. Henry Jekyll discovers that the “elixir of life” is the “female hormone” that transforms him into Mrs. Hyde MPAA rating: R Plot description: Dr. Henry Jekyll (Ralph Bates) is searching for the “antivirus, the universal panacea” that will target all common diseases, such as diphtheria, cholera, and so on. When he tells his friend Mr. Robertson (Gerald Sim) that it takes one or two years to develop protection against each disease, Robertson points out that he will be dead before he is able to fi nish the work. Challenged by this new thought, Jekyll decides to fi nd an “elixir of life,” the secret of eternal youth. Jekyll visits the coroner for a woman’s corpse from which he surgically removes something. In his lab, he slices off some red jelly to add to his solutions and works for ten days without sleep. He tests it on a male fly that lives for three days, which is as long as Jekyll sleeps. Jekyll tells Robertson that the insect lived the equivalent of 200 years, but Robertson points out that the insect is female. Jekyll is mystified that he could have made such an error. Jekyll soon depletes the coroner’s office of suitable female corpses and hires thugs Burke and Hare to procure fresh ones. (Note that Burke and Hare are the names of two real serial murderers who supplied fresh corpses, some of them prostitutes, to Dr. Robert Knox at the medical school in 1820s Edinburgh.) In the scene beginning at 24:45, Jekyll drinks his watery green elixir for the fi rst time. He doubles over in pain and staggers to the living room to sit in front of the mirror, where the reflection is a woman in a robe (Martine Beswick). She fluffs her hair and admires her ring. Neighbor Howard Banner knocks at the door and opens it just in time to see the nude woman fondling one of her breasts. Banner tells his younger sister Susan that Jekyll is entertaining a woman in his apartments. She has a crush on him and decides to pay him a visit. Jekyll answers the door and explains that his widowed sister Mrs. Hyde is visiting. Susan is relieved to hear this and so is Jekyll. He tells Robertson that they had both been correct about the fly experiment. In the meantime, the locals have taken the law into their own hands by hanging Burke and tossing Hare into a quicklime pit (calcium oxide), where he died an agonizing death. Commentary: This is the fi rst and only version in which Jekyll drinks a watery green solution. Jekyll’s original project is slightly confusing in that diphtheria and cholera are both bacterial diseases and are cured by antibiotics, not antiviruses. It is also the fi rst sound version in which Jekyll and Hyde were performed by different actors.
30
ReAction! Chemistry in the Movies
The Two Faces of Dr. Jekyll (1960) Production company: Hammer Productions, UK Director: Terence Fisher Screenwriter: Wolf Mankowitz, loosely based on Stevenson’s novella Short summary: Physician Dr. Henry Jekyll develops a drug to separate man’s two personalities Plot description: Dr. Henry Jekyll (Paul Massie) is a bearded, monotoned bore who transforms into a young, amoral, handsome, clean-shaven Mr. Edward Hyde (Massie again). Jekyll is married to Kitty (Dawn Addams), who is full of life but cheating on him with Paul Allen (Christopher Lee). To complete the triangle, Allen obtains short-term loans from Jekyll for gambling and entertaining Kitty that he never pays back. In the 5-minute scene beginning at 5:00, Dr. Jekyll talks to friend Ernst about his paper in which he argued that man has two personalities, each struggling for dominance. It was rushed to print before he could do the test to prove his theory. Next, Jekyll injects something into a spider monkey and returns it to its cage. After it begins to race around cage, Jekyll says that the drug’s effect will wear off in four days. In a short scene beginning at 14:45, Jekyll opens a small portable cabinet and withdraws a packet containing chemicals that he mixes and uses to fi ll a syringe that he self-injects. He records his pulse but then starts to shake. The scene cuts to a club where Kitty is dancing with Paul; he wants to break up even though Kitty doesn’t. In the lab, the monkey chatters and Henry wakes up with his back to the camera. He writes in his book that the experiment was a success and signs it “Ed.” He steps onto the street to get a cab to the dance hall, where he befriends Paul to experience debauchery. After many evil deeds, Jekyll runs into his lab to look in the mirror for the fi rst time at 1:16:00. The mirror image transforms to Hyde, who tells Jekyll that he must give up the struggle. “They’ll blame you for their deaths and you’ll have to hide as I’ve done.” Commentary: In this version, Hyde is capable of charming a snake charmer. In fact, fi lm studies professor Wheeler Dixon notes that director Fisher made his Jekyll bland and his Hyde suave because he believes in the “charm of evil” (Dixon 1991). When Hyde later takes an interest in Kitty, she does not reciprocate, making her the object of both Jekyll’s and Hyde’s unobtainable desire. Another interesting feature is the use of a mirror toward the end of the movie in which the real Jekyll and his reflection Hyde struggle for dominance with one another. The chemistry of the transformation is poorly explained but the pharmacokinetics is excellent in that the drug’s effect wears off naturally after four days. Dr. Jekyll and Mr. Hyde (1941) Production company: Metro-Goldwyn-Mayer Director: Victor Fleming
Dr. Jekyll’s Transformative Formula
31
Screenwriter: John Lee Mahin, based on the 1931 Paramount screenplay purchased by MGM Short summary: Physician Dr. Henry Jekyll develops a drug to separate man’s two personalities Plot description: During Queen Victoria’s Golden Jubilee week in 1887 London, Dr. Henry Jekyll (Spencer Tracy, one of the top actors of the time) donates his free time to treating mental disorders. He becomes interested in the case of a man named Higgins who heckled the pastor during his sermon about evil. Jekyll’s fiancée is Beatrix Carew (Lana Turner), the daughter of physician Sir Charles (Donald Crisp). At a dinner hosted by the Carews, Jekyll explains that the heckler bumped his head during an explosion and hasn’t been the same since. Jekyll believes he reverted to his animal self and then explains his theories about good and evil being chained together. Sir Charles is shocked and tells Jekyll to give up his theories. He should focus his attention instead on his practice and his social circle. On the way home, Jekyll happens to help barmaid Ivy (Ingrid Bergman) avoid being accosted by two men. She rewards him with her garter. In the 3-minute scene at 29:00, Jekyll is working hard in his lab because he knows Higgins’s health is failing rapidly. A formula in his notebook reads NaHCO2 [it should be a subscript 3]. After he feeds a solution to a gentle rabbit and an aggressive rat, the rabbit becomes ferocious and the rat docile. He is satisfied with the results even though he has missed the opera with Beatrix and Sir Charles. He arrives at the hospital with the formula only to learn that his insane patient had died. At 34:15, in the absence of a suitable human test subject, Jekyll decides to test the formula himself. He adds drops from several clear solutions to a dark solution and drinks it. His pulse is 72 and, at fi rst, he has happy visions of Beatrix and Ivy. Soon, though, he is whipping a carriage horse to make it go faster and the horses become Beatrix and Ivy. At 36:00, he struggles up from the floor in the dark and sees his reflection in the mirror. He laughs at his evil visage. When he fi nds that Beatrix is at the door, he transforms back to Jekyll only to learn that Sir Charles has decided to take her for an extended trip on the continent so that their relationship can cool off. In her absence, Jekyll becomes reclusive, and his butler Poole suggests he visit the music hall for some light entertainment. He reluctantly agrees but is stimulated by the sight of Ivy’s garter on his lab bench and decides to transform to Hyde before heading out. Commentary: This is the most lavish Jekyll and Hyde adaptation, featuring one of the most scientific Jekylls. When he takes the formula that he carefully and calmly developed, he knows it will transform him into an evil Hyde. Dr. Jekyll and Mr. Hyde (1920) Production company: Famous Players-Lasky Corporation (it merged with other companies to become Paramount Pictures in 1925)
32
ReAction! Chemistry in the Movies
Director: John S. Robertson Screenwriter: Clara Beranger, based on the 1887 Thomas Russell Sullivan dramatic play and the 1897 Forepaugh and Fish melodramatic play Short summary: Physician Dr. Henry Jekyll develops a drug to separate man’s two personalities Plot description: Dr. Henry Jekyll (John Barrymore) is a physician who works such long hours in his free clinic that he misses a dinner engagement at the home of his betrothed Millicent Carew (Martha Mansfield). After Jekyll arrives late, her father Sir Danvers Carew (Brandon Hurst) tells Jekyll that he needs experience because “a man has two selves” and that a weak man fears development of the bad self. Jekyll raises his eyebrows in response. Next, Carew brings the men to a tavern and tempts his future son-in-law with a dance hall girl, whom Jekyll rejects, but this causes him to ponder: “Wouldn’t it be wonderful if the two sides of man could be separated?” In the scene beginning at 24:00, Jekyll pours a dark-colored solution out of a test tube into a medium-sized graduated beaker that contains another dark-colored liquid. The potion releases a vapor that looks quite dangerous. He hesitates but then gulps it down. The camera doesn’t cut away while he gags, shakes his head, and convulses until he has an elongated face and an evil smile. The boney-fi ngered ghoul wearing a smoking jacket leaves the lab in search of a mirror. After confi rming his success, he reenters the lab, takes the potion again, and is restored. Commentary: This Jekyll’s theory for separating the two selves originates from his debauched future father-in-law, which gives a creepy feel to this version. Also interesting is the ghostly humanoid spider that enters Jekyll’s bed to transform him into Hyde. These horror aspects are enjoyable, but this version is chemically disappointing because there is so little screen time describing the transformative formula. In 1999, Klepper wrote a critical guide to 646 silent fi lms (Klepper 1999) and said Barrymore would have received an Academy Award nomination for this performance if it had been in existence in 1920. This movie can be viewed online at the Internet Moving Image Archive (www. archive.org). Dr. Jekyll and Mr. Hyde (1912) Production company: Thanhouser Films of New Rochelle, New York Director: Lucius Henderson Screenwriter: Unknown but based on the 1897 play by Luella Forepaugh and George F. Fish Short summary: Physician Dr. Henry Jekyll develops a drug to separate man’s two personalities
Dr. Jekyll’s Transformative Formula
33
Plot description: Dr. Henry Jekyll (James Cruze, who later became an important silent fi lm director) reads a book titled Graham on Drugs that says certain drugs can separate man into two beings, good and bad. The fi rst title card explains that Jekyll is going to test his theory in secret. A blonde Jekyll is wearing a dark lab coat in a room with bottles fi lling the adjacent shelf. He removes some powder from a mortar and uses it to fi ll a beaker about one-third full. He pours a dark liquid from a bottle into a large graduated beaker and shows that he has fi lled a second graduated beaker similarly. After he pours some of the liquid into the beaker containing the solid, it bubbles enough to create foam. He proudly gazes at it, looks to the distance, looks to the ceiling, heartily gulps it down, quivers and grabs his chest as he falls into the seat. He lowers his head to the bottom of the movie frame and his blonde hair dissolves into dark wild hair. He lifts his head to reveal Hyde, who has heavy mascara, false teeth, and fi ngers curled into two claws. He is hunched over as he walks to the mirror, and is then pleased with his reflection. He takes another draft, crouches down in the seat, and dissolves back to Jekyll. The rest of the story is choppy due to missing footage, but it seems that Jekyll is wooing Alice (Florence LaBadie), the parson’s daughter. When Jekyll sees his beloved, he spontaneously transforms to Hyde and attacks Alice. The Parson (Harry Benham) intervenes, so Hyde kills him with his cane. At the end of the story, Jekyll commits poison suicide, a chemical death. Commentary: This Dr. Jekyll follows a protocol from Graham on Drugs to make his formula. Even though that specific book does not exist, there were many similar compendiums used by physicians during this era, such as King’s American Dispensary. Their information was taken from the official U.S. Pharmacopeia (USP) but also included notes about recent research on sources, preparations, uses, and properties. The USP was founded in 1820 and is still the “official public standards-setting authority for all prescription and over-the-counter medicines, dietary supplements, and other healthcare products manufactured and sold in the United States” (U.S. Pharmacopeia 2006). It is a self-sustaining, nongovernmental body whose decisions have the power of law. “Pharmacopoeia” is Greek for “art of the drug compounder.” The fi rst pharmacopoeia was published in Germany in 1542, and its authors relied upon ancient texts for their instructions and therapies. By 1742, the London Pharmacopoeia was the fi rst to include only those preparations that were approved by a committee of men knowledgeable in the art of their preparation. Every official pharmacopoeia since then has further simplified the preparations, adding new compounds as they are identified, and deleting those proven to be ineffective. This is the earliest surviving American Jekyll and Hyde adaptation, although half of it is missing (11 minutes survive out of the original 23). When fi lms broke during projection, they would be quickly spliced
34
ReAction! Chemistry in the Movies
together to continue with the show. Theaters would resell or discard their fi lms at the end of a run because they didn’t have storage space, the fi lms were flammable, and they were considered to be ephemeral commercial product, not art. Nickelodeons thrived between 1905 and 1915 and showed fi lms that people called fl ickers because they had a shimmery look on the silver-painted screen. Movie fi lm speed was not fast enough in these earliest days of moviemaking, and each frame’s exposure was just a little bit different. During this era, the audience had to know the story to appreciate the movie. This version of the Jekyll and Hyde story (Thanhouser Company 1912; Thanhouser Company Film Preservation Inc. 1997) is based on the 1897 Luella Forepaugh and George F. Fish melodrama Dr. Jekyll and Mr. Hyde, or, A Mis-spent Life, which was created for American regional theaters (Rose 1996). It was performed so often that most people were probably aware of its narrative before they saw this fi lm adaptation. In fact, the earliest movie actors came from these regional companies and not from Broadway.
2 Invisibility Steals the Seen Chemistry Creates Criminal Opportunities
PHARMACOKINETICS AND SIDE EFFECTS The Invisible Man (table 2.1) and Jekyll and Hyde (table 1.1) movies share many features. One of the more interesting chemical threads is that the Hyde formula and the invisibility drug accumulate more realistic druglike properties with every adaptation. For instance, in the earliest Jekyll and Hyde fi lms, the effects of the transformative formula are permanent until the antidote is taken or is psychologically triggered. Jump forward to the 1960 version starring Paul Massie, and the injected drug doesn’t cause an immediate transformation and its effect wears off after four days. We can infer that Jekyll spontaneously reverts to Hyde because the compound is metabolized in the bloodstream just like any other drug by a process called pharmacokinetics. Once a drug is in the bloodstream, it must travel to its “site of action,” or receptor, before it can exert its effect. This process is called pharmacodynamics. In the 1960 version of Jekyll and Hyde, for example, the receptors must be skin and hair follicles since the transformation causes the facial hair to disappear, the skin to gain color, and the facial skin to tighten so much that Hyde now smiles. The injected compound is dispersed throughout the circulatory system by the pumping of the heart such that the average concentration of the compound rises to a certain value. When the concentration rises above the pharmacological threshold concentration, enough of it is bound to its targeted receptor that it elicits a physiological response. The invisibility effect wears off after a few hours in The Invisible Woman (1940) and The Invisible Agent (1942). Prior to those versions and in the 2000 version, invisibility was portrayed as being permanent. The two fi lms from the 1940s were produced by Universal Studios, which produced many of the classic horror fi lms of the 1930s, including The Invisible Man (1933). Soon thereafter, when Universal revived their monsters for a B movie titled House of Dracula (1945), the benign physician Dr. Edelman is able to explain each monster’s affl iction as the result of a disease that can be cured. For instance, Dracula’s vampirism is caused by a bat-shaped, blood-borne virus that Edelman proposes to cure with 35
36
ReAction! Chemistry in the Movies
Table 2.1. Chemical invisibility in the movies Title (Year)
Invisibility or Reversion Agent
Hollow Man (2000) Now You See Him, Now You Don’t (1972) Invisible Agent (1942) Invisible Woman (1940) The Invisible Man Returns (1940) The Invisible Man (1933) The Invisible Thief (1909)
Caine-126 Paint [Monocaine] Formula and ray Duocaine Monocaine Unnamed
a vaccine. Either the public understanding of diseases and their cures increased dramatically in the 1940s, or Hollywood was offering science to counter irrational fear during World War II even in its horror fi lms. A distinctive feature of the Invisible Man fi lms is that the invisibility formula produces side effects. Monocaine causes insanity in the archetype The Invisible Man (1933) version and in its sequel The Invisible Man Returns (1940). The physician/chemist in the sequel is the brother of Jack Griffi n, who died in the fi rst movie, so he knows that he must work quickly to fi nd the reversion formula before insanity takes effect. The Invisible Agent (1942) was the archetype’s second dramatic sequel, and its main character explains that the side effects of the drug are not predictable. His side effect is sudden sleepiness, or narcolepsy. Narcolepsy is based on the Greek word narke for numbness plus the word “epilepsy.” This demonstrates the principle that every person has a different set of drug-metabolizing enzymes in their liver. Each enzyme may be very slow, slow, average, fast, or very fast. Since there are a dozen or so major drug-metabolizing enzymes, each with its own speed, it is astonishing that so many of our drugs have the same effect on different people. It could also be that the Invisible Agent’s insanity receptor was different from the archetypal Invisible Man’s and that it didn’t have any affi nity for monocaine. According to a story told by director James Whale (Curtis 1982), after H. G. Wells watched The Invisible Man in 1933, he chided the director for the insanity side effects. Whale made Wells laugh when he responded, “If a man said to you that he was about to make himself invisible, wouldn’t you think he was crazy already?” The most serious type of side effect is one that causes the same harmful pharmacological response in many people. These occur when the drug or its metabolic derivatives binds to a receptor that is not the targeted receptor. When side effects occur at concentrations that are higher than the targeted pharmacological effect (called the toxic dose), it is possible to avoid them by using a lower dose. In real life, this is just one of the bits of information gathered during the clinical trials that test for the effectiveness and safety in any new drug lead. Any side effects that are
Invisibility Steals the Seen
37
revealed during clinical trials have to be reported, and some side effects will cause the pharmaceutical company to abandon the development of the drug they are testing. Another issue for pharmacokinetics is the interaction with food, drink, or other drugs. When The Invisible Woman drinks an alcoholic beverage, she learns that it prolongs her invisible state. By 1942, it was apparently common knowledge that alcohol can overwhelm the liver’s drug-metabolizing enzymes so that it metabolizes some drugs more slowly. At the movie’s end, rubbing alcohol causes her baby to disappear. But, this is another subject—genetics in the movies. In Hollow Man (2000), a research team works in the subbasement of a highly secure military research facility. When the movie begins, we learn that it has been possible to create invisibility and reversion formulas for a variety of animals. We see them successfully test their latest compound called serial irradiated protein Caine-125 to revert a gorilla. In the United States, it is regulated by statute that the safety and effectiveness of new drugs must be preclinically tested in animals before they are clinically tested in humans. Many compounds have the same effects in a variety of animals and humans. In these cases, it is simply a matter of scaling for dose, that is, so many milligrams of a compound for so much mass of the animal, to determine quantity to administer to humans. When compounds have different effects in different animals, it is necessary to test them on a wider variety of animals than usual before testing on humans. In Hollow Man, the researchers were able to prepare the fi rst human invisibility formula called Caine-126, but the human reversion formula was only 97% effective on cultured cells. There are 30 families of drug-metabolizing enzymes in the intestines and liver (Weinshilboum 2003). Seventeen of those families are cytochrome P450 enzymes, which oxidize (meaning that they add oxygen atoms) the compounds to which they bind. The oxidized compounds are further acted upon until they degrade into nutritional compounds or into compounds that can be eliminated through the bladder or bowels. As this occurs, the concentration of the active compound decreases below the threshold, and it no longer exerts its effect. Every human has a different set of drug-metabolizing enzymes, which is the cause of most human variability in drug effectiveness. The complexity is enormous considering that there are 17 different cytochrome P450 enzymes, each acting upon different compounds and each of which may have different levels of activity. To make sense of it, the P450 enzymes are designated by a number for its family, followed by a letter to signify its subfamily, and ending with a polypeptide number. The most abundant members in the human liver are CYP1A2 (CYP for cytochrome), CYP2A6, CYP2B6, CYP2C8, CYP2C9, CYP2C18, CYP2C19, CYP2D6, CYP2E1, CYP3A4, and CYP3A5. From this, you can see that the 2C subfamily contributes the most members. What is not apparent
38
ReAction! Chemistry in the Movies
from this list is that CYP3A4 and CYP3A5 are present in much higher abundance in the liver than the others, and they metabolize more than half of all drugs. Therefore, it is interesting that three-fourths of people with European heritage and half of people with African heritage do not express a functional version of CYP3A5 (Kuehl et al. 2001). This is important because when these people receive cyclosporine to suppress their immune systems prior to skin grafts or organ transplants, the drug is cleared very slowly. An even more dramatic example is CYP2D6, which comprises only 2% of all liver P450 enzymes but is responsible for metabolizing 30% of all drugs. CYP2D6 is also important because it exhibits the greatest variability between humans, with levels of activity that are closely related to ethnicity. The activities range from inactive to normal to ultra-rapid metabolizers. As you can imagine, the 2C family, the 3A family, and the 2D6 enzyme are the subjects of intense scrutiny in drug discovery, development, and clinical trials. Most of the variants (called isotypes) of the P450 enzymes have been identified, and efforts are under way to correlate each drug with its compatible isotypes so that in the near future, patients will be “typed” for their P450 enzymes before they are prescribed a drug (Ingelman-Sundberg et al. 1999). Finally, there are some compounds such as alcohol that inhibit the P450 enzymes. Unless degraded by other routes, drugs persist in the blood. A specific example is cocaine (figure 2.1) as an inhibitor of CYP2D6.
CH3
Ester Bonds
O CH3
N O
OH
O
CH3
CH3
N
Methyl-Ecognine
O
CH3
O
N OH
O
OH CH3
Cocaine
O
O
Ecognine
N OH O
CH3 N
O
Benzoyl-Ecognine
CH3 N
O OH O
Tropa-Cocaine
Pseudo-Tropine
Figure 2.1. Cocaine degrades to these products in the blood or during storage as a dry powder. None of the degradation products are pharmacologically active.
Invisibility Steals the Seen
39
Fortunately, cocaine is not very stable after it dissolves in water or blood. Ester bonds are more labile than most other covalent bonds, and cocaine has two of them. When either ester bond is hydrolyzed, cocaine loses its pharmacological effects of euphoria, enhanced strength, and enhanced intellectual focus.
THE SCIENCE FICTION OF INVISIBILITY Herbert George Wells, author of The Invisible Man (1897), wrote six novels in the 1890s that he called “scientific romances.” In 1934, Wells explained that he always provided the reader with some real scientific facts but then had the scientist bend one of them to make the improbable possible (Wells 1934). He said he achieved this through “an ingenious use of scientific patter.” The bulk of each of his novels speculated on the consequences of making that one thing possible. It is this playful aspect that makes his novels so enjoyable to read even while describing unimagined horrors. This approach was fundamentally different from their most famous predecessors Frankenstein (1815) by Mary Shelley and Strange Case of Dr. Jekyll and Mr. Hyde (1886) by Robert Louis Stevenson. The science of both is so vague that they barely qualify as science fiction and fit more properly within the gothic horror genre, where unexplainable curses abound. Instead, Wells’s novels are much closer to Twenty Thousand Leagues under the Sea by Jules Verne in 1870. Verne incorporated many scientific elements into his novels and provided scientific explanations for incredible features while playing with the facts. In chapters 19 and 20 of The Invisible Man (1897), the character Griffi n explains the theoretical basis for invisibility to his old medical schoolmate Kemp. In the book, Griffi n has a singular name that alludes to the mythological beast protecting the gold treasure of Scythia from the Arimaspians. Rejected as a “monster” by the other medical students because he was an albino, the brilliant Griffi n left medical school to earn a graduate degree in physics. After six months of work on “optical density,” he “found a general principle of pigments and refraction—a formula, a geometrical expression involving four dimensions.” That is, he developed a method that would make things invisible by lowering their refraction index. The only problem was the loss of color in some instances. Griffi n correctly explains to Kemp that we see things because they reflect, refract, and absorb light. He also correctly explains that glass becomes invisible when you place it in water because they have the same refractive indices. He overstates, however, the paucity of colored biomolecules in the human body. Nevertheless, Kemp responds, that he understands because “I was thinking only last night of the sea larvae and all jelly-fi sh!” (figure 2.2). After two and a half more years of graduate research, Griffi n accidentally discovers the unnamed bleach able to remove color from human
40
ReAction! Chemistry in the Movies
Figure 2.2. The transparent zebrafish embryo (Danio rerio) is used in biomedical research because it is possible to monitor development of each cell and tissue. Image courtesy of Shawn Burgess, National Human Genome Research Institute.
tissues without altering their chemistry. “You know the red colouring matter of blood; it can be made white—colourless—and remain with all the functions it has now!” Unfortunately for Griffi n, a lot more has been learned about blood since 1897. It is colored because of the iron ion bound to the heme molecule in hemoglobin, the most abundant protein in red blood cells. In fact, the iron ion gives blood its color. After the atmospheric gases are breathed into the lungs, they diffuse into the bloodstream. The oxygen binds to the iron in the hemoglobin, which becomes red-colored. When the blood reaches the oxygen-hungry tissues, the oxygen dissociates from the iron ion to create the purple-colored deoxyhemoglobin. This is why arteries carry red blood and veins carry purple blood. The hemoglobin iron is normally in its +2 ionic state, capable of binding to oxygen. Bleaches are oxidants and will oxidize the iron to its +3 state, making it incapable of binding oxygen. Knowing this, we might propose that Griffi n’s bleach oxidizes either the heme or protein portion of the hemoglobin so that the oxygen-bound complex no longer absorbs light, but this is not possible in real life. After three more years of hard work and enduring the badgering of his “provincial professor” to publish his work (but apparently not being asked to divulge the nature of his research), Griffi n robs his own father in order to rent a shabby apartment and fi nish his research in solitude. In the confi nes of his apartment, over a period of weeks, he develops a mechanical device to lower the refractive index of the transparent object. It consists of “two radiating centres of a sort of ethereal vibration . . . worked by a cheap gas engine.” When he tests his process on a stray white cat, he learns that he has not solved the problem of reflection because the iridescent tapetum lucidum at the back of its eyes continues to reflect light even though the rest of the cat is transparent. When Griffi n talks of “a general principle of pigments and refraction,—a formula, a geometric expression,” he was way ahead of his time. In 1897, there was a well-known geometric equation to describe refraction, but there was no connection to pigments or chemistry. Refraction is the
Invisibility Steals the Seen
41
angle of incidence, i
angle of refraction, r
imaginary lines normal to the surface Figure 2.3. A light beam refracts as it passes from air to water and back to air. Since water has a refractive index of 1.33, an incident angle of 60.0° yields a refracted angle of 40.6° [sin(refracted angle) = sin(incident angle)/(refractive index)].
term used to describe the disconcerting optical effect of a straight stick appearing to be bent at an angle when placed in water (figure 2.3). In fact, the Latin root for refraction is refringere, which means “to break up.” In 1637, René “I think therefore I am” Descartes (1596–1650) reduced the phenomenon to one value n that we call the refractive index (equation 2.1). His trigonometric derivation was included in his famous Discourse on Method. n = sin i sin r
In this relationship, n is the refractive index of the transparent material, sin i is the geometric sine of the incident angle, and sin r is the geometric sine of the refracted angle. Descartes used the sine function because it increases from 0 to 1 as the angle increases from 0° to 90°. Both the incident and refractive angles are relative to an imaginary line perpendicular to the interface between the two materials (figure 2.3). Descartes developed this equation without concern for its mechanism, only that this effect was produced.His formula has proven to be universally applicable to any solution or liquid. Descartes’s refractive index n was given new meaning in 1678 when Christiaan Huygens published his Traité de la Lumiere [Treatise on Light], in which he argued that light had wavelike properties. Basically, Huygens reasoned that these imaginary light waves travel at different speeds in different materials. When unobstructed in a vacuum, their speed is the
42
ReAction! Chemistry in the Movies
Table 2.2. Refractive indices for various transparent media Material
Chemical Components
Vacuum Air
— mixture of gases
Water Protoplasm
H 2O 60% H2O, 40% biomolecules
Plastic corrective lenses Soda-lime glass Quartz Ruby Zircon Diamond
Na, K, CO3, SiO2 SiO2 Al 2O3 ZrSiO 4 C
Refractive Index 1 exactly 1.000293 (at sea level) 1.330 1.4 1.498 1.510 1.545 1.766 1.920 2.417
Specific Gravity — — 1.00
1.31 2.65 4.1 4.7 3.525
fastest possible. Today, we use the symbol c (meaning constant) to represent the speed of light in a vacuum, and we know that c = 299,792 km/s, one of the universal constants of physics. According to Huygens’s wave theory, the refractive index is a measure of how much slower light travels through a given material relative to a vacuum, which can be represented as n = c/v, where v is the velocity of light in the transparent material. It is easier to measure the refractive index and then use it to calculate how much slower light travels through it. Any comparison of the refractive indices for various materials will show no relation to chemical composition (table 2.2). On the other hand, if we exclude the properties of diamonds from table 2.2, there is a very good relationship between refractive index and specific gravity (similar to density). The simple physical interpretation is that dense materials slow light the most. The chemical basis for the different refractive indices remained unsolved until H. G. Wells’s The Invisible Man of 1897. Unfortunately, Griffi n kept his notes in code, omitted key details, and then had his notebooks stolen by a tramp. It is only recently that real scientists have rediscovered the principle.
THE REALITY OF INVISIBILITY Except for Hollow Man in 2000, chemical invisibility has been replaced by magical invisibility in recent years; two examples are the ring in The Lord of the Rings movie trilogy and the invisibility cloak in the Harry Potter series. In the “Harry Potter” movies, Harry and Ron often use a cloak of invisibility to learn about something they are not supposed to know. Harry Potter’s youthful offenses are similar to those of Dexter Riley’s in Now You See Him, Now You Don’t from 1972. In that fi lm, Medfield College student Dexter Riley creates an invisibility paint based
Invisibility Steals the Seen
43
on the theories of the nineteenth-century Russian chemist Bersakov, who wrote on the topic of preventing light from reflecting. A bust of Bersakov in the movie is the spitting image of Dmitri Mendeleev (1834–1907), the real chemist who published his periodic law of the chemical elements in his 1869 textbook. Riley explains that Bersakov was not able to test his theory because he was locked up after he went insane. Even though the Medfield College dean is angry that Dexter spends his time working on a crazy guy’s theories, Riley’s formula works after a lightning storm electrifies his solution. Once again, invisibility is achieved only after a combination of chemical and physical changes, but this time reflection has been conquered and not refraction or absorption. Dexter and his friend use their invisibility to enter the bad guy’s offices at night so they can thwart his plans to turn the campus into a gambling casino. Magical invisibility is far older than chemical invisibility. The earliest invisibility story was recorded by Plato in 360 B.C.E. as part of book II of The Republic. In this, Socrates is said to have asked for a defense of the opinion that people don’t practice justice for itself, but only for fear of what would befall them if they don’t. Glaucon responds with the story of The Ring of Gyges in which Gyges the Lydian found a ring that made him invisible and then used his invisibility to kill the king, rape the queen, and become the king himself. John R. R. Tolkien based his epic fantasy on the power of the “One Ring.” Tolkien taught Anglo Saxon and English Literature at Oxford University. In 1938 he published The Hobbit and in 1954 and 1955 its three-part sequel The Lord of the Rings. Both stories revolve around the “One Ring,” forged by Lord Sauron to control nine other ring wearers. The ring’s sole power is to make its wearer invisible. While this doesn’t make the wearer a bad person, long-term use of the ring eventually corrupts the user’s morals. Wells was inspired to write The Invisible Man after reading the humorous poem The Perils of Invisibility (1870) written by William S. Gilbert (of Gilbert and Sullivan operetta fame), published in the newspaper Fun. The important lines are: “Old Peter vanished like a shot; But then—his suit of clothes did not.” This amazing image encapsulates within it the basis for visualizing an invisible person. Invisible objects will have volume and mass but no surface unless something defi nes their form. If we expand our view beyond the visible portion of the electromagnetic spectrum, we realize that radar-evading materials such as those used on the Stealth bomber are “invisible” to radiowave detectors. In fact, it has recently become possible to design materials that have a “negative refractive index” so that infrared light “bends” around them. The concept of a negatively refractive material is radical because it implies that the material can reverse the direction of the light wave. For instance, when n is negative, the only solution to the equation n = c/v is a negative value for velocity v given that c is a universal constant with a large, positive value. These negatively refractive materials were shown to be possible on
44
ReAction! Chemistry in the Movies
theoretical grounds by Russian physicist Victor G. Veselago (1968). In 2000, English physicist John B. Pendry used Veselago’s theory to develop such materials (Pendry et al. 1996, 1999). In 2007, an American and French team created nanofabricated material with a two-dimensional negative index in the blue-green visible region (Lezec et al. 2007). In their own words, “the negative response arises from the in-plane isotropic dispersion properties of a transverse magnetic mode involving coupled surface plasmon polaritrons at each metal-dielectric interface.” The theoretical basis for negative refraction derives from Scottish physicist James Clerk Maxwell’s (1831–1879) idea that light has both electric and magnetic properties. In fact, the current mathematical description of the refractive index relates it to the effects electricity and magnetism have on the speed of light through material. The equation is n = √Hµ, where µ is the symbol for electric field effect, called permittivity, and H is the symbol for magnetic field effect, called permeability. This equation indicates that the refractive index, permittivity, and permeability are all positive values. By 1968, Veselago redefi ned it to be n = ±√Hµ and argued that the negative sign should be chosen when both H and µ are negative. He discounted the problem presented by the equation n = c/v by noting that it is not the best way to defi ne light speed. Specifically, he argued that dispersive materials do not conform to that relationship because they have a refractive index that varies with wavelength. For dispersive materials, each wavelength will be diffracted in a different direction, and the total properties of the light beam have to be defi ned based on properties other than wavelength.
LACK OF SELF-REFLECTION LEADS TO CRIMINAL ACTIVITY One rule seems to apply when fashioning a chemical horror movie character: Form follows formula. While the question for Dr. Jekyll and Mr. Hyde may be which form, it is surely what form in the case of the Invisible Man. To tease out an answer, we must return to the mirror. There is a disconcerting undressing scene in The Invisible Man (1933). As he readies himself for bed, Griffi n stands in front of a mirror and begins to unwind a long strip of bandage. He has been using this gauze to define the shape of his head and neck in order to talk to Kemp without scaring him. As the downward spiraling motion of his arm causes the bandage to fall away we see, well, nothing. His head is gone from view, with only the collar of his pajama top guiding us to where his neck would be (scene at 44:00). This scene plays on a reversal of rather standard bedtime bedroom movie fare. In this elegant room, with a candelabra reflected in the mirror over the dressing table, couldn’t we almost expect to see a beautiful woman brushing out her hair before retiring for the night? What we see instead is a man who is symbolically mentally wounded (remember that gauze) admiring a self who isn’t visually there. The scene gives us an important
Invisibility Steals the Seen
45
insight into this chemist; he has found a way to make his form imperceptible, but it has been a vain pursuit. Most people know the myth of Narcissus, the boy who looks into a pool and becomes so enamored of his reflection that he remains frozen in that position forever as punishment for his vanity. Though something of a minor myth in the time of the Greeks, its theme of unhealthy selfabsorption speaks to our individualistic era to the point that some call ours the “Age of Narcissism.” In his seminal early paper On Narcissism, Freud connected self-regard with narcissistic libido, a psychic or emotional energy (Freud 1957). In later psychoanalytic work, a pathological narcissist would be defi ned as someone who exhibits a lack of self-esteem coupled with a compensatory grandiosity. In other words, something critical has gone missing: a cohesive self with growth and development potential. The Invisible Man is analogous to the pathological narcissist because he is lacking by defi nition; he has been negated from view. In the updated movie version of the story, the title Hollow Man curiously brings us closer to an image we can see in our mind, that of an empty shell (figure 2.4). This is the papier-mache figure plotting world domination with an army of chemical changelings like himself. The Invisible Man’s looking glass merits a closer look. The mirror is a potent symbol that can illuminate understandings about the structure
Figure 2.4. The transparent Dr. Sebastian Caine (Kevin Bacon) sees his shape. Hollow Man © 2000 Global Entertainment Productions GmbH & Co. Movie KG. All Rights Reserved. Courtesy of Columbia Pictures. Photo courtesy of the Mary Riepma Ross Media Arts Center.
46
ReAction! Chemistry in the Movies
of the self and our mental functioning. It is deeply connected to human social organization. The terms “reflection” and “reflexive” are often associated with self-concepts and imply the ability to refer back to oneself. In self-reflection, for instance, some trait, behavior, or action we hadn’t seen clearly before can be “faced” and taken in for evaluation. There is also the connotation that any changes resulting from this process will be for the better. Self-reflection is a positive activity (see the discussion of shadow in chapter 1, page 20). In earliest childhood, the interaction between a mother, or caregiver, and child is characterized as a mirroring experience. Healthy self-esteem develops through the process of an infant internalizing the admiration it excites in its mother’s eyes (Knight 2006). Our ability to learn by imitation, so critical to human culture, has a biological basis. Neurologists have identified brain cells called mirror neurons that act in systematic involvement with several areas of the brain to “mirror another creature or another person’s activities” (Solomon 2006). These neurons are basic to the functions of understanding and imitation, and they may be of profound importance to moral decision making, those circumstances when one has to negotiate the antagonisms of the ego’s needs and social obligations (Tancredi 2005). We cannot, however, touch on the topic of morality without fi rst defi ning empathy, which for our present purposes is a “vicarious affective response to another person” (Hoffman 2000). Mimicry, also called the imitation/feedback loop, is a neurologically based empathy that is likely hard-wired. Studies have shown it to be an involuntary, fast mechanism that assures a match in emotional, nonverbal expression in face-to-face encounters. Mimicry-based empathy promotes the “prosocial motives” of solidarity and involvement with others and may even be a “prosocial act” (Hoffman 2000). Sociopaths and psychopaths are noted for their lack of empathy, compassion, fear, and remorse. Recent research on psychopathic patients shows that these individuals’ brains appear to be misfiring, leaving them unable to generate the emotions necessary to constrain violent, impulsive actions. Imaging techniques reveal they have little or no activity in the amygdala area, believed to be the part of the brain where feelings of empathy are generated. Further studies point to a concomitant deficit in the reasoning center of the psychopathic brain that mirrors a dysfunctional amygdala. Such individuals are unable to rein in destructive, often murderous, instincts, and their moral competence is impaired. Neuroscience research might change the course of this seemingly deterministic biology; one of its hopeful notes is that psychopathy may respond favorably to treatment in the future (Tancredi 2005; Brandt 2007). Why does the invisibility theme keep returning in the movies, and what does it give us on reflection? When the Invisible Man in the 1933 version proudly asserts that he can use his invisibility to “rob and rape and kill,” we hear a cold echo from the Ring of Gyges story in part II of Plato’s Republic. We should ask ourselves how we would respond to the moral
Invisibility Steals the Seen
47
challenge if, like Gyges, we could act in ways that evaded both detection and punishment. One infamous psychological experiment from the early 1970s offers us some uncomfortable data. The experiment, conducted at Yale University by Stanley Milgram, was conceived as a simple investigation into obedience and authority. One participant, a “teacher,” administered electric shocks that increased in intensity when another participant, playing the role of “learner,” gave wrong answers to questions. The learner was an actor behind a partition who received no actual shocks at all, but made escalating vocalizations of pain that could be heard by the teacher. Present in the room with the teacher is the “experimenter,” the authority figure who urges the teacher to keep administering the shocks. The aim of the investigation was to discover conditions under which people would defy authority when faced with a “clear moral imperative” (Milgram 1974). The “surprising and dismaying” results of the experiment showed that a large number of participants continued to administer the shocks up to the highest level. The obedient teacher subjects didn’t see themselves as being responsible for their own actions. Instead, they accepted the collateral invisibility that was conferred by the agent of authority. As Milgram said, “the disappearance of a sense of responsibility is the most far-reaching consequence of submission to authority” (Milgram 1974).
MONOCAINE AND LOCAL ANESTHETICS Monocaine is one of the best-known imaginary compounds in the movies since it is the active ingredient in the mixture prepared by The Invisible Man (1933), the archetype fi lm for this chapter. Most viewers don’t realize, however, that the -caine suffi x signifies that it is a local anesthetic, such as Novocain (procaine), lidocaine, and, of course, cocaine. Consider the effect on your body when one of your nerves has been blocked through the injection of a local anesthetic. You can’t feel the surrounding tissue. It’s as if it isn’t there. Carrying this even further, if cocaine is a local anesthetic, Griffi n’s insanity is a metaphor for cocaine addiction. In the 1933 fi lm, Dr. Jack Griffi n is an assistant food preservation chemist. In his spare time over a period of five years and with the full knowledge of his employer (who is also his future father-in-law), he works in a locked room in his employer’s basement on a secret project. After it is too late, his employer Dr. Cranley and his coworker Dr. Kemp discover that Griffi n extracted the compound monocaine from “a rare Indian flower that draws color from everything it touches.” Cranley worries that Griffi n may not know of a research project published in an obscure German journal. When a dog was injected with monocaine, it turned “dead white” and then went mad. Jack Griffi n was obviously an excellent experimentalist because he succeeded in isolating the compound, but it was his lack of scholarship that got him into trouble. In his subsequent madness, he knocks over a baby carriage, strangles a policeman to death,
48
ReAction! Chemistry in the Movies
causes a train wreck that kills an entire load of people, and then pushes Kemp and his car over a cliff. At the end of the ensuing manhunt, the police shoot Griffi n, and he dies while hemorrhaging transparent blood in a hospital bed. Monocaine, the invisibility compound, was an invention of screenwriter Robert Cedric Sherriff for the movie The Invisible Man (1933), directed by James Whale. Even though Sherriff wrote an autobiography (Sherriff 1968) and Whale was the subject of two detailed biographies (Curtis 1982; Gatiss 1995), neither person ever revealed the genius behind the choice of the name monocaine. Nevertheless, both men were very creative, and The Invisible Man was the third of the five fi lms they made together. They helped each other become established in the London theater circuit. Sherriff had written Journey’s End (1928), a play about life in the trenches during the Great War that consisted of three men in a bunker in the middle of the stage talking about life and the war. Everyone who read it thought it was brilliant, but no one wanted to produce a play for a war that was now part of history, which contained little action, and which had no female role. James Whale had been an actor and stage designer who wanted to direct, so he accepted. The fi rst few performances were to small appreciative crowds, but after it moved to a theater in London’s West End, it was a smash hit that played for two years to capacity crowds. They brought the play to New York, and then Whale directed the screen version in Hollywood. Whale stayed to direct Frankenstein (1931) and The Old Dark House (1932) for Universal Studios. Universal was so pleased with the box office receipts from those two fi lms they asked him to direct The Invisible Man. Even though more than a dozen screenwriters had written treatments of the novel, none of them was satisfactory. Whale convinced Sherriff to travel to Hollywood to write the screenplay. Sherriff mentions he was embarrassed that the studio executives thought his adaptation was brilliant since he felt he hardly strayed from the novel. The most chemically divergent aspect of the screenplay is the presence of monocaine and its insanity side effect. According to Whale, they both decided they didn’t want Griffin to be as rationally evil as he was in the book; otherwise, no one would want to identify with him when he committed his crimes. They decided to gain sympathy by having the invisibility drug affect his mind. Duocaine was the name of the fi rst human invisibility reversion compound according to the 1940 sequel The Invisible Man Returns. According to this naming scheme, the next compounds in the series would be tricaine invisibility, tetracaine reversion, pentacaine invisibility, and so on. Fast forward to Hollow Man (2000), in which Dr. Sebastian Caine develops Caine-125, the gorilla reversion formula, shortly after the movie begins. His colleague chides him for naming the compounds after himself and says they might work better if he didn’t. Later, Caine convinces his team to inject him with Caine-126 but doesn’t tell them that he was denied permission for this. In a surprising lack of judgment for which
Invisibility Steals the Seen
49
the team should rightly be reprimanded, they inject him with it before they’ve developed the reversion formula. Subsequently, his colleague spends the next three days working to develop Caine-127, but it is not completely effective on cells in culture. Caine, who has been confi ned for this time period due to his invisibility, becomes angry and decides to wreak havoc. The story of cocaine as the fi rst local anesthetic is a remarkable one that bears retelling. Cocaine is familiar to us today as an illegal recreational drug that causes euphoria when snorted, injected, or ingested. Much less generally known is how cocaine revolutionized surgery because of its local anesthetic effects when topically applied. Cocaine (coca + amine) is the amine compound the coca plant produces to kill insects that try to feed on it (Nathanson et al. 1993). The coca plant, Erythroxylum coca, grows wild on the eastern slopes of the Andes Mountains. Peruvians and Bolivians have historically chewed its leaves for stimulation, to reduce fatigue, to stave off hunger, and to protect against the tiring effects of high elevation. They preferred to chew the sweet-tasting coca plant leaves, which we now know contain less of the molecular compound cocaine and more of its precursors and derivatives. In contrast, Europeans would later focus their attention on the bitter-tasting coca leaves, which are particularly rich in the alkaloid we call cocaine. Since chewing coca leaves or drinking coca extracts is much less addictive than imbibing purified cocaine, European chemists unwittingly created a social problem when they developed the methods for its extraction in pure form. Amerigo Vespucci’s published account of his world circumnavigation in 1499–1500 describes how the South Americans chewed coca leaves (Van Dyke and Byck 1982). The stimulating effects of chewing coca became widely known, however, only after the conquest of Peru by Francisco Pizarro in 1530. For the centuries that followed, visitors would bring back samples of the coca leaves only to fi nd that they had degraded during transport. Interest was greatly revived in 1859, when Dr. Paolo Mantegazza published a prize-winning essay about the use of coca by the Peruvians and of his own self-experiments with chewing leaves and drinking a tea made from its leaves (Mantegazza 1859). He had just returned to Milan from Peru, where he had been practicing medicine. Mantegazza self-tested varying doses and then measured his pulse rate and level of intoxication; in other words, Mantegazza engaged in the fi rst psychomotor pharmacological experiments ever undertaken. He concluded his paper with the following: “I prefer a life of ten years with coca to one of a hundred thousand without it.” In about 1863, the Corsican-born Parisian apothecary Angelo Mariani was inspired by Mantegazza’s essay and began cultivating coca plants in his greenhouse in Neuilly on the Seine, just outside Paris (Mortimer 1901). First, he learned how to stimulate the plants to produce the highest amounts of the bitter alkaloid. Then, he masked the bitter taste of the coca extract by adding it to red wine from Burgundy to create a marriage
50
ReAction! Chemistry in the Movies
made in heaven that he named Vin Mariani. He also marketed Mariani elixirs (coca dissolved in alcohol), Mariani lozenges, and Mariani teas. In 1872, he published his fi rst paper on the topic of coca’s history and restorative powers; it went through at least four editions (Mariani 1886). As the years passed and he uncovered more of its cultural and scientific history, the paper grew to book length. The third edition of Coca and Its Therapeutic Application was translated for the American market in 1890 and went through at least four editions itself (Mariani 1890). In 1917, the American Medical Association published the cocaine content of a variety of patent medicines (Street 1917) and reported that Vin Mariani contained 6–8 milligrams per ounce, which is enough for a person to feel its effects. The widespread use of Vin Mariani among actors and music hall singers can be traced to Parisian laryngologist Dr. Charles Fauvel, who prescribed it for vocal fatigue in one of his patients. Vin Mariani worked so well that Fauvel wrote to a French medical magazine that it anesthetized the pharyngeal mucous membrane while also invigorating the muscles (Mariani 1902). As the prescriptions began to roll in, Mariani used his cunning promotional skills to develop an even wider market, the largest of which was as an export to the United States. His most famous advertising gimmick was to send a case of Vin Mariani to prominent artists, royalty, scientists, heads of state, authors, religious leaders, and so forth, and then ask them for an endorsement and a photo. Their responses would be sent as advertisements to newspapers. The endorsements were also combined as plates in a series of more than 14 volumes between 1894 and 1925, each with between 60 and 90 endorsements (Mariani 1894–1910). Among the hundreds who endorsed Vin Mariani were Thomas Edison (1896, vol. 2); Pope Leo XIII, who gave him a golden medal in appreciation (1899, vol. 4); U.S. President William McKinley (1899, vol. 4); Henri Poincaré (1910, vol. 12); Jules Verne (1896, vol. 2); and H. G. Wells (1910, vol. 12). Mathematician Poincaré wrote an equation beneath his photo: 20 Mariani = 100 T. Author Wells sent “before” and “after” cartoons of himself that were published instead of his photo (slouching and unhappy vs. erect and beaming with energy and happiness). The story of cocaine is different from that of coca in that addiction now enters the equation. In 1858, the Austrian government sent the frigate Novarra around the world to study trade opportunities. The trip was led by Karl von Scherzer, who returned with many items, including a quantity of coca leaves (von Scherzer 1861). These he gave to the German chemist Dr. Friedrich Wöhler at the University of Göttingen, who in turn gave them to his graduate students Albert Niemann and Wilhelm Lossen. Niemann developed a reproducible procedure to isolate nearly pure cocaine from the leaves (Niemann 1860). Today we know that cocaine is found at anywhere between 0.7% and 1.8% of the dry weight of the coca leaves. Niemann’s analysis showed that its chemical formula was C32 H2 O Az O2 (today written as C32H2NO3). Even though
Invisibility Steals the Seen
51
his extraction procedure was the best of its time, his elemental analysis was incorrect. Three years later in 1865, Wilhelm Lossen reported that the chemical formula for cocaine is C17H21NO4, and he got it right (Lossen 1865). In his 1862 thesis, Niemann noted that cocaine had a bitter taste and numbed his tongue, an observation that many scientists and nonscientists had reported when chewing coca leaves (modern chemists no longer report the taste of their compounds). Even though Niemann died within the year of mysterious circumstances, his isolation procedure was used by the developing medicinal chemistry companies in Europe and the United States to isolate the alkaloid and make it available to physicians for use in their practice. In 1884, a recently graduated, unmarried, and pre-psychoanalytical Freud tried to cure a colleague of his morphine addiction with some cocaine that he had purchased from Merck Chemical Company. The colleague had become addicted to morphine after he mangled his thumb and suffered great pain because of it. His colleague was cured of morphinism but soon became a cocaine addict and died two years later from his drug abuse. In the meantime, Freud began self-experimenting with cocaine. He monitored his pulse, breathing rate, strength at pushing weights, and general feeling before and after he ingested various amounts of cocaine. Freud described these things in his fi rst paper on the subject “Über Coca” (Freud 1884) that brims with such enthusiasm it is easy to overlook the science. Freud had two of his friends from the University of Vienna help him with these experiments. One of them was Dr. Carl Koller, a newly minted eye doctor working at the university’s teaching hospital, the largest in Vienna. Despite numerous medical reports that coca, and subsequently cocaine, numbs the tongue, Carl Koller was the fi rst physician prepared to use this information in surgical practice. While attending the University of Vienna medical school, Koller had learned that general anesthesic agents such as ether or chloroform could not be used during eye surgery because the patient had to be awake to cooperate during the operation. After graduation, he decided to fi nd a suitable anesthetic for use during this type of surgery. Although he had also learned in his studies that cocaine numbs the tongue, it did not occur to him it might be the anesthetic he was looking for until he self-tested it on Freud’s request. Shortly afterward, Koller raced to the lab and found that a drop of a very dilute cocaine solution was enough to anesthetize the eye of a frog. He could touch the frog’s cornea and it didn’t fl inch. After also testing a rabbit and dog, Koller selfexperimented and found that he could touch his own cornea with the head of a pin without feeling a thing (Koller 1884a). On September 15, 1884, Koller’s friend Dr. Josef Brettauer read his paper for him at the annual meeting of the Deutsche Ophthalmologische Gesellschaft and performed a demonstration of cocaine’s local anesthetic effects on the eye (Knapp 1884; Koller 1928). The audience’s response was electric. American surgeon Henry Noyes heard the lecture and sent
52
ReAction! Chemistry in the Movies
a summary of it to the Medical Record (USA) (Noyes 1884). Within days, doctors across Europe and the United States began to use this new surgical advance on all the mucosal tissues. Soon, Koller himself presented his fi ndings at a Viennese Medical Conference (Koller 1884b). This lecture was translated and appeared in medical journals such as the Lancet (England) (Koller 1928) and Semaine medical (France), as well as in many newspapers. By November 1884, the medicinal chemical companies had nearly doubled the price of cocaine as their supplies were outstripped by demand. Koller’s daughter later wrote that among the many letters of congratulations to Koller for his discovery is a note from Freud that is addressed ‘Seinem lieben Freunde Coca Koller’ [To my dear friend Coca Koller] (Becker 1963). Among natural products, cocaine is an alkaloid, which means that it has at least one nitrogen atom that confers alkaline, or basic, properties to it. Most alkaloids are extracted from plants. It is important to know the acidic and basic properties of a compound to determine how readily it will pass through cell membranes, which is a predictor of how fast it will enter the blood stream. Cocaine is produced by only two plants, both in the genus Erythroxylum, a shrub native to the South American Andes. Erythroxylum coca is found mostly in Bolivia, whereas Erythroxylum truxillense is found primarily in Peru. The Erythroxylum genus has very few relatives. It is a dicotyledon (a type of flowering plant) of the subclass Rosidae, order Linales, and family Erythroxylaceae. The order Linales has only one other member, the family Linaceae, the rather large flax family, none of which produce these same alkaloids. In fact, the only other family of plants to produce tropane alkaloids (figure 2.5; tropanes are any compound with the bicyclic ring structure that is common to cocaine and atropine) is Solanaceae, a very distant relative with members such as Atropa belladonna (deadly nightshade), Datura stramonium (thornapple), Hyoscyamus niger (henbane), and Mandragora officinarum (mandrake). Many members of Solanaceae produce pharmacologically important tropanes such as atropine (racemic hyoscyamine), scopolamine (also called hyoscine), and hygrine. Its
CH3
CH3
O
N
N
CH3
CH3
N
O O
O
HOH2C
O O
Tropane
Cocaine
Figure 2.5. Three tropane compounds.
Atropine
Invisibility Steals the Seen
53
ancestry merges with Erythroxylaceae only at the dicot stage, which is very ancestral indeed. The lineage of Solanaceae is subclass Asteridea and order Solanales. Recent determination of the biosynthetic pathways for these two flower families shows that they converged on the tropanes independently. The chemical evidence is that the 3-hydroxyl is in the alpha position rather than beta in atropine and scopolamine. This single change requires the activity of a completely different enzyme from the one in Erythroxylum coca. The handful of responsible drug manufacturers (in the United States: Squibb, Eli Lilly, Upjohn, Abbott, and Parke-Davis) that sold directly to physicians and pharmacists added cocaine to their stable of products. The medical era of cocaine use had begun, but it would soon end from its overuse by the self-medicating public. Shortly after Koller’s great discovery was publicized, many patent medicine manufacturers entered the cocaine business (Spillane 2000). They earned robust sales until their products were regulated by the 1906 Pure Food and Drug Act and then eliminated by the 1914 Harrison Narcotics Tax Act. Cocaine was included among the narcotics even though it is a stimulant. Patent medicine manufacturers advertised to drug wholesalers, pharmacists, and to the general public. Before 1906, there was no federal oversight of any part of this trade. There were no laws to determine whether the claims for any products were true, or if the product was effective. If you had the money, you could buy it. Coca wines made up the lion’s share of the coca/cocaine products and Vin Mariani was the biggest seller in the United States probably because it had been established even before the 1884 reports of its anesthetic effects. Among the alternative coca wines was the one produced by the Atlanta, Georgia, pharmacist John Pemberton called “Peruvian Wine Coca.” After Atlanta voted itself “dry” in 1886, Pemberton replaced the wine with cola extract (which added caffeine to the drink) and called it Coca-Cola. Coca extract was in the drink at a level of about 1 mg cocaine per fluid ounce until 1907 (Street 1917; Pendergrast 1993), one year after the Pure Food and Drug Act began hurting its sales. This amount of cocaine was significantly less than in Vin Mariani but was probably enough to satisfy and refresh. Muckraking journalists and Nathan Davis of the Department of Agriculture were well aware of the social problems created by this legal cocaine use, and their efforts were primarily responsible for creating and then enforcing the 1906 Pure Food and Drug Act. One of the most memorable movies to use cocaine addiction as a theme is D. W. Griffith’s For His Son (1912). The story features a pharmacist who creates a product called Dopokoke that contains pure cocaine, not just coca extract. The product sells like crazy but also brings tragedy to him when his son becomes addicted and dies. Between 1884 and 1892, medicinal use of cocaine for local anesthesia was linked to at least 18 deaths and 200 cases of cocaine intoxication (Armstrong Davison 1965; Anonymous 1979). In fact, Koller was the
54
ReAction! Chemistry in the Movies
fi rst to report that cocaine was toxic (Koller 1884a, 1884b). During this same time period, cases of addiction began to be reported. The most famous of these was Freud’s own admission (Liljestrand 1971). Although nonaddictive, local anesthetics were available from the 1890s on, and the incidence of intoxication or death from local anesthesia remained a concern. In 1924, the American Medical Association formed a council to examine the issue (Mayer 1924). They studied 43 fatalities in detail and concluded that 40 of them were due to the action of local anesthetic drugs. In half the cases, the local anesthetic was cocaine. The study concluded that cocaine should not be used as an injectable drug, and it was used only topically in clinical practice afterward. Since the 1920s, the clinical use of cocaine has been reduced to only a few applications. For instance, it is still used as a topical anesthetic during ear-nose-throat surgery because it constricts the blood vessels, acting simultaneously as a local anesthetic. Chemists searched for other compounds that had local anesthetic properties but that weren’t toxic, addictive, or euphoric (Liljestrand 1971; Brain and Coward 1989; Ruetsch et al. 2001; Calatayud and Gonzalez 2003). The fi rst substitute was a natural compound marketed as Eucaine (meaning “true” caine) by Schering and Glatz in Germany in 1896 (figure 2.6). Although Eucaine wasn’t addictive, it also wasn’t as effective as cocaine but just as poisonous. The search continued. The structure of the tropine part of cocaine and atropine was fi nally determined in 1898 by Richard Willstätter in Munich, Germany
HN
CH3
CH3
O
CH3
CH3
N
N
O
H3C
O
O
O
H3C
HOH2C
O
O O
Cocaine
Eucaine
Atropine (and Mandragorine) O
H N
H3C
O CH3
O H3C H3C
NH2
Butethamine (Monocaine®)
N
CH3 H3C
O
H N N
H3C
Procaine (Novocain®)
NH2
Lidocaine
O H3C
Figure 2.6. A variety of local anesthetics and atropine. The top three compounds are natural, the bottom three synthetic.
Invisibility Steals the Seen
55
(Willstätter 1898). The tropine part has the nitrogen and double rings (see figure 2.5). It was now possible to consider the parent compound when making synthetic derivatives. From its structure, we can see that the tropine part is very rigid. It holds the positions of the benzoyl and methylamine parts at a certain inflexible distance from one another. In 1904, Alfred Einhorn synthesized and patented Novocain (meaning “new” caine) for Hoechst Chemical Co. near Frankfurt, Germany (figure 2.5) (Einhorn and Uhlfelder 1909). Its flexible structure was very different from cocaine’s, but it did have an ester and two ethylamines. Its low toxicity and acceptable effectiveness when used in combination with adrenaline were reported in 1905 (Braun 1905). In the United States, it was patented as procaine. Hoechst marketed it in the medical press as being less toxic than cocaine and non-habit-forming. It remained the top dental anesthetic for more than five decades (Hadda 1964). The main problem with Novocain was that it had to be used in high doses, which meant that some patients and health professionals became allergic to it. Einhorn synthesized quite a few compounds that were variations of the Novocain structure, but none offered substantial improvements on its properties. In 1923, after 25 years of work, Willstätter in Munich and the chemists at Merck Laboratories in Darmstadt, Germany, copublished a method to synthesize cocaine in the laboratory (Willstätter et al. 1923). This synthetic version had the same properties as the natural compound, providing the most important proof that its chemical structure was correct. In 1937, real chemists proved they had watched The Invisible Man when they named their latest local anesthetic monocaine (see advertisement in chapter 10). Its structure is very similar to Novocain. The most recent advance occurred in 1943, when Nils Löfgren and Bengt Lundquist developed xylocaine (marketed as lidocaine). It required a much lower dose than Novocain, making it less likely to induce an allergic response (Löfgren 1948). Lidocaine has an amide linkage rather than an ester linkage, causing it to degrade more slowly. Another important feature is that it also has a slightly less flexible structure than Novocain. As a result, lidocaine is four times as potent as Novocain, has rapid onset, and is effective when injected. It quickly displaced Novocain from the market and is still widely used, whereas the original Novocain is rarely, if ever, used. All subsequently marketed local anesthetics have the amide linkage and not the ester linkage.
THE ARCHETYPE MOVIE: THE INVISIBLE MAN (1933) Production company: Universal Pictures Director: James Whale Screenwriter: R. C. Sherriff, based on the same-titled 1897 novel by H. G. Wells
56
ReAction! Chemistry in the Movies
Short summary: Assistant food preservation chemist Jack Griffi n takes monocaine invisibility formula, can’t reverse invisibility, goes insane, and kills people Plot description: Drs. Jack Griffi n (Claude Rains) and Arthur Kemp (William Harrigan) are assistant chemists working for Dr. Cranley (Henry Travers), whose food preservation laboratory is located in his home. Griffi n is engaged to Cranley’s daughter Flora (Gloria Stuart), although Kemp loves her, too, but no one knows it. Cranley trusts Griffi n so much that he has allowed him to work for five years in his spare time in a locked room in his basement on an unknown project. After Griffi n disappears, Cranley and Kemp search that laboratory for clues and discover a list of compounds, the last of which is monocaine (brief scene beginning at 26:30). Cranley explains that monocaine is derived from an Indian flower that draws color from everything it touches. He worries about Griffi n because he may not know about the German monocaine experiments with a dog that turned white and then went crazy. The audience already knows that Cranley’s fears have been realized. In the 7.0-minute scene starting at 16:45, the invisible man is at the Lion’s Head Inn working with his chemical apparatus. He assaults the innkeeper Mr. Hall, takes off his bandages to reveal nothing underneath, laughs maniacally, and assaults a policeman and a few citizens before heading out onto the street. We learned earlier that he chose to work on invisibility because it will bring him wealth, fame, and power. He had taken a few injections every day for one month and, after becoming invisible, his brain lit up with the realization that he could rule the world. But fi rst, he knew he must develop the invisibility antidote. He thought he could work quietly in the resort town of Iping during the off-season, but his bandages and behavior aroused too much suspicion. Griffi n heads to Kemp’s house because he thinks Kemp will want to help him rule the world. He explains he is not always invisible; he can be seen when it rains and he must wait a few hours after eating. While he is telling Kemp these things, Griffi n senses that he’s uneasy and can’t be trusted. By now, the Iping policeman has contacted his chief, who thinks the invisible man is a hoax and heads to the Lion’s Head to fi nd out what’s going on. Kemp drives Griffi n to the inn to retrieve Griffi n’s notebooks. The unbandaged and invisible Griffi n kills the police chief and then retrieves his notebooks. The police begin a search for a man who is “invisible and mad.” A radio broadcast causes citizens to call the police station with suggestions for ways to detect an invisible man, such as throwing ink on him, or watching for his breath in the cold air. The police use a net to prove that Griffi n is not in the room with them as they make their plans. In the meantime, Kemp contacted Dr. Cranley who brings Flora with him to visit the bandaged Griffi n. At fi rst Griffi n is tender with Flora, but then he declares that “even the moon is afraid of me” and she knows
Invisibility Steals the Seen
57
he’s mad. The police show up, but Griffi n escapes to kill a trainload of people, then Kemp, and two other men before the police fi nally catch and shoot him. At the hospital, the surgeons can’t remove a bullet they can’t see. Griffi n tells Flora: “I meddled in things that man must leave alone” and dies. The invisibility wears off and we see his skeleton, skin, and then actor Claude Rains for 15 seconds. Commentary: In Bryan Senn’s Golden Horrors: An Illustrated Critical Filmography, 1931–1939 (Senn 1996), a group of 30 prominent writers, editors, and fi lmmakers in the horror genre ranked The Invisible Man the fifth greatest horror fi lm of the 1930s, the Golden Age of Horror.
CHEMICAL INVISIBILITY MOVIES Hollow Man (2000) Production company: Columbia Pictures Director: Paul Verhoeven Screenwriter: Andrew Marlowe, inspired by the plot device in Wells’s 1897 novel Short summary: Dr. Sebastian Caine develops a human invisibility compound and convinces his colleagues to test its effect on him MPAA rating: R Plot description: Chemist Dr. Sebastian Caine (Kevin Bacon) leads a team of researchers working for the Defense Department in a heavily guarded underground facility. Linda Mackay (Elizabeth Shue) is his fi rst assistant and former girlfriend who now secretly dates Matthew Kensington (Josh Brolin), another team member. General Caster (J. Patrick McCormick) is the head of the oversight committee for the project. Caine and his team have already developed invisibility compounds for all types of animals from dogs to apes by placing the animal “out of quantum synch.” The caged, invisible animals are visualized with infrared cameras. It seems that reversion from invisibility becomes more difficult as the animals become more complex. The gorilla has been invisible for some time and has become aggressive. In the opening scene at about 3:30, Caine uses a computer to model a small molecule that he calls a protein. He adds a four-atom bridge to the molecule and subjects it to a computergenerated energy stress test.It remains intact. He is delighted and calls his fi rst assistant, Linda Mackay, to let her know that he is heading into the lab to try out his new formula. In the 5-minute scene starting at about 8:00, the new antidote for Isobella the gorilla (Tom Woodruff Jr.) is called “serial-irradiated protein Caine-125.” The gorilla’s arm is spray-painted to visualize it and the freshly irradiated red serum is injected into her vein. The gorilla’s body becomes visible from the inside out, flowing fi rst to the heart, other organs, bones, muscle, flesh, and fi nally hair.
58
ReAction! Chemistry in the Movies
At a meeting with Defense officials, Caine claims that he and his team haven’t quite solved the reversion problem yet. This prompts the officials to tell him that they won’t fund him for too much longer. On the way back to the lab, Caine says: “Let’s go to phase III: human trials.” When his team demures, he says: “You don’t get anywhere by following rules.” Back at the lab, Caine tells the others that he has volunteered to stay “shifted” for three days so that they can test the reversion formula on him. Veterinarian Sarah Kennedy (Kim Dickens) says, “We’re going too fast. This is bad science.” You suspect right then that she’ll be the fi rst one killed. When Caine has been invisible for 10 days, he becomes angry with the others and decides to leave the facility to escape its claustrophobic environment. After arriving at his apartment, Caine notices that a female neighbor has just arrived home; he enters her apartment and rapes her. This scene earns the fi lm a deserved R rating. Though it leaves you with a sick feeling, it demonstrates that this is a thoroughly modern update of the tale of an amoral scientist. Back in the lab, Kensington fi nally develops an “irradiation signature” that allows the formula to survive the “cohesion test.” When he tests it on a cell culture, however, it proves to be only 97% effective, and everyone is disappointed. Things go from bad to worse for Caine. He kills an invisible dog, General Caster, and then hunts down the members of his team. The cat-and-mouse chase is punctuated by numerous clever ways to visualize the transparent man. At one point, Caine announces over the intercom that he can’t let them take his invisibility away because, “It’s amazing what you can do when you don’t have to look yourself in the mirror.” Commentary: Cocaine has a nitrogen bridge atom that is critical for creating its rigid structure. Now You See Him, Now You Don’t (1972) Production company: Disney Pictures Director: Robert Butler Screenwriter: Joseph L. McEveety, inspired by the plot device in Wells’s 1897 novel Short summary: Undergraduate Dexter Riley invents invisibility spray and uses it to save the school from developer Mr. Arno MPAA rating: G Plot description: Dexter Riley (Kurt Russell), Richard Schuyler (Michael McGreevey), and Debbie Dawson (Joyce Menges) are undergraduates at Medfield College taking Professor Lufkin’s Creative Science course. At a faculty meeting secretly overheard by the students over a walkie-talkie, Dean Eugene J. Higgins (Joe Flynn) says that the school is in fi nancial trouble and that it can’t afford new equipment for the science lab. He tells Lufkin (William Windom), “Just do the same smelly labs you’ve always
Invisibility Steals the Seen
59
done.” Mr. A. J. Arno (Cesar Romero) has recently gotten out of jail and loaned the school enough money to keep it afloat. Arno tells the dean he can pay him back anytime. We know he is really waiting for the right moment to turn the campus into a casino. In the 3.75-minute scene beginning at 6:00, the students are being assessed by Dean Higgins and Professor Lufkin for the likelihood that one of them could win the $50,000 Forsythe science prize. First, Schuyler pours breakfast food and milk into a graduated beaker and then eats it. A second student is working on the fl ight dynamics of the bumblebee. Finally, they visit Riley, who is working on an invisibility formula, using the theories of the nineteenth-century Russian chemist Bersakov who talked of preventing light from reflecting. Unfortunately, Bersakov was not able to test his theory because he was locked up after he went insane. Higgins and Lufkin decide to promote the bumblebee fl ight project. During the night, Riley’s solution is given an electrical charge by a lightning strike, and the next morning he discovers it works. After his fingers become invisible, Schuyler and Dawson enter. Schuyler reluctantly agrees to test it, too, but then immediately worries, “How do we make our fi ngers visible again?” Riley hypothesizes that the solution should wash right off. It does but only after they wash vigorously. The students decide that Arno must be up to something. That night Dawson sprays Dexter and Schuyler with the solution so they can sneak into his office. They discover that Arno is going to foreclose soon and that a 1912 law allows gambling on the college’s land. Next, they help the dean win a game of golf against Timothy Forsythe (Jim Backus) so he will allow Medfield to compete for the science prize. Then, Arno and his gang figure it all out and steal the invisibility formula in order to rob a bank in the middle of the day. The students spray the crooks with water from a fi re hydrant and a car chase ensues. It ends when Arno’s car drives into someone’s pool. The students retrieve the formula and race back to the dean. In the fi nal scene, Higgins and Lufkin are at the Forsythe Award ceremony, where the dean admits that State’s heliospectrograph improvement was incredible. The students arrive to show Forsythe how their formula works, but the dean jumps in the way. He doesn’t disappear until after he walks away and then only his face and torso disappear. Riley says that the spray smells like chlorine and that maybe some pool water diluted it. After the dean faints from not being able to see himself in the mirror, the students toss water on him. Forsythe awards $50,000 to the dean. Commentary: This may be the only Invisible Man feature fi lm in which the principle of action is said to be the inability to reflect light as opposed to refraction or absorption. It is also the best example of washable invisibility paint in the movies. Prior to 1972, invisibility inks and paints were used in a number of cartoons, notably in “The Vanishing Private” (1942) by Disney. That cartoon features Army Private Donald Duck using invisibility paint to camouflage a cannon and then himself.
60
ReAction! Chemistry in the Movies
The bust of Bersakov in the Creative Science Lab looks an awful lot like Dmitri Mendeleev, who discovered elemental periodicity. To our knowledge, it is the only example of “Mendeleev in the movies.” The joke about Mendeleev is that he had his hair cut once a year on New Year’s Day whether he needed it or not. Naturally, most photographers took his portrait on New Year’s Eve. In 1869, Mendeleev was writing an introductory chemistry textbook and was searching for a way to summarize the known elements. After arranging the elements according to mass, he noticed that certain properties repeated themselves. He also noticed that certain elements were missing and predicted their properties. By 1886, three of the missing elements were discovered by other researchers, and they had properties predicted by Mendeleev. He was famous in his own time for his discovery. The Invisible Agent (1942) Production company: Universal Pictures Director: Edwin L. Marin Screenwriter: Curt Siodmak, inspired by the plot device in Wells’s 1897 novel Short summary: Frank Raymond uses his grandfather’s invisibility formula to outwit the Nazis Plot description: Frank Griffi n (Jon Hall) is the grandson of Dr. Jack Griffi n, the fi rst person to develop an invisibility compound. Frank has immigrated to the United States, changed his name to Frank Raymond, and now runs a printing shop in Manhattan. German SS Agent Conrad Stouffer (Cedric Hardwicke) and Japanese Agent Baron Ikito (Peter Lorre) follow his trail and offer to buy his grandfather’s formula. Frank refuses, so they search his place, fi nd nothing, and begin to torture him. Stouffer says, “Weapons are made to be used.” Frank says he’ll talk, and then removes a vial from a drawer, smashes it, and runs. When Frank tells a group of American military brass he was attacked by foreign agents, they ask him to give them the formula. Frank refuses. After Pearl Harbor is bombed, he changes his mind and tells the Defense Department he will give them his formula as long as he is the only one to use it. They agree and ask him to fi nd out when the Germans are going to carry out an aerial attack of New York City. They fly him to Berlin, where he parachutes to his target. After landing, he injects himself with the formula, becomes invisible, and makes his way to his contact, who tells him that Maria Sorenson (Ilona Massey) has a list of Japanese spies in the United States. We learn that the side effect of the formula is that it makes you sleepy at the strangest times. Commentary: The only chemistry in this entry of Universal’s Invisible Man series concerns the drug’s side effect of sleepiness. An amusing plot problem is that Dr. Jack Griffi n must have fathered a child even though
Invisibility Steals the Seen
61
we know from the 1933 fi lm that he was killed before he had a chance to marry his fiancée Flora Cranley. Another bit of trivia is that this fi lm was written by Curt Siodmak, who fled Germany for America in 1937 and then wrote or helped write the screenplays for three Invisible Man sequels, a single Invisible Woman movie, and a variety of other horror and science fiction fi lms, including The Wolf Man (Siodmak 2001). The Invisible Woman (1940) Production company: Universal Pictures Director: A. Edward Sutherland Screenwriters: Robert Lees, Frederic Renaldo, and Gertrude Purcell, from a story written by Joe May and Curt Siodmak, inspired by the plot device in the 1897 novel Short summary: Kitty Carroll uses Professor Gibbs’s invisibility process to revenge herself against her boss Mr. Growley Plot description: Professor Gibbs (John Barrymore) hasn’t invented anything in 10 years, and his benefactor, the wealthy lawyer Dick Russell (John Howard), can’t pay him because he’s broke. It’s bad timing for Gibbs because he just invented a machine and drug combination that confers temporary invisibility. He knows it works on a cat, and he’d like to test it on a human, so Russell gives him just a bit more time. In a short scene at 5:00, Gibbs is working near some chemical apparatus. He adds drops of a solution to a beaker of dark fluid and flames erupt. Gibbs drinks it just as the maid (Margaret Hamilton) approaches and she exclaims: “You drank it!” Gibbs gulps and says: “Yes . . . I had indigestion.” Gibbs puts an ad in the newspaper seeking a volunteer to become invisible. Among the respondents, he picks “K. Carroll,” thinking that “K” is a man. Kitty Carroll (Virginia Bruce) works at the Continental Dress Co. as a model and is able to convince the professor to use her. She doesn’t tell him that she intends to revenge herself on her nasty boss Mr. Growley (Charles Lane). She disrobes behind a translucent screen, extends her arm for Gibbs to inject his solution, and then he turns on his electronic marvel. It works. When Gibbs leaves to fi nd his benefactor, Carroll says, “Growley, Growley, here I come,” and climbs out the lab window. Gibbs returns to the lab with Russell, but when they can’t fi nd the invisible woman, Russell leaves for his vacation. Then, Carroll comes back and Gibbs brings her to Russell’s vacation lodge, where he is convinced of Gibbs’s invention. Carroll is cold from being invisible and naked, so she drinks brandy to warm up. Gibbs cautions her that there’s no telling what affect the alcohol will have. They soon fi nd out that it prolongs the invisibility effect. Gangster Blackie Cole (Oscar Homolka) is living in Mexico to avoid prosecution, but he’d like to return. Upon seeing Gibbs’s newspaper ad,
62
ReAction! Chemistry in the Movies
he sends three henchmen to steal the machine. After they have problems, they kidnap Gibbs and Carroll. Gibbs tells them that the machine will do funny things without the formula. Carroll drinks some alcohol to become invisible again and defeats the crooks with help from the late-arriving Russell. The movie ends with the former Miss Carroll married to Russell and with a one-year-old son. After she applies rubbing alcohol to the baby’s back, it disappears. Gibbs says “Hereditary.” Commentary: This comedic adaptation uses an invisibility process that is closest to the one described in Wells’s book. In the book, Griffi n uses an unnamed bleach for the pigments and two dynamos to adjust the refractive index of the uncolored tissues. He tests the process on a stray cat before he tries it himself. Professor Gibbs in the movie may have been named for Josiah Willard Gibbs (1839–1903), who developed the theories of thermodynamics and statistical mechanics. Gibbs rarely left his native New Haven, Connecticut, during his entire life. He was educated at Yale (1863 Ph.D. Engineering, the fi rst engineering Ph.D. ever awarded) and then became a professor of mathematical physics at Yale (1871–1903). In 1902, Gibbs published his textbook Elementary Principles of Statistical Mechanics that gave his Phase Rule, which describes the energy relationship between the three phases of matter: solids, liquids, and gases. In fact, it was only after the publication of his book that his research caught the attention of other scientists. Today, his theories relating chemical reactions with heat energy, work energy, temperature, pressure, and other physical forces are still used as originally written. The Invisible Man Returns (1940) Production company: Universal Pictures Director: Joe May Screenwriters: Curt Siodmak and Lester Cole, from a story by Curt Siodmak and Joe May, inspired by the plot device in the 1897 novel Short summary: Frank Griffi n, Jack’s brother, helps his partner Geoffrey Radcliffe escape a death sentence and works on the duocaine reversion formula Plot description: Dr. Frank Griffi n (John Sutton) is one of two physicians at the Radcliffe Coal Company. Griffi n gives his dead brother Jack’s invisibility formula to Geoffrey Radcliffe (Vincent Price), the other physician at the company, because he has been unjustly sentenced to death for killing his brother Michael Radcliffe, the owner of the company. In the 3-minute scene at 9:15, Inspector Sampson (Cecil Kellaway) of Scotland Yard visits Griffi n’s laboratory. He blows cigar smoke everywhere and says he knows Griffi n and Radcliffe have been working on a project together. He implies that it was duocaine, the invisibility reversion formula that Griffi n’s brother had been working on before he died.
Invisibility Steals the Seen
63
Griffi n says as little as possible, and Radcliffe is hiding in a remote location, getting irritated from being invisible for so long. In a short scene at 18:00, Griffi n is working with bubbling apparatus and lab animals. He restrains one of his invisible guinea pigs, injects the formula, and watches as its body materializes, skeleton fi rst. It struggles briefly and then dies. In the meantime, Radcliffe tracks down the real killer Richard Cobb (Sir Cedric Hardwicke), who happens to be his cousin. Cobb eludes him by knocking out a light, saying, “I’m as invisible as you are now.” During the fi nal chase scene, Detective Sampson shoots Radcliffe and Cobb, and then Cobb confesses to the murder just before he dies. Radcliffe needs a blood transfusion but Griffi n can’t see the internal bleeding so he tries the antidote. The circulatory system appears, then muscles, and then skin. Radcliffe is alive and happy to see his hand. Commentary: Chemically, this fi lm is important because it names the invisibility reversion formula duocaine. Otherwise, this fi lm is better known as Vincent Price’s fi rst monster movie. Even though he is bandaged except for about one minute of screen time, his omniscient voice-over transcended this fi lm to become his trademark and has been repeated, mimicked, and parodied innumerable times since. For instance, Universal’s Abbott and Costello Meet Frankenstein (1948) ends with Price’s disembodied voice and laughter, and Michael Jackson’s 14-minute Thriller (1984) music video begins with it (Meikle 2003). Also, this was director Joe May’s fi rst movie after arriving from Germany, and he spoke only German. That was fi ne for Price because he spoke the language, but the other actors had to rely on primitive sign language (Smith 2002). The Invisible Thief (1909) Production company: Pathé Freres, France Director: Ferdinand Zecca Short summary: A young thief uses the recipe in Wells’s novel to prepare an invisibility formula Plot description: A young man in a bold plaid suit purchases L’Homme Invisible by G. H. [sic] Wells from a bookstall. He takes the book to his apartment, where he reads it while mixing the powders from three bottles. After he adds cream and pours it into a cup, he drinks it until his head and hands become invisible. He quickly disrobes, enters someone’s house, and steals the silverware. Back in his apartment, he dresses but also adds gloves and a mask. In this form, he returns to the street, where he steals the wallets from a window-shopping couple. The police chase him to his room, where he barricades the door, and calmly waits for them to break it down. They grab him, but he disappears, and they are left holding his clothes. Once the furniture begins flying, they run downstairs, leaving him triumphant.
64
ReAction! Chemistry in the Movies
Commentary: Like many fi lms during this early period, there were no intertitles to explain the plot. The actors pantomimed the entire story. The title of Wells’s book shown at the beginning of the movie is the only clue that a French company produced this one. Indeed, it was distributed in many countries simultaneously, although with different titles. In France, it was L’Homme Invisible and is often listed as such in fi lm catalogs. This 6-minute silent short was produced only 12 years after the novel was published. A copy of it exists at the British Film Institute and at the UCLA Film Archives. The special effect appears to have been achieved by fi lming a man dressed in black against a black background, and then double exposing that image with his apartment room. He sheds his clothes to reveal just the lightest shadow on the screen, which allows us to follow his movements.
3 Isomorphs of Paranoia Chemical Arsenals
CHEMICAL WEAPONS, CHEMICAL WARFARE, AND CHEMICAL ATTACKS In the United States after the September 2001 attacks, citizens were advised to protect themselves from toxic dusts by covering their windows with plastic sheeting and duct tape that could be purchased from any hardware store. One hundred years ago, terrorists would not have had ready access to today’s common chemicals to create makeshift explosives, and citizens would not have had access to plastic sheeting or duct tape to protect themselves from aerosols or gases. Chemical weapons have engendered a cloud of fear since their introduction into warfare during World War I. Recently, the large-scale use of chemicals as lethal weapons has drifted from warfare to terrorism. Chemical weapons are often equated with poison gases (either asphyxiation or nerve agents), but as can be seen in the list of movies for this chapter (table 3.1), they are actually the most diverse type of weapon. Some of these weapons are discussed elsewhere in the book (psychedelic agents, chapter 5; explosives, chapter 9). The chemistry in nuclear weapons movies is discussed in the commentary sections for those movies that use them. The movies in this chapter are closely linked to spy movies, which lie at the nexus of the action and thriller genres. Spy movies are appealing in part because these charming, good-looking government employees live by their wits and gut reactions to make split-second decisions that are best for the spy and the government. But a spy is only as good as the villain; otherwise, it wouldn’t be challenging or fun. So, the fi nal ingredient for the movie choices in this chapter is that many of them refer to actual chemical weapons, which grounds them in the real world. The audience knows these weapons are dangerous and can be misused by the wrong person. Only about 70 chemical compounds have been put to use during military confl icts over the past century, and they are classified based on their effects. Asphyxiating and blistering agents were created for WWI (1914– 1918); nerve agents were developed for WWII (1940–1945) but never 65
66
ReAction! Chemistry in the Movies
Table 3.1. A selection of movies about war, sabotage, terrorism, and chemistry Title (Year)
Chemical Agent
United 93 (2006) The Quiet American (2002) Spider-Man (2002) The Sum of All Fears (2002) The Rock (1996) Fat Man and Little Boy (1989) Die Hard (1988) Bell Diamond (1986) Scanners (1981) The Molly Maguires (1970) Dr. Strangelove (1964) The Incredible Shrinking Man (1957) Miracle in Harlem (1948) Sabotage Agent (1943) The Testament of Dr. Mabuse (1933)
Jet fuel Plastic explosive Performance enhancement Plutonium bomb VX nerve gas Uranium and plutonium bombs C-4 plastic explosive Defoliant Psychokinesis Gunpowder Fluoride Insecticide and nuclear fallout Knowledge and suspicion Poison gas Poison gas and chemical fi re
used in that war; napalm was also created for WWII but it generated public comment only when used in the Vietnam War; nonlethal psychedelics were tested extensively during the 1950s but haven’t been documented as having been used yet; herbicides and tear gas were used tactically during the Vietnam War. Even though many nations, including the United States, stockpiled chemical weapons after WWI and continued to develop new ones, they were not used during WWII. This is partly because they are most useful for preventing stalemates such as trench warfare, but other strategies have been developed to prevent entrenchment.
Nerve Agents Nerve agents (also called nerve gases even though they are actually aerosolized liquids) are the most toxic of all chemical weapons (table 3.2). A few pounds of nerve agent can wound and kill thousands of people and will persist in the area. In contrast, hundreds of pounds of chemical explosives yield a short-term event that might wound hundreds in the right location. Since citizens and troops fear nerve agents above all other weapons, they are often equated with “chemical warfare” even though they have rarely been used in war. A number of movies in this chapter deal with nerve agents, and their history is described here, but the inventor of modern chemical warfare is described in the next section. In 1933, Gerhard Schrader became head of pest control development under the new National Socialist regime (Hoenig 2002). In December 1936, Schrader and his assistant developed a toxic compound they named tabun (figure 3.1). Nazi law required a report concerning the discovery of anything with potential military use, and they complied. Tabun
Isomorphs of Paranoia
67
Table 3.2. Relative inhalation toxicities of some past and present lethal agentsa Agent
Codeb
Chlorine Mustard Tabun Sarin Soman VX
CL HD GA GB GD VX
Developed
Type
1915 1917 1936 1938 1944 1952
Toxicity (1)c 13 50 200 400 2,000
Asphyxiating Blistering Nerve Nerve Nerve Nerve
a Data collected from R. Noyes, Chemical Weapons Destruction and Explosive Waste/Unexploded Ordnance Remediation. Westwood, NJ: Noyes Publications, 1996. b “Code” means U. S. Army Code. c Relative to chlorine; all but chlorine are stockpiled the United States in 2008.
O N
C
P
H3C
N
O O CH2 CH3 CH3
Tabun (GA)
F
P CH3
O O
CH CH3 CH3
Sarin (GB)
F
P
CH3 O
CH3
CH C
CH3
CH3 CH3
Soman (GD)
Figure 3.1. The first three nerve agents were discovered and developed by Gerhard Schrader’s team for the Nazis between 1936 and 1944 but were never used during WWII.
proved to be just as lethal against test animals (and presumably humans) as insects and other pests. A few years later, tabun was shown to interfere with nerve transmission and is now considered to be the fi rst deliberately developed nerve agent. The Nazis set Schrader up with a secret military laboratory, and by 1938 his team had also discovered sarin, an even more toxic nerve agent. In January 1940, the Germans built a plant to produce tabun in Dyenfurth-am-Ober in Silesia, now part of Poland. The precursors and intermediates are so corrosive that the vessels had to be lined with silver or quartz. By April 1942, the plant began producing significant quantities of tabun but not without accidents; at least 10 workers died while handling the material. The Allies never learned about this facility during the war despite all of their gathered intelligence. Soon, the Germans had enough tabun to kill everyone in London, and it could have been used to great effect against the Allied invasion at Normandy. One of the mysteries of WWII is why the Nazis never used it. Churchill had warned Hitler that any use of chemical warfare would receive equal retaliation, but he
68
ReAction! Chemistry in the Movies
had stockpiles only of American-produced phosgene at his disposal, and it was ten times less deadly than tabun. In June 1944, Schrader’s team at the Dyenfurth plant developed the third nerve agent, called soman. When the Russians took the plant in August 1944, they dismantled it, brought it to Russia, and began producing their own stockpiles of tabun and sarin. At the end of the war, it was widely publicized that many nations loaded their chemical weapon stockpiles onto ships and released them at sea. Of these three agents, only sarin has ever been used in a war: Iraq used it against Iran in their 1983 confl ict. Sarin has been used most infamously by the AUM Shinrikyo cult that released it during rush hour on the Tokyo subways in 1995. The most toxic and persistent nerve agent is VX (figure 3.2). It was discovered in 1952 by Dr. Ranajit Ghosh, who was searching for new pesticides at Imperial Chemical Industries in the United Kingdom. The tools to discover and assess nerve agents were well established by this time. It is now clear that many countries were engaged in the search for nerve agents after WWII, and many similar chemical structures and potencies were simultaneously discovered. When the British renounced chemical warfare in 1956, they passed the secret of VX to the United States, which began producing it in Newport, Indiana. It has never been used in any war, but many countries have stockpiled it.
Nerve Agent Mechanism Nerve cells have a large central body containing a nucleus just like any cell, but also many short arms called dendrites and one long arm called an axon that ends in short arms called termini. A nerve impulse moves from the dendrites to the termini and involves the rapid directional release of ions across the nerve cell membrane. This electrical impulse, or action potential, is communicated to the dendrites of the next nerve
CH3 H3 C
O
CH N CH2 CH2 S
H3 C
CH
P
O CH2 CH3
CH3
CH3
VX Figure 3.2. VX nerve agent was discovered in 1952. Its structure and potency were kept secret until it was passed to American military officials in 1956.
Isomorphs of Paranoia CH3
H3C
N
69
O
H3C
Acetylcholine
CH3
H 2O
CH3
+
-O
Acetylcholinesterase
+ O
O
CH3
H3C
N+ OH
H3C
Acetate
Choline
Figure 3.3. The enzyme acetylcholinesterase is located in the space between nerve cells. It catalyzes the hydrolysis of acetylcholine to its parts.
cell by way of chemical communication. Specific types of nerves and impulses release specific chemicals to the space between the nerves called a synapse. Nerve impulses that stimulate muscles communicate using the small molecule acetylcholine, an ester of acetic acid and the alcohol choline (figure 3.3). When the nerve impulse reaches the cell terminus, it releases acetylcholine to the synapse. When enough acetylcholine molecules diffuse across the synaptic space and bind to the next cell’s acetylcholine receptors, the electrical impulse of the next cell begins. Two processes prevent acetylcholine from continually stimulating the next cell. An enzyme called acetylcholinesterase is permanently located in the synaptic space and catalyzes the hydrolysis of acetylcholine to acetate and choline, which are not excitatory. At the same time, proteins in the termini of the preceding cell continually recapture acetate and choline for resynthesis of acetylcholine and repackaging. The organophosphate nerve agents such as tabun, sarin, soman, and VX all act by forming a covalent bond with the active site of acetylcholinesterase (figure 3.4). Each of these compounds resembles the structure of acetylcholine, but they tend to be larger and bind to the enzyme more strongly. After they bind, they each have a group, such as fluorine, attached to the phosphorus that is easily and simultaneously displaced upon reaction with the enzyme’s reactive group. The inhibited form of the enzyme is incapable of hydrolyzing acetylcholine. The downstream dendrite is then continually stimulated by the resulting high levels of acetylcholine in the synaptic space.
CH3
H3C O H3C
P
F– F
-O
Sarin (GB)
O H3C
O
Acetylcholinesterase
CH3
H3C
P
O
O
Sarin-Inhibited Acetylcholinesterase
Figure 3.4. The four nerve agents in figures 3.1 and 3.2 react with acetylcholinesterase to prevent it from performing its function.
70
ReAction! Chemistry in the Movies
Nerve Agent Antidotes The Mark I Nerve Agent Antidote Kit (Mark I NAAK) used by the U.S. Army contains an autoinjector with 2 mg atropine and a second autoinjector with 600 mg pralidoxime chloride (PAM) (figure 3.5) (Hoenig 2002). They are intended to be used on yourself or your buddy and are sequentially injected into the thigh (not into the heart with a long needle as shown in 1996 movie The Rock). From there, they travel to the brain, where atropine works more rapidly than PAM. Atropine binds to the acetylcholine receptors to cause them to release bound acetylcholine. Note that atropine has an ester and a nitrogen, like acetylcholine, but that it has a much more rigid structure. When it is bound to the acetylcholine receptor, it stimulates the cell differently from the acetylcholine. PAM reacts chemically with the nerve-agent–inhibited acetylcholinesterase so that the organophosphate group is released and the enzyme can now work again. There are numerous side effects associated with both these antidotes, but they are better than those of the nerve agent. The side effects from this amount of atropine are inhibition of sweating, pupil dilation, dry mouth, mild sedation, and increased heart rate. The side effects of PAM are dizziness, blurred vision, nausea, and vomiting. The Chemical Weapons Convention of 1993 outlaws the production, stockpiling, and use of chemical weapons and their precursors (Noyes 1996). Its official name is “Convention on the Prohibition of Development, Production, Stockpiling and Use of Chemical Weapons and on Their Destruction.” It was signed by many countries, including the United States, in 1993 and came into force in 1997. The Convention gives power to the Organization for the Prohibition of Chemical Weapons to conduct inspections of military and industrial plants in member countries. The inspection team has a complex task because many of these agents are precursors in either the dye or insecticide industries. The most significant outcome of the Convention has been to eliminate production of mustard and nerve gases. As of 2007, more than 120 nations have ratified the Convention, and about 30% of the world’s declared stockpiles have been destroyed.
CH3
CH3
HOH2C
Cl –
N O
N+
N
OH
O
Atropine
PAM (Pralidoxime Chloride)
Figure 3.5. The combination of atropine and PAM can act as an antidote to nerve agents.
Isomorphs of Paranoia
71
FRITZ HABER: CHEMICAL WARFARE CREATOR, NOBEL PRIZE WINNER Fritz Haber fled Germany wearing a false beard at the end of the WWI, fearing he would be tried as a war criminal for having developed chemical weapons (Hoenig 2002; Stoltzenberg 2004). The next year, he was awarded the 1918 Nobel Prize in Chemistry for his prewar development of ammonia synthesis. These events coming in such close proximity to one another demonstrate this brilliant scientist’s belief that science is for all humanity during peace but for one’s country during war. More belligerently, in his Nobel speech, he said: “In no future war will the military be able to ignore poison gas. It is a higher form of killing.” Before WWI, Haber was a chemist with ambition (Stoltzenberg 2004). His early research concerned the effects of electricity and then gas pressures on chemical reaction rates and yields. This allowed him to rise from a professorship at the University of Heidelberg to directorship of the Kaiser Wilhelm Institute for Physical Chemistry in Dahlem, near Berlin. He was able to reach that exalted position in 1911 because he had solved the challenge of producing ammonia from hydrogen and nitrogen (3 H2 + N2 o 2 NH3) in 1909. This was an economically important reaction because ammonia is one of the simplest fertilizers and Germany did not have easy access to imported fertilizers. Even though it appears to be a straightforward reaction, it had confounded other prominent chemists, such as Friedrich Wöhler, who were still living and who didn’t believe it was possible. Haber found that the reaction would yield products under the right combination of an iron catalyst, high temperature, and high pressure. The process is widely used even today to create nitrogen fertilizer and has played an essential role in feeding the world’s continually growing population. It is usually called the Haber-Bosch process because engineer Carl Bosch tested many variables to make the reaction work on an industrial scale. Even though Bosch didn’t share Haber’s 1918 Nobel Prize, he later shared the 1931 Nobel with Friedrich Bergius for the hydrogenation of coal and oil. When the Great War broke out in 1914, Haber saw an opportunity to show the Kaiser he was grateful for his directorship, to demonstrate his love for his country, and to advance his position even further. Almost immediately, he lobbied for and was appointed head of the Chemistry Division in the Ministry of War. During his later war trial, he said he conceived of the idea to develop chemical weapons after the French forces used tear gas to force troops out of the trenches. Haber’s wife Clara committed suicide two months after he began spending all his time preparing for the chlorine attack on Ypres, Belgium. She wrote several farewell letters, none of which survive, made a test shot with Haber’s pistol, and then shot herself in the heart and died within a few hours. Haber’s biographer notes that the reason for her suicide remains unknowable but suggests she was jealous he was having an affair or that
72
ReAction! Chemistry in the Movies
Table 3.3. Chemical weapons used during the First World Wara Agent
Codeb
Introduced
Relative Toxicity
Vapor Density
Chlorine Phosgene HCN Mustard
CL CG AC HD
1915 1915–1916 1916 1917
(1) 20 12 40
2.45 3.4 0.9 5.5
a
Data collected from H. D. Crone, Banning Chemical Weapons: The Scientifi c Background. Cambridge: Cambridge University Press, 1992. b “Code” means U. S. Army Code.
she opposed his involvement in gas warfare (Stoltzenberg 2004). During the early years of their marriage, Haber and his chemically educated wife had a harmonious relationship that apparently ended with the birth of their son in 1901. They drifted apart as Haber spent greater amounts of time away from home. Within days of her death, Haber returned to work and later expressed great guilt over her death. Haber knew that a chemical weapon must have a variety of features (Haber 1986; Crone 1992). It must be deadly—although experience soon proved that an ability to maim had just as much psychological value. It must be inexpensive to produce because it will be needed in large quantities (Haber relied on products that were currently being industrially produced). It must be heavier than air so that it doesn’t float away (most agents were liquids that were vaporized during dispersal). Finally, it must be stable enough to withstand transport in military vehicles and to serve as payload in missiles. Only five chemical weapons were used during battles in WWI (table 3.3, figure 3.6). Each kills and maims by a different mechanism, and none of them is being stockpiled today—they aren’t deadly enough. Fritz Haber fi rst chose chlorine gas because it was produced in large amounts by the German dye industry. In fact, it is still being produced in large amounts by the world’s chemical industries for dyes, paper, plastics, and medicines. Unfortunately, its high toxicity was well known because so many dye chemists had accidently inhaled it. Chlorine gas is the second most electronegative element, meaning it is highly reactive with a
Cl H
C
N
C
O
Cl
S
Cl
Cl
Cyanide Gas (AC)
Phosgene (CG)
Distilled Mustard (HD)
Hydrogen Cyanide
Carbonyl Chloride
Bis(2-chloroethyl)sulfide
Figure 3.6. Structures of three chemical weapons used during WWI.
Isomorphs of Paranoia
73
huge number of substances. The immediate danger is to eyes, nose, and respiratory tract, so it is considered to be a choking (or asphyxiating) agent. The lung’s membranes swell, causing them to fi ll with fluid and, effectively, drown the victim. During its fi rst use, the German Army didn’t expect it to work and didn’t follow up tactically when it did. In subsequent battles, it either floated away from the battle site too soon or killed the German soldiers setting it off. Within a year, the Allies had developed their own chemical warfare agents. Their fi rst was phosgene (Foulkes 1934), which was also used to synthesize dyes. It is a choking agent like chlorine except that it is 20 times more toxic. Since the Germans had the world’s largest dye industry at the time, they quickly retaliated with phosgene of their own. Next, the French used hydrogen cyanide gas, but it did little harm because it was lighter than air and floated away. It is absorbed into the blood stream and is carried to the oxygen-utilizing enzyme in cells, where it binds and inhibits. The victims die from an inability to get enough oxygen to their tissues. The chemical compound most strongly equated with WWI is “mustard gas,” which is actually an oily liquid. It is properly called a sulfur mustard oil, one of many mustards that have been developed. It kills by asphyxiation because its two chlorine atoms are good leaving groups. It probably also cross-links proteins in an unnatural way to cause blistering when it lands on the skin and in the eyes. Many troops returned home blinded, and mustard oil became the most feared of all the chemical weapons. The Treaty of Versailles ended the Great War in 1919 and forbade Germany to use, manufacture, or import poisonous gases. The United States led an international discussion to ban chemical weapons generally, but it went nowhere. The fi rst real agreement most countries signed was the Geneva Protocol of 1925, more formally called the “Protocol for the Prohibition of the Use in War of Asphyxiating, Poisonous or Other Gases, and of Bacteriological Methods of Warfare.” Ironically, the United States did not sign the Protocol, in part because it was moving toward isolationism. More problematic was that two-thirds of the signing countries did so with the caveat that they would not use them fi rst, leaving their options open to use it in retaliation. A third problem with the Protocol was that it provided no method of enforcement. When it was broken by a signing nation, there was no retaliation: Japan signed in 1925 but broke it in 1938 against China; Iraq signed in 1931 but broke it in 1983 against Iran; and Italy ratified in 1925 and broke it in 1935 against Abyssinia. Despite the Geneva Protocol, Haber continued his research on gas warfare under the guise of “pest control.” In 1920, his team had developed Zyklon B [literal translation is Cyclone B] that the Nazis would use 20 years later in the Holocaust to eliminate people they deemed inferior, such as gypsies, homosexuals, the developmentally disabled, and people with any percentage of Jewish ancestry. Zyklon B consists of liquid hydrogen cyanide in an inert porous solid support, such as diatomaceous earth.
74
ReAction! Chemistry in the Movies
The boiling point of HCN is 26°C (78°F) and must be kept under modest pressure to maintain its liquid state. When the National Socialists came to power in 1933, they funneled even more funding into chemical warfare research. Apparently, Hitler remained impressed by his own mustard gas incapacitation during his time as a soldier in the trenches of World War I. The Nazi authorities told Haber he could remain in charge of the program despite his Jewish background (neither he nor his parents had participated in the Jewish faith) but that all of his Jewish workers would have to resign. Haber resigned instead, moved to Switzerland, and died the next year.
LOOSE LIPS AND CHATTER The Molly Maguires (1970) occupies a unique place in movie history. It was made at the turning point when movies with sabotage themes were about to be eclipsed by terrorism themes. Sabotage is the older term and is defi ned as a deliberate action aimed at weakening an enemy, oppressor, or employer through any means including destruction. Terrorism is defi ned in greater detail in the next section but includes the intention of instilling fear in one’s opponent. This movie shows the intertwined nature of these themes. The story line is set in 1876 Pennsylvania, almost 100 years before the fi lm was made. The historic Molly Maguires were a secret society of mostly Irish coal miners who were not paid a living wage for their work. They decided their best protest option was to sabotage their own workplace, thereby terrorizing the company owners. The violent tactics of the Molly Maguires presaged the violence used by such groups as the Weather Underground and the Black Liberation Army in the early 1970s (International Spy Museum and History Associates Inc. 2004). The historic incidents portrayed in The Molly Maguires would have had the effect of contextualizing the radical edge of the civil disobedience movement taking place in the United States at the time the movie was made. There is an insider/outsider dynamic at work in sabotage and terrorism. How do insiders communicate without having outsiders learn what they’re discussing, and how do outsiders discover when some threatening activity is truly about to happen so they can stop it? In WWII the government urged self-imposed censorship under the “loose lips” campaign (figure 3.7). Individual soldiers were given a list of prohibited subjects that related to military information, such as troop movements, locations, and facilities. This campaign also had a home front, or civilian component. Methods of communication have changed since WWII. The Internet and satellite communications networks have made it easier and faster to communicate globally and for those communications to be monitored. Chatter is an intelligence-gathering term for the frequency of communication between suspected terrorists. A fluctuation in the level of chatter is the signal that something is about to happen.
Isomorphs of Paranoia
75
Figure 3.7. “Loose lips might sink ships” poster published by the U.S. government, 1941–1945. Image courtesy of the National Archives and Records Administration.
An examination of sabotage and terrorism themes in the movies shows that sabotage was at its highest level in the 1930s and 1940s but declined to its current level in the 1950s (figure 3.8). Terrorism was an underutilized theme until the 1960s. Since then it has continued to rise unabated. Therefore, our current societal interest lies in movies about terrorism. The selected movies in this chapter show how hard it is to separate these two themes, however. Because of this, chapter 3 is the only chapter with two archetypal movies.
FEAR IN A TIME OF TERRORISM Terrorism, though not new to history, has taken on a new dimension after the events of September 11, 2001. That day, the world’s leading democracy, the United States, came face to face with the brutal force of terror within its borders. On a nihilistic scale and of a planned sophistication never before seen, a new order of disorder seemed to prevail. It
76
ReAction! Chemistry in the Movies
Figure 3.8. The transition from sabotage to terrorism themes in films. The films were obtained from a keyword search for the indicated terms on the Internet Movie Database (www.imdb.com). The raw list was curated by removing TV shows, straight to video, films with non-English titles, and redundancy.
became clear that terrorism would be a constantly evolving phenomenon (Seger 2003) about which free societies will have to remain ever vigilant. But this watchful attitude will need to be of a two-part nature, directed equally toward the terrorist actors themselves and also toward their audiences’ emotional response, including actions that response may provoke, for it is difficult to understand terrorism without considering it in a psychological light (Crenshaw 1990). Defi nitions of terrorism vary. Most include the inducement of fear as a destabilizing strategy of the perpetrator. Fear is an emotional response to a real-world threat or danger; panic, anxiety, and horror are its relatives (Solomon 2006). Terror is fear out of the normal range of experience— the extreme fear that accompanies extraordinary violence. Other goals of terrorist acts include the death and injury of persons and the destruction of property, governments, and ways of life. This destruction may take the form of bombing, assassination, hijacking, hostage taking, murder, torture, and rape, and this is by no means a complete list. For the purposes of this book, some movie examples dealing with terrorism use chemical means to achieve their ends (table 3.1). Although embodying a “creative evil at its worst” (Seger 2003), there is very little actually created by these terrorist methodologies except perhaps the emotion of fear in its direct and indirect victims. It is instructive to note that we have not moved very far from the themes of chapters 1 and 2. We can say the tone of chapter 1 is lateral: left/right.
Isomorphs of Paranoia
77
This opposition reaches into the territory of good/evil. The tone of chapter 2 is of the inner/outer variety and has more connection with a positive/negative opposition. In this chapter we look at defense/offense, but fi nd that good/evil and positive/negative have not left the building. The characters of Mr. Hyde and the Invisible Man form an amalgam of what we would recognize as a contemporary terrorist. In the 1941 version of Dr. Jekyll and Mr. Hyde, Hyde acts in a very graphic way as tormentor to Ivy by psychologically and physically abusing her. Ivy is intimidated and demoralized by Hyde’s constant threats and has lost her personal volition. She is contained in a room as captive prey to Hyde when he makes his horrifying visits. You might say Hyde is subjecting Ivy to a protracted campaign of terror and Ivy is symbolic of all innocent victims of terror. Her palpable fear and hatred of him drive her fi nally to seek the kindly Dr. Jekyll’s help. She feels her only option is to end it by killing herself, because no matter where she goes she believes Hyde will fi nd her. Jekyll assures her that he will see to it Hyde never bothers her again. In a poignant scene, we see Ivy in her room, lulled into a sense of security, now celebrating her freedom by toasting her image in the mirror with a glass of champagne. At this very moment, Hyde dramatically bursts into the room and Ivy’s expression registers shock, disbelief, apprehension, and sad resignation. She recognizes that her world and all the hopes she has for her future are about to be destroyed. He says, sardonically, “Surprised? Could it be that you didn’t expect to see me?” (scene beginning at 1:28:15). Like a terrorist, Hyde is an unpredictable, ruthless, violent, destructive force. Also like a terrorist, everyone is his potential victim: men, women, and children. The only thing saving one from another is sheer chance. There is no predicting where or when he will strike again. Hyde moves in society with the anonymity of a masked man, hiding behind, or in front of, Dr. Jekyll. He has no personal identity. The Invisible Man also shares with Hyde the same terrorist profi le. Unlike Hyde, however, he could be anywhere since he cannot be seen. While Hyde’s name and modis operandi may link him with the word “hide,” the Invisible Man is truly hiding in plain sight. In recent years we have seen governments and law enforcement agencies grapple with the tremendous difficulties associated with apprehending terrorists and preventing terrorist acts, primarily because terrorists blend into society so well. There is often little to distinguish them from ordinary citizens living ordinary lives (International Spy Museum and History Associates Inc. 2004). The invisibility of the terrorist actors goes hand-in-hand with the imperceptibility of their weapons. By using common chemicals that escape detection in crudely destructive ways, the terrorists of today have our chemical world at their disposal. In recent years we have seen swift transitions from normalcy to nightmare with substances as ubiquitous as fertilizer, jet fuel, and peroxide plus acetone, used, respectively, in the Oklahoma City bombing of 1995, the 9/11 attacks of 2001, and the London
78
ReAction! Chemistry in the Movies
Tube bombings of 2005. It has been a reminder to our progress-oriented culture that weapons don’t have to be developed from advanced technologies to have devastating impact; common substances in large enough quantities or unstable mixtures will serve perverted interests just fine. It is unrealistic for any society to grind to a halt simply because it has experienced acts of terrorism, but how compatible are the ideas of a fearful populace with a free society? We know that fear functions very efficiently in our emotional system, and we are able to retain it as a memory that can play over and over again in our minds (Hammond 2003). So, if terrorism is extreme fear, what happens when terrorism is coupled with the word “threat”? There is a curious doubling of the fear component in the variant phrase terror/terrorist/terrorism threat that involves an iteration of fear. The quantity is larger and the volume is louder, compounding the impact of fear. The word threat also extends the fear into the future, a future that may have no known end point because terrorism is unpredictable and the risks associated with the threat may not be fully knowable or measurable (Runciman 2006). When government officials and the media use this phrase in the course of developing strategies to thwart further terrorist acts and inform the public, the idea of fear has been pushed to a drastic limit. Many have suggested that by keeping fear alive in this way, the terrorists’ interests are served (Rogers 2003). In fact, terrorists understand the fear factor and can cause great societal disruption by employing even empty threats and hoaxes. A rational assessment of the overall terrorism risk is called for (see chapter 7 for a discussion of risk). The secrecy inherent in intelligence gathering and its evaluation, however, leaves the public unable to judge whether political leaders are being forthcoming about credible evidence or manipulating, even manufacturing, the evidence of terrorist activity for political ends. As the world now knows, leaders in the United States and Britain had faulty evidence that Iraq possessed chemical, biological, and nuclear weapons of mass destruction (WMDs) and thus posed an international security threat. When speculation linked Iraq to the terrorists behind the 9/11 attacks, it was hard to forget that the leader of Iraq, Saddam Hussain, had used chemical weapons in the 1980–1988 war with Iran and also against his own people during that war. Although faulty, the so-called evidence was used to rally public support for an invasion of Iraq in 2003. When the physical security and survival of the free world is said to be at stake, transparency and accountability in government, the hallmarks of free societies, can easily disappear. In such an environment, the terrorism threat can grow into a political doctrine that could justify any kind of governmental action, including those that might prove detrimental to free societies in a variety of ways. How can citizens weigh, with proportion and balance, the political options to deal with terrorism effectively if they are told by their leaders that they have only one option: fight a preemptive war on terror now to keep from living in the vulnerable and unforgiving emotional clutches of fear for the long haul?
Isomorphs of Paranoia
79
The fear quotient helped to make the case for war in the years immediately following 9/11. Donald Rumsfeld, then U.S. Secretary of Defense, was asked at a news briefi ng in 2002 whether Iraq was willing or attempting to supply WMDs to terrorists, because there were reports that no evidence was found of a direct link between Baghdad and some of the terrorist organizations. In an attempt to justify the idea of invading the nation of Iraq regardless of factual considerations, Rumsfeld infamously responded that there are “known knowns” and “known unknowns” and also “unknown unknowns—the ones we don’t know we don’t know . . . it is the latter category that tend to be the difficult ones” (Rumsfeld 2002). In evaluating the risks and benefits to all possible responses to terrorism, we also need to take into account the consequences of getting assessments wrong and taking actions that could turn out badly. Is a terrorist attack truly the worst thing that can happen to a country, and are all means justified in response, including the taking of innocent lives? Global warming and changes to the environment may pose greater risks to ways of life, political stability, and all human survival than acts of terrorism, yet this issue has not received equivalent attention (Denny 2005, chap. 9). It has been argued that terrorism does not pose risks that are “off the scale,” and although those risks may not be quantifiable, comparative judgments can still be made about them (Runciman 2006). One could say that chemical weapons are feared out of proportion to their lethality. As such, they can become a very successful psychological weapon. In Fritz Haber’s conception during WWI, troops would abandon the trenches, even under a barrage of gunfi re, when they saw a cloud of poisonous gas approaching. Advances in protective gear meant fewer soldiers died from gas attacks after they lost their initial element of surprise early in the war. But because poisonous gas was often invisible, it became a tactical advantage for one side to cause the other side to think a gas attack was imminent, whether it was or not. This would necessitate that the targeted soldiers would have to wear uncomfortable and cumbersome protective gear. The advantage was a tired and slow opponent. After WWI, use of chemical weapons in warfare received international condemnation, although it was still surprising to many that they were not used in WWII. It wasn’t until the Iran-Iraq War (1980–1988) that chemical (and biological) weapons were used on a large scale again. Some of the lasting images of the 2003 march to Baghdad were U.S. soldiers in “clumsy, hot, awkward” chemical-protective suits (Miles 2004). It is widely acknowledged that not enough preliminary consideration went into the planning for the war with Iraq, leaving the people of that country divided amongst themselves and the United States without an exit strategy. Because of this we can also speculate that not enough consideration has been given to the consequences of U.S. domestic spying and surveillance of its own citizens on the health of its democracy, or the damage to its democratic values at home and abroad when harsh methods are employed in the name of intelligence gathering. In insidious and acute
80
ReAction! Chemistry in the Movies
ways, these responses to terrorism can undermine the working of democracy now and in the future. In the latter case especially, such a response can compromise the core values on which a free society sustains itself. Let us consider two manifestations of these responses to terrorism. Elaine Tyler May, a history professor at the University of Minnesota, has written about the character of public and private space in different political eras, especially that of the cold war. She is examining the changes being made to public spaces in the United States as a result of its priority to protect itself from terrorism. Citing Washington, D.C., as an example, May sees the restrictions in access to public sites of national power through barricades, security gates, and screenings as significant symbolic change. Equating the shape of public life with the shape of public space, she questions what impact this “lockdown” atmosphere will have on citizens’ full participation in their government in the future (LaChance 2007). An even more corrosive and damaging response in the war on terror has been the use by the United States of methods the Red Cross has described as “tantamount to torture” (Goodman 2007a). Ostensibly employed preemptively to gain “actionable intelligence,” despite repeated studies that it doesn’t work and yields unreliable results, the use of torture instead acutely corrupts democratic values, laws, and the medical and psychological disciplines that are pressed into its service (Rubenstein 2007). In August 2007, as an act of protest, best-selling writer and psychologist Mary Pipher of Lincoln, Nebraska, returned a prestigious award she had received from the American Psychological Association the previous year (Goodman 2007b). At its annual meeting, the association had refused to put a moratorium in place banning psychologists from participating in harsh interrogations at Guantanamo Bay and illegal U.S. detention centers around the world, or CIA “black sites.” While some techniques were banned, not all were, and Pipher characterized this as providing loopholes allowing psychologists to continue to participate in interrogation techniques such as sensory and sleep deprivation with the purpose of learning more skillful methods to break people down and legitimize the process of torture (Young 2007). Pipher felt strongly this compromised her profession and linked it to “reverse engineering” medical and scientific knowledge by turning something for the betterment of human life into the tormentor of human life (Goodman 2007b). We have seen an antecedent of this inverted dynamic at work earlier in this chapter in the story of Fritz Haber and the weaponizing of chemicals. As the Pipher scenario illustrates, individual medical professionals and scientists can bring their inner moral compass to bear on their profession and make their voices heard. In a 2007 Aspen Ideas Festival panel discussion about global terrorism, it was noted that Osama Bin Laden’s goal had never been to defeat the United States, but rather to bankrupt it. And by the word bankrupt, the 9/11 mastermind did not mean monetarily, but morally. Panelist Hussain
Isomorphs of Paranoia
81
Haqqani, director of the Center for International Relations and associate professor of international relations at Boston University, said the United States was “getting crazy, they’re going crazy, about civil liberties et cetera,” seeing a form of victory for the terrorists because they are “making Americans non-American” by engaging with the terrorists and “diminishing their value as a nation, as a society” (Haqqani et al. 2007). Daily injections of fear can act as a kind of invisibility formula that changes a free society’s values. The temptation is great to enter a cycle of revenge and counterrevenge, or match “terror with terror” (Soyinka 2004). The danger is that the society that was there before can lose itself and disappear. It is useful to remember that when the Invisible Man made his transition from positive to negative, he went insane in the process. The United States was formerly a beacon to the world in the cause of human rights. Since 9/11, it has employed routinized torture and dehumanizing interrogation methods in a systematic fashion in responding to terrorism. One might ask where the country that upheld human rights as a guiding principal went, and what will be required to fully restore those values. Chemical arsenals can play a significant role in destructive human activity, but the question needs to be asked: Does the greater risk for harm lie in the chemicals themselves, or in our response to them? It seems to me, more and more, that we are being governed by waves of mass emotion, and while they last it is not possible to ask cool, serious questions. One simply has to shut up and wait, everything passes. But meanwhile, these cool, serious questions and their cool, serious, dispassionate answers could save us.(Lessing 1987, page 42)
THE FIRST ARCHETYPE MOVIE: THE TESTAMENT OF DR. MABUSE (1933) Distribution company: Sherman Krellberg; The Criterion Collection Director: Fritz Lang Screenwriters: Fritz Lang and Thea von Harbou, from the novel by Norbert Jacques Short summary: Berlin Police Inspector Karl Lohmann traces crimes to mastermind Dr. Mabuse, who is locked in an asylum Plot description: Hofmeister (Karl Meixner) is a disgraced policeman who is working to ingratiate himself with the police to get his old job back. In the opening scene at a counterfeit shop, Hofmeister is pursued by men who are trying very hard to kill him. After temporarily eluding them, he calls Berlin Police Inspector Karl Lohmann (Otto Wernicke) to tell him the counterfeit ring is part of a much larger criminal conspiracy. Just after he says he knows the name of the leader, the thugs reach him,
82
ReAction! Chemistry in the Movies
turn out the lights, and shoot into the dark. Inspector Lohmann yells into the phone for Hofmeister to tell him what is happening. After a long pause, he hears Hofmeister singing but nothing else. Lohmann sets out to fi nd Hofmeister. Meanwhile, former criminal mastermind Dr. Mabuse (Rudolf KleinRogge who played mad scientist Rotwang in Lang’s Metropolis of 1926) has been locked in an asylum for years and is apparently mad. He writes pages upon pages of notes that his psychiatrist Dr. Baum (Oscar Beregi, Sr.) collects and reads later. In the short scene beginning at 29:45, he sees that Mabuse has written about thin-walled glass bulbs fi lled with a gas. They explode on impact. Dr. Born realizes that this exact event was reported in yesterday’s newspaper. In other scenes, a shadow behind a curtain gives instructions to various thug leaders, who in turn tell their thugs what to do. The shadow talks of a “reign of terror” and a “reign of crime.” At 1:02:30, the audience sees a doodled face on one of Mabuse’s note pages. There are circles for eyes and the word MORD (meaning death) spelled vertically between them that acts as a nose. Below the face is an oval with the word GAS written around the edge and once inside the oval. The chemical formula for water is written below the word GAS. Just as Dr. Mabuse dies, he transfers his personality to Dr. Born. We know the transformation is complete when the dead-eyed Dr. Born says: “My name is Mabuse, Dr. Mabuse.” At 1:48:00: Inspector Lohmann and informer Kent are in Dr. Born’s office searching for clues. First, they find a record player/radio system that the jailed Mabuse had somehow used to communicate as the man behind the curtain. Then, they find a Berlin map showing the location of “Chem. Werke,” plus a date and time that occurred one hour ago. Lohmann reads Mabuse’s notes about a chemical factory fire that cannot be subdued by firefighters and then the scene cuts to firemen trying to subdue a factory fire. Commentary: Dr. Mabuse engages in sabotage to induce a reign of terror, choosing a chemical factory where there are highly reactive compounds that cannot be extinguished with water and, in fact, may be stimulated by it. For instance, when sodium metal is added to water, it reacts violently to form the products that include hydrogen gas, which burns in the generated heat. The National Socialist party had taken control of Germany just before director/co-producer/co-writer Lang wanted to release his fi lm. Even though the Minister of Propaganda Joseph Goebbels decided to ban the fi lm, Lang was invited to meet with Adolph Hitler, who had watched and admired it. At the end of the meeting, Goebbels offered Lang the position of official Nazi documentarian, and Lang accepted. Lang left the country the next day and did not return until long after Nazis had been defeated.
Isomorphs of Paranoia
83
THE SECOND ARCHETYPE MOVIE: DR. STRANGELOVE; OR, HOW I LEARNED TO STOP WORRYING AND LOVE THE BOMB (1964) Distribution company: Warner Brothers Director: Stanley Kubrick Screenwriters: Stanley Kubrick and Terry Southern, loosely based on the Peter George novel Red Alert Short summary: General Jack D. Ripper acts on his belief that public water fluoridation is proof the communists have already begun to take control of the United States Plot description: U.S. Air Force General Jack D. Ripper (Sterling Hayden) sends the 843rd Bomb Wing of the Strategic Air Command to drop their nuclear arsenal on their Soviet Union targets. He is able to do this because he has invoked Plan R, which allows senior military personnel to retaliate if the Soviets have killed the president in a sneak attack. It is clear, though, that the president is not dead and that Ripper is bonkers. U.S. President Merkin Muffley (Peter Sellers), U.S. Air Force General Buck Turgidson (George C. Scott), and other authorities gather in a secret underground control room to stop the bombers and, in anticipation of failure, plan for the next step. Royal Air Force Group Captain Lionel Mandrake (Peter Sellers) is at Burpelson Air Force Base as part of an officer exchange program and has been acting as Ripper’s executive officer. He has been locked in General Ripper’s office and attempts to learn the truth. In the two-minute scene beginning at 45:30, Mandrake learns that Ripper will only drink distilled water, rainwater, and grain alcohol. For Ripper, the government-sponsored water fluoridation projects are proof of a Communist plot within the United States. In another two-minute scene beginning at 55:30, Ripper explains that there are studies under way to fluoridate a variety of other products, including, “Ice cream, Mandrake! Children’s ice cream!” The fi nal proof of this postwar Commie conspiracy is that public water fluoridation began in 1946. The essence of Ripper’s paranoia is that a foreign substance was introduced into the water supply against his personal choice. Commentary: The public water fluoridation programs in the United States are considered to be the greatest success of any public health program. By 2005, 75% of Americans drank publicly fluoridated water. The benefits are significant (60% fewer cavities regardless of age, gender, or economic status), the risks are almost nonexistent, and the cost is very low. The fi rst public water fluoridation projects began in 1945 in Grand Rapids, Michigan. The link between fluoride in drinking water and reduction of tooth decay began its 30-year journey in 1901 in Colorado Springs (Peterson
84
ReAction! Chemistry in the Movies
1997). A young Boston-trained dentist named Frederick McKay noticed that children born in the area had a characteristic tooth problem that was different from children and adults who had moved there. Locally born residents had brown mottled teeth that were virtually caries-free (lacked tooth decay). McKay visited many other communities throughout the West to study the brown tooth problem and never found the causative agent, although he did fi nd that it was related to the local water supply. In 1930, an industrial chemist named H. V. Churchill determined that fluoride was present at high levels in all of McKay’s towns (between 2 and 14 ppm). Further research showed that lower fluoride (about 1 ppm) significantly reduced tooth decay but did not cause browning or mottling. After safety trials and epidemiological studies, fluoride was added to the water supply of Grand Rapids, Michigan, in 1945. At the end of the trial period 11 years later, the average number of cavities in 30,000 Grand Rapids children had dropped from seven to three. Hydroxyapatite is the mineral in teeth and bone that crystallizes between the collagen fibers and gives strength. When fluoride ion displaces hydroxyl ions, it becomes fluorapatite: ⎯⎯ ⎯ → Ca 5 (PO 4 )3 F ⫹ OH⫺ Ca 5 (PO 4 )3 (OH) ⫹ F⫺ ← ⎯ Fluorapatite is even stronger and prevents tooth decay by slowing demineralization and speeding remineralization. Dr. Strangelove is the most critically acclaimed and popular fi lm listed in this book. In 1989, the U.S. Library of Congress selected this fi lm for preservation in the National Film Registry. In 2008, it was ranked #21 in the top 250 fi lms at the Internet Movie Database and was the #1 ranked comedy.
MOVIES ABOUT WAR, SABOTAGE, TERRORISM, AND CHEMISTRY United 93 (2006) Distribution company: Universal Pictures Director: Paul Greengrass Screenwriter: Paul Greengrass Short summary: Hypothetical reconstruction of the highjacking of United Airlines fl ight 93 on September 11, 2001, based on a true story MPAA rating: R Plot description: Multiple simultaneous events are shown as intercut scenes. The movie begins with a Muslim man praying in a hotel room. As his praying continues as a voice-over, the image changes to scenes of city streets and then other Muslim men in a hotel room. When the praying
Isomorphs of Paranoia
85
ends, all of the men drive to Newark Airport, where they pass through airport screening procedures, and one of them gets through with a small knife. Interspersed with this are mundane scenes of the crew preparing the plane and other passengers waiting to load. In the 5.5-minute series of scenes beginning at 7:15, the airplane fuel tank is fi lled from a truck that is labeled “Flammable.” The intercut scenes show the activity in a control tower and people boarding the plane that is only about one-third full. When the tank is full, the plane door is closed and locked. The camera moves effortlessly between many points of view for the remainder of the movie. In brief: the plane takes off; the terrorists kill the pilots and hijack the plane; the passengers learn via their cell phones of similar hijacked planes that crashed into prominent buildings and caused great damage; the passengers devise a brave plan to wrest control of the plane from the terrorists. Finally, the plane crashes in a field in Pennsylvania, killing everyone on board. Commentary: The notes that accompany the DVD indicate that Greengrass and his coworkers visited with nearly all of the families of the passengers of United Airlines Flight 93 to reconstruct the events of that day as truthfully as possible. Some imaginary material was used to fi ll in the gaps. The Quiet American (2002) Distribution company: Miramax Films Director: Phillip Noyce Screenwriters: Christopher Hampton and Robert Schenkkan, based on the same-titled 1955 novel by Graham Greene Short summary: London Times reporter Thomas Fowler vies with American undercover agent Alden Pyle for Phuong’s affection in 1950s Vietnam MPAA rating: R Plot description: Thomas Fowler (Michael Caine) is a reporter for the London Times stationed in the 1950s Saigon. He enjoys his post because he loves his mistress Phuong (Do Thi Hai Yen). He is estranged from his Catholic wife in London, who later refuses his request for a divorce. When the Times calls him home because not much is happening in Vietnam, he decides to follow a rumor about a Communist attack in the North. Alden Pyle (Brendan Fraser) is an undercover CIA agent who imports plastic explosives for the “Third Force,” as opposed to French or Communist forces. He is posing as an American government employee on an economic mission. When Fowler and Pyle meet at a nightclub, Pyle dances with Phuong and falls in love with her, too. When Fowler reaches Phat Diem, Pyle is already there, and together they discover that the townspeople have been massacred. The French Colonial soldiers exchange fi re with some group throughout the night,
86
ReAction! Chemistry in the Movies
but it is not clear who they are. Fowler can’t make sense of it because neither the French nor the Communists would want to kill the people of Phat Diem. Back in Saigon, Fowler learns that Pyle is a supporter of the megalomaniac General Thé, so he decides to interview the general. Phuong discovers from Pyle that Fowler’s wife will never give him a divorce and that Pyle would like to take her away. She switches her allegiance to Pyle, and Fowler is greatly disappointed. He wants to get even and then discovers that Pyle has been supplying plastic explosives to the general just before an explosion in a nearby city square. Commentary: The events depicted in this novel are fictional but are based on Graham Greene’s actual experiences in Vietnam. In 1952, General Trinh Minh Thé did take control of Saigon after a series of explosions in the city that he attributed to Communists but that his troops had actually set. It was later determined that the Americans helped Thé engineer those explosions. After World War II, the French attempted to regain control over Vietnam but were fi nally repelled in 1954 by Ho Chi Minh, a Vietnamese patriot and Communist. The United States fomented opposition to Ho Chi Minh because it feared the spread of Communism throughout Asia. When a civil war fi nally broke out between South and North Vietnam in 1957, the United States formed an alliance with the South and began sending supplies and troops. In Vietnam, this war is called the American War. It ended in a cease-fi re in 1973 with the withdrawal of U.S. troops. Fighting returned soon thereafter and ended with Vietnamese unification and self-government for the fi rst time since 1858. Spider-Man (2002) Distribution company: Sony Pictures Entertainment Director: Sam Raimi Screenwriter: David Koepp, based on characters created by Stan Lee Short summary: Military Industrialist Norman Osborn takes untested performance enhancement drug and becomes the Green Goblin MPAA rating: PG-13 Plot description: Peter Parker (Tobey Maguire) is a science major being raised by his Aunt May (Rosemary Harris) and Uncle Ben (Cliff Robertson). While touring a science exhibit at Columbia University on the Upper West Side of New York City, Parker is bitten by a genetically modified superspider that has gotten loose. He soon takes on the traits of a spider, becoming able to climb walls and shoot webs from his wrists. After Uncle Ben is killed by a criminal, Parker decides to become a crimefighter named Spider-Man. Industrialist Norman Osborn (Willem Dafoe) runs Oscorp and Dr. Mendell Stromm (Jack Betts) is the Director of Research and Development. During a presentation to Army officials, Stromm reports
Isomorphs of Paranoia
87
that the experimental human performance enhancement drug was tested on rodents and increased their strength by 800%. When General Slocum (Stanley Anderson) asks whether there are any side effects, Stromm says that they were observed in one trial, at which point Osborn cuts in to say that it was an aberration. When the general continues to press Stromm for details, Stromm says that it caused insanity and that he recommends they return to the formula. Osborn is angered at this revelation, and General Slocum says he’ll cut funding if there is continued lack of progress. When Osborn and Stromm are alone, at 24:00, Stromm explains that methylbromochloroparacine “catalyzes metabolism” and that it will only take two weeks to change the formula. Osborn says he can’t wait and that he’ll take it himself because “risks are part of science.” Stromm straps him in and then treats him with a watery green gas. Osborn’s muscles and heartbeat increase to dangerous levels and then his pulse stops. When Stromm revives him, Osborn is now the Green Goblin and kills Stromm. Commentary: The references to a Jekyll-and-Hyde type of transformation are numerous. First, the gas is green-colored like Dr. Jekyll’s watery green solution. Next, the Hyde character is stronger and more primitive than the Jekyll character. Finally, there are scenes with either the Green Goblin or Osborn looking into the mirror and one in which his transformative other talks back to him. When the name of the molecule is uttered in the fi lm, there is so much noise in the scene that it is difficult to fully hear its name. This makes it impossible to draw its full structure, although it is possible to say that methyl, bromine, and chlorine substituents are attached to a molecule that may be named paracine. Unfortunately, paracine is not the name of a known molecule. Most fictional movie chemists don’t make weapons. This fi lm is a rare exception, along with Moonraker (1979) and Scanners (1981). The villain develops a “nerve gas” in Moonraker and a psychokinetogenic birth defect pill in Scanners. The Sum of All Fears (2002) Distribution company: Paramount Pictures Director: Phil Alden Robinson Screenwriters: Paul Attanasio and Daniel Pyle, based on the same-titled 1991 Tom Clancy novel Short summary: The source of a terrorist nuclear bomb is discovered by isotope analysis MPAA rating: PG-13 Plot description: This is a story about a nuclear bomb being used as a terrorist weapon. In the opening scenes of the movie, the bomb is being transported in an Israeli plane during the 1973 Yom Kippur War when
88
ReAction! Chemistry in the Movies
the plane carrying it is shot down. The bomb doesn’t ignite and is recovered by scavengers in 2002. They sell the bomb to middle men, who sell it on the black market for $50 million to Richard Dressler (Alan Bates), an Austrian neo-Nazi. He intends to use it to incite the United States and the Russian Federation to fight one another. This way they will be distracted and leave the other nations alone. Jack Ryan (Ben Affleck) is a young analyst at the Central Intelligence Agency who has researched the life of Alexander Nemerov (Ciaran Hinds). Ryan’s knowledge becomes important when Nemerov becomes the president of the Russian Federation and is expected to use his military power with a strong hand. When Nemerov later admits that he authorized the use of gas weapons in a battle in Chechnya, the U.S. president sends peacekeeping troops. This incites Nemerov’s ire and makes the two wary of one another. In the meantime, Dressler has paid three Russian nuclear scientists to reactivate his bomb and then has it delivered inside a cigarette vending machine to a Baltimore sports stadium. After Ryan learns of the bomb’s location, he tries to contact CIA Director William Cabot (Morgan Freeman) only to fi nd he is attending a football game in the stadium with the president. Even though the message is garbled, Cabot realizes something is afoot and removes the president from the stadium only minutes before a nuclear explosion destroys a large part of the city. The president survives but Cabot is in critical condition in a makeshift hospital. The suspicions and actions of the United States and Russia escalate to the point that both order destruction of limited targets but prepare for massive attacks. In the one-minute scene beginning at 1:34:00, Ryan locates the Army Radiological Assessment Team just as they test some samples taken from the explosion site to determine its isotopic composition. One worker says, “Wow, check out that gadolidium reading” and the other says, “Yeah, the mass fraction is huge.” They use a book (we only see names of production plants) to conclude it was produced at the Savannah River plant. They can even tell which section of the reactor and on which day any given sample was produced. Gadolidium is typical of nuclear weapons material produced at Savannah River, whereas material produced at the Hanford site in Washington State always produces promethium. Ryan has only minutes to fi nd out who detonated this bomb created by the United States so he can provide the president with good information to prevent a nuclear war. Commentary: Plutonium-239 (Pu-239) is the plutonium isotope used in nuclear weapons. Its nucleus has 94 protons and 145 neutrons (239 = 94 p + 145 n) and is stable for tens of thousands of years. When it is bombarded with neutrons, some Pu-240 is formed and its nucleus immediately splits into two roughly equal halves by a process called fi ssion. Fission also releases vast amounts of energy, along with two or three neutrons, which can initiate the decay of adjacent Pu-239 nuclei in
Isomorphs of Paranoia
89
a chain reaction. At high density of highly pure Pu-239, the chain reaction occurs fast enough to generate a nuclear explosion. The Savannah River site in South Carolina produced Pu-239 for America’s nuclear weapons between 1953 and 1992. It is now a nuclear materials processing plant. So many steps are involved that the effort requires the dedicated activities of many government-sponsored engineers. Pu-239 has to be produced artificially because its natural abundance is so low. The process begins with one of the uranium-containing ores, usually pitchblende, that are mined around the world. The uranium extracted from these ores is a mixture of three natural isotopes: 0.01% U-234, 0.71% U-235, and 99.30% U-238. Of these, only the U-235 nucleus is prone to rapid decay when struck by a neutron, while the other two simply absorb the neutron. So, the natural uranium is bombarded with neutrons until nearly all the U-238 atoms have been transformed to Pu-239. As in every purification procedure, the desired product is isolated or created but is also accompanied by contaminants that are characteristic of the specific purification and synthetic procedures. These contaminants create a unique signature, or tracer, that can be used to identify its source. The Rock (1996) Distribution company: Buena Vista Production company: Don Simpson/Jerry Bruckheimer Pictures Director: Michael Bay Screenwriters: David Weisberg, Douglas Cook, and Mark Rosner Short summary: Rogue marine commandos threaten to use VX gas missiles against San Francisco MPAA rating: R Plot description: Marine Brigadier General Francis X. Hummel (Ed Harris) and a rogue team of his former Marine commandos take Alcatraz Island and the 81 people touring it at the time. They position 15 VX missiles, stolen from a secret government base, toward San Francisco and demand that the government pay $100 million to the families of commandos lost in covert actions around the world. To respond to the threat, FBI Director James Womack (John Spencer) sets up a command center in San Francisco and arranges with the Pentagon to have some recently developed thermite bombs brought to use against the terrorists. They will burn hot enough to destroy the VX poison and everyone on the island. Womack also gathers John Patrick Mason (Sean Connery), Stanley Goodspeed (Nicholas Cage), and a team of U.S. Navy SEALs. Mason is the only person to have ever escaped from Alcatraz, but he can’t be trusted to complete his mission. Mason was a highly trained captain of the British Special Air Service who was sent to Alcatraz for possessing a U.S. Government microfi lm. We learn the fate of this microfi lm
90
ReAction! Chemistry in the Movies
at the end of the movie. His home country denied any knowledge of him or his actions when he was tried and imprisoned. After he escaped, then FBI agent Womack recaptured him and sent him to a maximumsecurity prison, erasing all records that he ever existed. Mason proves several times that he can escape at will. Dr. Goodspeed is an FBI warfare specialist who earned his B.A. at Columbia and his M.S. and Ph.D. in Biochemical Toxicology at Johns Hopkins. Goodspeed has problems with his girlfriend and has never seen field service but is willing to do his best. He explains: “Well, I’m one of those fortunate people who like my job, sir. Got my fi rst chemistry set when I was seven, blew my eyebrows off . . . been into it ever since.” In the 16-minute scene starting at 12:15, Goodspeed is wearing full body protection as he works in a sealed room checking the contents of a package for sarin, identified as a poisonous corrosive gas. After he pulls out a doll, it begins spewing gas and everyone outside the sealed room flees in terror. Goodspeed is told, “Inject atropine into your heart now!” Commentary: VX nerve gas and the atropine antidote were described earlier in the chapter. The Rock was the seventh highest grossing fi lm of 1996, and its usefulness as a potential teaching tool with regard to VX nerve agent has already been noted (Wink 2001). The movie certainly sparks curiosity about atropine’s nerve gas antidote effects. Fat Man and Little Boy (1989) Distribution company: Paramount Pictures Director: Roland Joffé Screenwriters: Bruce Robinson and Roland Joffé, from the story developed by Bruce Robinson Short summary: General Leslie Groves persuades physicist Robert Oppenheimer to begin and then to complete the Manhattan Project MPAA rating: PG-13 Plot description: It is September 1942, three years into WWII and nine months after the United States entered the confl ict. The chain of events begins when General Leslie Groves (Paul Newman) is given a noncombat job to work with scientists. He’s not happy about it. To get up to speed, he visits Leo Szilard (Gerald Hiken) in Chicago (short scene at 5:00). Szilard explains that you purify the U-235, arrange for two portions of it to be brought together suddenly, and the resulting mass will undergo a spontaneous, self-generating reaction. The general fi nds it easy to convince Robert “Oppy” Oppenheimer to lead the team of scientists by appealing to his vanity. Oppy believes every word of it and tells his mistress that he will soon have the general wrapped around his fi nger. By April 1943, the Los Alamos complex is ready and the scientists begin arriving. At one of the fi rst meetings, Oppenheimer says they need three things to make the bomb: physics,
Isomorphs of Paranoia
91
uranium from Oak Ridge, and engineers to build it. In the fi rst half of the movie, the scientific discussions concern the best way to bring the uranium or plutonium together to create the critical mass. For instance, at 22:00, the demolition expert talks with physicist Michael Merriman (John Cusack) about the problems he’s been having. While doing so, it occurs to him to crush an orange in analogy to a hollow sphere collapsing to form a smaller solid sphere with higher density. In the scene at 38:45, Oppy is pondering some notes he wrote on a blackboard. We never learn his thoughts, but the equations appear to concern the movement of detonated materials down a cylinder. This relates to their next problem, which is how to bring two smaller masses of uranium together from the opposite ends of a cylinder. The general storms into the room like a bullet to accuse Oppy of jeopardizing the project. His intelligence men tracked him as he left the base without permission to meet, talk, and spend the night with his mistress, a member of the Communist party. In scene 12 on the DVD, titled “Pros and Cons,” the European war has ended and Oppy believes there is no further need to continue the project. He also belatedly learns that the Germans weren’t even close to creating their own atomic bomb. The general tells him that the work must continue because just having these bombs will act as a deterrent to other aggressor nations. The scene cuts quickly to Oppy’s wife Kitty (Bonnie Bedelia) holding a crumpled piece of paper written by Oppy (figure 3.9).
CONTINUE? YES
NO
1) Responsibility to the project
1) responsibility to humankind
2) seen as a failure of courage to give up
2) effects
3) the future of energy
3) can we control it?
4) [unreadable] can make the [unreadable]
4) Pandora’s box???
5) might fall into the wrong hands
5) No going back
6) [unreadable]
6) weapon! inhumane?!
7) [unreadable]
all weapons!
8) [unreadable] 9) Won’t necessarily be used THE THREAT ALONE
Where are you God! Hah!
Figure 3.9. In the movie, J. Robert Oppenheimer tabulated these pros and cons as to whether to continue leading the development of nuclear bombs after the European War ended.
92
ReAction! Chemistry in the Movies
Later in the movie, she’ll say to Oppy: “What about love and understanding? I thought that’s what science was all about.” Commentary: The best part of the fi lm takes place in the second half after Oppenheimer answers “Yes” to the “Continue?” question at the top of his note (figure 3.9). The fictional Merriman is the one who relays a petition from scientists in Chicago, represented by the Leo Szilard character, to Oppenheimer and the other scientists working at Los Alamos. The scientists request a demonstration during which one of the bombs is detonated while being observed by representatives from all nations before they are used to kill innocent people. Most characters in the movie represent real people, but the character of Michael Merriman played by John Cusack is a composite of two scientists (Newtan 2007). In the movie, the amiable Merriman uses a screwdriver to open and close a half-sphere shell surrounding a sphere that glows blue. After he accidentally spills the contents and retrieves it by hand, he tells everyone to mark their positions and then get out. His slow gruesome death is intercut with scenes of testing the fi rst nuclear bomb at the Trinity site, letting the audience see the effects that will soon be delivered to the citizens of Hiroshima and Nagasaki. The dramatic goal was given precedence over historical accuracy so that the audience would have some understanding of the effect that a nuclear explosion would have. Since the movie is presented as being mostly historically accurate, this episode makes Merriman and, by reflection the other workers, appear to be overconfident about their abilities and underconcerned about the effects of the consequences they are producing. The fi rst actual nuclear accident at Los Alamos occurred less than a month after the two bombs had been deployed over Japan. First, lab assistant Harry Daghlian accidentally dropped a tungsten carbide brick onto a plutonium core during criticality testing and died 21 days later from the energy released. Eight months later, lead scientist Dr. Louis Slotin and others were performing criticality tests on the same plutonium core. Slotin held the neutron-deflecting upper beryllium hemisphere with one hand while using a screwdriver in his other to open and close the sphere. Apparently, the screwdriver slipped, causing the hemisphere to fall fast onto the plutonium sphere so that a large burst of radiation emanated. As Slotin pulled away, his hand pulled the hemisphere with it but then dropped it. The fi rst half of the movie is frustrating for a scientist viewer because so little of the science is described. For instance, the audience never learns that the Fat Man bomb contained plutonium-239 (Pu-239) or that Little Boy contained uranium-235 (U-235). Instead, we are shown the shell casings and only know they are the names of the two bombs. We learn that plutonium was harder to work with than uranium but the reason it had to be imploded isn’t described. In the summer of 1942 at the University of California–Berkeley, J. Robert Oppenheimer held the fi rst conference to consider the creation
Isomorphs of Paranoia
93
of atomic weapons (Los Alamos National Laboratory 2004). The consensus of the conferees was that it would be possible to create such a bomb and that it could be detonated by shooting one piece of either Pu-239 or U-235 into the other to create its critical mass. Ultimately, this proved suitable for uranium, which is why it required a long, thin casing, and was originally code-named Thin Man but later Little Boy. To maintain a nuclear chain reaction, it is necessary to have an amount of material called the “critical mass.” In the case of nuclear weapons, the critical mass is determined by the rate of neutrons generated during the reaction and the rate at which adjacent atomic nuclei absorb those neutrons. When the nuclei of certain isotopes absorb neutrons, they undergo fi ssion. Fission occurs when the nucleus roughly splits in two, coincidentally sheds neutrons, and releases large amounts of the energy that had been holding the nucleus together. Each isotope has its own critical mass, which differs depending on its purity and the intended use. When nearly 100% pure and used as the energy source in atomic weapons, the critical mass of U-235 is 52 kg, which would form a 17-cm diameter sphere. When the U-235 is less pure, a larger mass is required. The critical mass for Pu-239 is only 10 kg, which would form a 10-cm sphere. In nuclear weapons, the isotopes must be kept at a subcritical mass until detonation. In the spring of 1944 at Los Alamos, the fi rst samples of Pu-239 became available and were found to emit high levels of powerful gamma radiation. It was determined that some Pu-241 was cocreated and copurified with the Pu-239. Pu-241 decays rapidly to Am-241, which is very longlived and emits strong gamma rays. As the plutonium ages, the amount of Am-241 builds up and becomes increasingly dangerous to handle. If this plutonium were detonated like uranium, there would be a low-grade detonation that would blast the two pieces of metal apart to create a fi zzle rather than an explosion. In August 1944, Oppenheimer reorganized Los Alamos to focus on developing a suitable implosion mechanism for plutonium, some of which was presented in the movie. Die Hard (1988) Distribution company: 20th Century Fox Director: John McTiernan (who also directed Medicine Man; see chapter 9) Screenwriter: Jeb Stuart, from Roderick Thorp’s 1979 novel Nothing Lasts Forever Short summary: New York Policeman John McClane single-handedly thwarts a group of terrorists in Los Angeles using his wits and their C-4 plastic explosives MPAA rating: R Plot description: New York policeman John McClane (Bruce Willis) arrives in Los Angeles at the Nakatomi Corporate building, where his
94
ReAction! Chemistry in the Movies
wife Holly Gennaro (Bonnie Bedelia) is the vice president. Just before he arrived, she gently fended off the advances of a coworker at the Christmas party while wishing she knew whether McClane was coming for his semiannual visit with the children. They separated so that she could take this job, and she has been very successful. He still doesn’t want to leave New York, but he does love her, although he is incapable of saying so. After McClane arrives in Gennaro’s office on the 30th floor, they greet and argue briefly, and then he steps into her executive washroom to clean up after his journey. At that exact moment, a Pacific Courier van races into the building’s garage, and men emerge carrying weapons. Their leader is Hans Gruber (Alan Rickman); they speak German, and they later call themselves terrorists even though their only goal is to steal some bonds held in the company safe. After cutting the telephone lines and overriding the building’s electronic controls, the terrorists gather all of the company personnel into one room on the 30th floor. McClane escapes unnoticed up the fi re stairs. Soon thereafter Gruber shoots a bullet into the head of the company president when he refuses to give the safe’s access code. Gruber has a backup plan to retrieve $640 million in negotiable bonds from the safe. Next, McClane sets off a fi re alarm that unfortunately only alerts Gruber to his presence on the 32nd floor. One of the terrorists hunts McClane down but is killed by McClane instead, almost accidentally. McClane uses the dead thug’s walkie-talkie to send out a general S.O.S., which is picked up by the police, who think it is a crank call but who respond anyway. McClane is also now in possession of a machine gun, a small cache of C-4 plastic explosives, and detonators. Commentary: The chemistry of explosives is described in greater detail in chapter 9. Briefly, C-4 plastic explosives consist of a small portion of RDX explosive mixed with a plastic material. Since RDX is not particularly shock-sensitive, it has to be detonated with a capsule containing a powerful shock-sensitive explosive. This fi lm is ranked #133 out of the top 250 fi lms on the Internet Movie Database, which indicates that there are a great many people who enjoy this sort of over-the-top action fi lm. It is probably the most viewed movie about plastic explosives, which have been around since their development by the U.S. Army for use during WWII. The fi rst appearance of plastic explosives in the movies may be the 1958 version of The Quiet American. The 2002 remake of that fi lm is described in this chapter. Bell Diamond (1986) Distribution company: www.jon-jost.com Director, cinematographer, editor: Jon Jost Scenario and dialog: Improvised by Jon Jost and the cast Short summary: Vietnam veteran suffers emotionally after the war, possibly from Agent Orange exposure
Isomorphs of Paranoia
95
Plot description: Jeff (Marshall Gaddis) watches TV sports and drinks beer all day. There is no work for anyone since the Bell Diamond mine was closed in Butte, Montana. Cathy (Sarah Wyss), his wife of seven years, decides to pack up and leave him because they don’t do anything together and he won’t do any chores. He insists that he wants to hear the truth, so she says she really wants to have a baby but “I know you can’t do that for me.” He’s upset by this revelation, but he knows it’s the truth (figure 3.10) and helps her pack. Jeff follows friends Dan and Nick around the empty town and then to the Bell Diamond mining complex. Dan and Nick trade swigs from a whisky bottle until it is nearly empty. Dan wants to offer some to Jeff, but Nick says it will only make him angry. At fi rst it does, but then Jeff takes a swig and woo-hoos to create an echo in a tunnel. Cut to Cathy, who talks to her painter friend Hailey about the places they’ve lived as children and adults. Back to the trio, who climb to its highest platform. Dan says he feels that someone is going to open fi re on him and then describes tactical issues with regard to height that he learned in Vietnam. Jeff starts sobbing, and Dan asks if he’s alright. Nick says to leave him alone, and then Dan leaves in anger at both of them. Dan visits Ron and he tells him that “Jeff got killed over there in ’Nam and he doesn’t even know it.” Dan is drunk on whisky and high on
Figure 3.10. Jeff (Marshall Gaddis) is upset upon hearing the truth from his wife. BELL DIAMOND © 1986 Jon Jost, Filmmaker, Cinematographer. Image courtesy of Jon Jost.
96
ReAction! Chemistry in the Movies
marijuana when he leaves. Ron drinks more whisky and listens to a radio call-in program. In the scene beginning at 1:06:00, a woman calls in to radio announcer Alan’s “Party Line” show to ask about the Legion. Ron gets fed up with the conversation and calls in. Ron says: “You don’t know what it’s like with your awards and parades. I have a friend who’s on this gallows thing up there and I think he’s gonna jump.” Alan wants to know why and says that many are out of work. He then asks “Agent What? Try to tell me where he is.” Commentary: Many of Jost’s fi lms remind you that life can seem to be a series of unrelated events that make sense. In this one, Jeff and Cathy haven’t been able to have a baby, but the audience doesn’t know how long they’ve been trying or whether they’ve sought help. The only thing we know is that everyone believes it was because Jeff was exposed to Agent Orange during his duty in Vietnam. At the end of the fi lm, Cathy returns home to Jeff because she’s pregnant. When they reunite, they are both happy to see each other again. Americans were the fi rst to use an herbicide as a weapon during a war (Galston 1969). Agent Orange was used from 1961 to 1971 during the Vietnam War to clear jungle growth to reduce the potential for ambush, and to deprive the enemy of food. The next chapter deals with herbicides and insecticides in greater detail, although the adverse health effects of Agent Orange are briefly given here. Returning Vietnam Veterans complained of a variety of health maladies. Agent Orange was suspected for its ability to cause some of the problems because it was known to contain contaminating dioxin. Subsequent studies found weak links between dioxin and a variety of health effects, with the strongest being diabetes. In the face of insufficient and inconclusive information, the U.S. government provided compensation to some exposed Veterans in 1984, two years before this fi lm was made. Film critic Joe Morgenstern coined the term “size-to-content ratio” to describe “little films with a lot in them” (Morgenstern 2007). He was speaking of independently produced fi lms that offer an intimacy with the characters that can never be achieved in a Hollywood-type fi lm. This is relevant to Jost because Kevin Thomas wrote in the Los Angeles Times, “Jon Jost may well be the most important American fi lmmaker who remains virtually unknown to moviegoers” (Thomas 2007). He produced his fi rst feature-length fi lm in 1974 and has chronicled America since. Even though his fi lms are hard to see because they are shown publicly only at festivals and in special presentations, they are available for purchase through his website. His accessible experimental fi lms with improvised narratives are performed by people who aren’t actors. As producer, director, cinematographer, editor, and distributor, Jost has more control over the fi nal product of his fi lms than anyone since the silent era. In 1991, the Museum of Modern Art presented all of his feature fi lms in a complete retrospective, and that same year, he was the inaugural recipient of the John Cassavetes Award during Independent Film’s Spirit Awards ceremony.
Isomorphs of Paranoia
97
Scanners (1981) Distribution company: Avco-Embassy Pictures, Canada Director: David Cronenberg Screenwriter: David Cronenberg Short summary: ConSec Corporation has created people with telekinetic and telepathic powers, including one who intends to dominate the world and another who intends to stop him MPAA rating: R
Plot description: ConSec is a chemical weapons corporation that purchased Biocarbon Amalgamated from Dr. Paul Ruth (Patrick McGoohan) and renamed it their Scanner Division. Dr. Ruth is a psychopharmacologist who founded Biocarbon Amalgamated in 1942 to produce ephemerol. It is never clear what the product was supposed to do, but in a short scene at 1:31:15, we see a Biocarbon Amalgamated ad in a 1947 Life magazine for EPHEMEROL targeted to pregnant women. The ad was successful enough that ephemerol was given to women during 236 pregnancies, the fi rst two of which were to Dr. Ruth’s wife’s. All of the children were born with telepathic and telekinetic powers and are now called scanners. Dr. Ruth’s two sons don’t know their father, but they have the most powerful scanning powers of all. When the movie begins, they are 35-year-old adults named Darell Revok (Stephen Lack) and Cameron Vale (Michael Ironside). At the beginning of the movie, Vale is a homeless, paranoid man who is able to read people’s thoughts. Some ConSec men capture him by shooting him with a dart. When he awakens, Dr. Ruth tells Vale that he is a scanner and can’t help reading people’s thoughts. In the scene beginning at 23:00, Dr. Ruth injects Vale with ephemerol, which he says is a scan suppressant, and then shows him a black-and-white psychiatric fi lm of Darell Revok when he was 22. Revok had just drilled a hole in his head to let the voices out. Today, Revok knows how to use his powers. Earlier in the movie, Revok caused another man’s head to explode and a policeman to kill two others and then himself. In fact, Revok has created an underground organization of scanners who intend to rule the world. He has been able to woo all scanners away from ConSec, who realize that they’ve lost credibility in the scanner program. Dr. Ruth was called in to solve the problem, and now Vale agrees to help fight Revok. After a series of deadly chases, Vale uses mind control to force a gunman to give him a vial with an Erlenmeyer flask graphic on it. In the scene beginning at 58:15, he and scanner Kim Obrist (Jennifer O’Neill) enter Biocarbon Amalgamated, whose corporate symbol has the same graphic Erlenmeyer flask. Vale and everyone else is in full protective body suits and the building is on fi re. Using his mind to explore the company’s computers, Vale discovers that they make ephemerol for ConSec as part of the Ripe program, but he is not able to access any more information.
98
ReAction! Chemistry in the Movies
Commentary: The name of the drug suggests that its properties were supposed to be fleeting since the Greek word ephemeros means “lasting only a day.” In any event, ephemerol is clearly modeled after thalidomide, with the Hyde formula twist that it calms the telepathic reception for those who were irreversibly transformed by it during fetal development. Thalidomide was developed and marketed as a sedative that found a niche market as relief from the symptoms of morning sickness in pregnant women. It is infamous for causing birth defects in more than 12,000 babies in 46 countries, with the largest numbers occurring in Germany, England, Canada, and Japan, where it was fi rst released to market. Thalidomide prevents neoangiogenesis, the growth of new blood vessels (Melchert and List 2007). At the end of the fi rst trimester of pregnancy, some mothers experience nausea and vomiting brought on by the hormonal surges accompanying uterine implantation of the fetus. This is also the time during fetal development for arms and leg formation, a process that involves angiogenesis. This explains the thalidomide birth defect in which the babies are born either without arms and legs, or stunted arms and legs. After it was being marketed in many countries in 1957, a company applied to the U.S. Food and Drug Administration to market the drug in September 1960. Despite considerable pressure to approve the application, Dr. Frances Kelsey, who had held her position for less than a year, requested further safety tests under the 1938 Federal Food, Drug, and Cosmetic Act. That law had been passed in response to the deaths caused by ingestion of a popular over-the-counter drug that was reformulated so that its solvent was ethylene glycol, a toxic compound. While waiting for the tests to be completed, the news broke in November 1961 in Germany that thalidomide was causing human birth defects. By the time the fiasco had ended, there were almost 8,000 deformed babies in 46 countries. It is now clear that humans are much more susceptible to thalidomide than were the rabbits used in the tests. In the United States, the result was quick passage of the Kefauver-Harris Drug Amendment in 1962, which required that all new drugs be tested for birth effects in two animals at two doses. The Molly Maguires (1970) Distribution company: Paramount Pictures Director: Martin Ritt Screenwriter: Walter Bernstein Short summary: A group of Irish coal miners calling themselves the Molly Maguires sabotage and terrorize the Pennsylvania coal company that employs them Plot description: No words are spoken for the fi rst 15 minutes. First, men and boys work the coal above and below ground. At 9:00, the whistle
Isomorphs of Paranoia
99
blows and the mine empties except for a group of men led by Jack Kehoe (Sean Connery). They set gunpowder at strategic points inside the mine, light a fuse, and leave before it explodes. Next, a Carbon & Schuykill train pulls into a station in 1876 Pennsylvania, and James “McKenna” McParlan (Richard Harris) disembarks. He heads straight for the bar, where a fiddler plays an Irish jig and McKenna asks for a “lager” (the movie’s fi rst spoken word). He plays cards with two men, they start a fi stfight with him, and they are all hauled off to the police station. While talking to McKenna alone in his office, the Police Captain Davies (Frank Finlay) acknowledges him as the company’s undercover agent. He is to report everything he discovers to the captain. After obtaining lodgings and a job as a miner, there is an excellent sequence of McKenna tossing large coal chunks into a wagon, cutting wood, raising a support beam, and mining the coal with a pick axe. On payday, there is a long orderly line to the office. Even though McKenna earned $9.52, they deduct his materials until he receives only 24¢. He’s angry but doesn’t do anything about it. Commentary: The historical sources of information about the Molly Maguires come from the report that undercover Pinkerton detective James McParlan gave to the coal company and from the trial summary at which the men were sentenced to death by hanging. Both of these are strongly biased against the Molly Maguires. In this movie, the men are sabotaging and terrorizing the company to induce them to pay a living wage. The Incredible Shrinking Man (1957) Distribution company: Universal Studios Director: Jack Arnold Screenwriter: Richard Matheson, based on his 1956 novel The Shrinking Man Short summary: Scott Carey shrinks after exposure to an insecticide and gamma radiation Plot description: During opening credits, a man’s outline shrinks as a nuclear cloud grows. When the fi lm begins, Robert “Scott” Carey (Grant Williams) and his wife Louise (April Kent) are boating on the sea surrounded by clear skies. Just after Louise goes below deck for a beer, the boat passes through a scintillating cloud. When she returns, the cloud has passed. Six months later, Scott discovers that his pants are too large. Louise says he’s just losing weight. Two visits to his doctor confi rm that he is steadily shrinking so his doctor sends him to the California Medical Research Institute. In the 3.25-minute institute scene beginning at 13:00, Scott has swallowed a barium solution, been subjected to radioactive iodine, and had a “protein bond test” performed. During the third week of testing, Dr. Silver performs a paper chromatography test for the loss of nitrogen, calcium, and
100
ReAction! Chemistry in the Movies
phosphorus and another for phospholipids and amino acids. While analyzing the results, Dr. Silver tells him he’s had “a rearrangement of molecular structure in cells, an anti-cancer.” The doctor asks whether he’s been exposed to an insecticide and Scott says “Yes, two months ago.” The doctor then asks about his exposure to radioactivity and Scott remembers the mist on the boat. Commentary: In 1954, the crew of the Lucky Dragon, a Japanese fi shing boat, accidentally sailed into a cloud of radioactive fallout from an American hydrogen bomb test on a Pacific atoll. Soon afterward, many of the crew fell ill or died (Lytle 2007). Given that Rachel Carson published Silent Spring in 1962, it is prescient that this 1959 movie supposes that insecticides were carcinogenic. The fi rst step in carcinogenesis is a mutational event. The second step leads to the duplication or expression of the mutated DNA sequence. If the mutated sequence leads to a change in the regulation of cell growth, then the subsequent cell growth will fail to respond to normal regulatory restraints. Both steps can be caused by the strong gamma radiation emitted by some radioactive elements, by some viruses such as papillomavirus, and by certain types of chemical compounds. In a benign cancer, the cells grow locally. In a malignant cancer, the cells break through the tissue membrane cells and invade other locations. The scenario in this fi lm is that the fi rst and second steps are caused by an insecticide and gamma radiation. Miracle in Harlem (1948) Distribution company: Screen Guild Productions Director: Jack Kemp (with an all-black cast and crew) Screenwriter: Vincent Valentini Short summary: Who swindled Aunt Hattie out of her candy making business and then killed two men? Plot description: Aunt Hattie (Hilda Offley) has a well-respected small business producing handmade chocolates in her Harlem apartment kitchen. When the movie begins, she passes responsibility for the company to her niece Julie Weston (Sheila Guyse) and nephew Bert (William Greaves). Weston just returned from New Orleans, and Bert just returned from the Army. He was attached to the Chemical Warfare Unit in the poisons division and is now studying for the ministry. Jim Marshall (Kenneth Freeman) is the son of a large Harlem candy manufacturer. He was sent to the University of Chicago to learn chemistry because “chemistry is a good thing to know in the candy business.” Since returning home, though, Jim has been hanging around with shady characters. He and one of these characters pay a visit to Julie and Bert, who are working in their kitchen. They inform them that Jim’s father now owns Aunt Hattie’s recipe, so they are shutting them down. After
Isomorphs of Paranoia
101
they leave, Julie vows to fi x Mr. Marshall and, a few scenes later, Julie Weston is taken to police headquarters for questioning. It seems that Mr. Marshall died from eating a poisoned chocolate made in Aunt Hattie’s house, and the police fi nd arsenic in her medicine cabinet. Commentary: Chemical education of young African-American men plays a big role in character defi nition in this movie. Bert’s training in the Army would have resonated with the movie’s release shortly after the end of WWII. It’s surprising that no one suspects him to be the poison murderer, but it could be because he is training for the ministry and therefore above suspicion. This movie is considered to be the best of the so-called race movies of the 1930s and 1940s (Jones 1991). These fi lms were produced for African-American audiences and projected in theaters that catered to this audience. Many of these mostly black cast-and-crew fi lms consisted of a dramatic story interrupted by musical or dance numbers; there is something to appeal to everyone. The fi lm scenarios posited a world in which all screen characters had dark skin and in which even the “white” characters were light-skinned African Americans. Hollywood movies of this time featured few darker-skinned characters except as maids, porters, janitors, and so on. This particular musical murder mystery was considered lost until 1983, when it was discovered along with more than 100 similar movies targeted to African-Americans in a warehouse in Tyler, Texas. The “Tyler Texas Black Film Collection” is now housed at Southern Methodist University in Dallas, Texas. Sabotage Agent (aka Adventures of Tartu) (1943) Distribution company: Metro-Goldwyn-Mayer Director: Harold S. Bucquet Screenwriters: John Lee Mahin and Howard Emmett Rogers, based on a story by John C. Higgins Short summary: Captain Terence Stevenson goes undercover as Jan Tartu to destroy a German poison gas factory located in Czechoslovakia Plot description: In 1940 London, Captain Terence Stevenson (Robert Donat) is a bomb defuser with unique credentials. He was born and grew up in Romania, where his father was the consulting chemist at the O. I. Refi nery. He then earned his chemical engineering degree from the University of Berlin. When the British fi nd out that the Nazis are making poison gas on an industrial scale in Czechoslovakia, they ask Stevenson to undertake a covert operation to blow it up. After he accepts, they provide him with the cover that he is Jan Tartu, a chemical engineer who is wanted for using his chemical knowledge against the Romanians. Once he is undercover in Czechoslovakia, he obtains a supervisory job in a bomb assembly factory. He has great difficulty in contacting the
102
ReAction! Chemistry in the Movies
Czech underground to help him in his task because of strong suppression by the Nazis and lack of trust on the part of the resistance. He earns a promotion to the new poison gas factory on the outskirts of town before he has made contact with them. Inside his new workplace, he sees huge gas retorts and meets Dr. Billendorf, who developed the facility. In a 1-minute scene from 1:14:45, he learns that the factory produces two chemicals that are mixed together before they are placed in the bombs. One of the lab pages contains a blurred chemical structure of a methylbenzene connected to at least three carbons. The lab page remains as the background to a montage of Tartu mixing various solutions together as he learns the procedure. At day’s end, Tartu says, “Your new formula is amazing.” Billendorf responds that production is two weeks ahead of schedule and that they will begin shipping out tomorrow. Tartu rushes home make one fi nal attempt to attract the attention of the underground, and this time he is successful. In the scene at 1:25:45, he is brought to their crude chemical lab, where he synthesizes the explosive nitrocene, powerful enough to blow up two factories. Commentary: It is remarkable that this fi lm postulated a nerve agent factory in Czechoslovakia since it wasn’t known until after the war ended that the Germans had built a plant to produce large quantities of tabun in nearby Silesia, now part of Poland. This also happens to be the second fi lm to feature a character identified as a chemical engineer. The fi rst appears to have been the comedy crime mystery The Night Before the Divorce (1942) directed by Robert Siodmak.
4 Bad Company The Business of Toxicity
A COMPANY’S LEGACY IS NOT ITS WIDGETS BUT THE RIVER FILLED WITH DEAD FISH In the movies, chemical companies maximize profits by poisoning their customers, workers, neighbors, and the environment, or they terrorize or outright kill the heroic insider who becomes a whistleblower. English professor Phillip Lopate argued in the New York Times that movies about business in general present a cartoon view of corporate structure (usually there isn’t one), making them the “fantasy villain,” a nearly faceless evil represented in the narrative by a “wall of Suits” (Lopate 2000). Business professor Ribstein goes further and asserts that the overwhelmingly negative view of business in American fi lm narratives is fueled by fi lmmakers who feel their artistic vision is constrained by profit-making capitalists (Ribstein 2005). Ribstein begins his argument with a summary of nine movies about “Evil Corporations.” He doesn’t appear to realize that seven of them were companies that handle or produce chemicals: The China Syndrome (1979), Silkwood (1983), The Fugitive (1993), A Civil Action (1998), The Insider (1999), Erin Brockovich (2000), and Mission: Impossible II (2000). All of these fi lms, and many others, were considered for inclusion in this chapter but, as the fastest growing category of chemistry in the movies, only two from this evil seven made it into the present chapter: Silkwood (1983) and Erin Brockovich (2000) (table 4.1). The evil chemical company theme plays out in several ways. In the deeply satiric comedy Kids in the Hall: Brain Candy (1998), the pharmaceutical company’s happiness drug provides a foundation upon which the comedy troupe bases their humor. This chemical gravitas also lends weight to a number of fictional dramas that explore the theme of toxicity, such as One Man (1977), I Love Trouble (1994), and The Constant Gardener (2005). The company presidents in these movies murder, or hire thugs to murder, the individuals who choose to expose the toxicity of their products. Evil chemical companies are found in “based on a true story” dramas such as in Silkwood (1983), Erin Brockovich (2000), and Bhopal Express (2001). Knowing that the story is based on true tales of
103
104
ReAction! Chemistry in the Movies
Table 4.1. A selection of movies about companies that produce or handle chemicals Title (Year)
Type of Company
The Constant Gardener (2005) Bhopal Express (2001) Erin Brockovich (2000) Kids in the Hall: Brain Candy (1998) Safe (1995) I Love Trouble (1994) Silkwood (1983) The Incredible Shrinking Woman (1981) One Man (1977) Soylent Green (1973) Riders of the Whistling Pines (1949) I Know Where I’m Going! (1947)
Pharmaceutical Pesticide Gas and electric Pharmaceutical Suburban nonpoint sources Milk Plutonium processing Household products Gasoline additive Industrial ecodisaster DDT War chemicals
toxic chemicals lends considerable weight to these story lines. Finally, there are fi lms that don’t take chemical companies to task but instead have something to say about our culture of consumption, which includes the chemical companies, such as Soylent Green (1973), The Incredible Shrinking Woman (1981), and Safe (1995). The two earliest movies discussed in this chapter were included to show different aspects related to the themes of toxicity and chemical company presidents. In the singing cowboy western Riders of the Whistling Pines (1949), DDT is used to deal with a tussock moth infestation. It was the first movie to use DDT in its narrative. In the light romantic comedy I Know Where I’m Going! (1947), the female protagonist is engaged to the president of a chemical company, whom she travels to meet but the audience never gets to see or hear. His company is producing chemicals for use during the war, but he is ensconced on a remote Scottish Isle far from harm. Because most of the current movies pit chemical companies against the environment, it is reasonable to propose that Rachel Carson’s 1962 nonfiction book Silent Spring serves as the intellectual foundation for this theme. Carson used her moving prose to explain the profound interdependence of life on Earth and to call for more thoughtfulness when applying DDT and other insecticides over the towns, cities, forests, and lakes of the United States. She documented numerous cases in which people were killing the good insects along with the bad to cause unintended consequences up the food chain. The strongest opposition to her book came from the chemical companies that were producing DDT, such as Monsanto, Velsicol, and American Cyanamid (Matthiessen 1999). In addition, a U.S. Department of Agriculture scientist was quoted in Chemical & Engineering News as saying that the world would not be able to feed itself without pesticides (Darby 1962). Actually, Carson had written that “it is not my contention that chemical insecticides must never be
Bad Company
105
Figure 4.1. The number of bald eagle breeding pairs has been increasing since 1991 in Nebraska, which is typical of many other regions across the United States. Data are from Nebraska Game and Parks Commission.
used.” Instead, she called for “zero tolerance” toward residues of chlorinated hydrocarbons (e.g., DDT) and organic phosphorus compounds on food, for a more vigilant and more highly staffed force of FDA inspectors, for the use of less toxic chemical insecticides such as pyrethrins, rotenone, and ryania, and for better public education about chemicals in general. Even though the chemical industry (and some members of the federal workforce) opposed her message, the public strongly endorsed her themes. Her book was an immediate best-seller such that the Book-ofthe-Month Club was compelled to run 150,000 copies of her book for its fi rst distribution (Anonymous 1962; Harvey 1962). On the book’s dust jacket, Supreme Court Justice William O. Douglas called the book “the most important chronicle of this century for the human race.” It wasn’t until the 1969 that the U.S. Congress created the precursor to the Environmental Protection Agency (EPA). Among its earliest highprofi le acts was endorsement of a 1972 congressional ban against insecticide DDT use and the 1973 creation of the Endangered Species Act (ESA). Even though the bald eagle had been declared an endangered species in 1967, America’s official bird was among the prominently mentioned species on the new ESA list. In the 1800s, the bald eagle was abundant in every one of the lower 48 states but by 1900 was no longer breeding in the central United States due to overhunting. The widespread overuse of DDT during the 1950s reduced that further until in 1963 there were only 417 known nesting pairs in the Lower 48. The DDT caused their eggshells to soften such that they were too easily crushed during incubation in the nest. The ban allowed the bald eagle to recover (figure 4.1). It
106
ReAction! Chemistry in the Movies
was reclassified as threatened in 1995 and removed from the list in 2007. Its recovery is proof that conservation works and was necessary. Even though Carson’s Silent Spring jump-started the environmental movement in 1962, two high-profi le environmental disasters in 1969 were the immediate impetus that gave President Nixon enough clout to endorse the formation of the EPA so that the federal government could deal with these complex problems (Corwin 1989; Welsh 1989; Sales 1993). On January 29, 1969, a Union Oil Company platform stationed six miles off the coast of Santa Barbara, California, suffered a blowout. Later investigation concluded that the U.S. Geological Survey had been wrong to give Union Oil a variance to use substandard drilling procedures. After the blowout was capped, the resulting pressure buildup led to five breaks along a fault that allowed even more oil and gas to seep out of the submerged earth. It took another 12 days to cap all of these ruptures but the leakage continued for several months. The oil spill received wide news coverage and was described as an environmental disaster. Thirty-fives miles of coastline was littered with oilsoaked diving birds, such as cormorants and grebes. Local citizens from all economic levels volunteered to help clean the birds and the beaches. The fi nal death toll was 3,600 birds, 10 seals, and dolphins, and countless fi sh and marine invertebrates. In a perfect demonstration of how you can’t write better fiction than the truth, Fred L. Hartley, president of Union Oil at the time, was quoted as saying: “I am amazed at the publicity for the loss of a few birds.” On January 17, 1970, President Nixon signed the National Environmental Protection Act, requiring that the environmental consequences of federal projects be considered and that public hearings be held before any permits can be issued. It is likely that the public oversight afforded by this legislation would have prevented the Santa Barbara oil spill. Whereas the Santa Barbara oil spill appears to have been driven by the desire to cut corners for economic gain (for greater company profits and lower consumer costs), the other disaster had been brewing for more than a century. Plants that manufactured consumer products in Cleveland and Akron, Ohio, had been dumping their waste into the Cuyahoga River since the late 1800s. On June 22, 1969, the seemingly permanent oil slick on the river caught fi re in Cleveland (Ohio Historical Society 2005). The Cuyahoga is a shallow, winding river that runs through Akron and then Cleveland before it empties into Lake Erie, the least deep of the Great Lakes. The fi re lasted only 30 minutes and destroyed a mere 50,000 dollars worth of railroad bridges that spanned the river. Far more significant is that the Cuyahoga had caught fi re nine other times between 1868 and 1952. Nevertheless, timing is everything, and the antagonistic connection was forged between uncontrolled manufacturing and the environment in the minds of the citizens. Time magazine used the Cuyahoga River fi re to memorialize the imminent death of Lake Erie and to note the excellent work begun by
Bad Company
107
the citizens of Cleveland to clean up their contributions to the problem (Anonymous 1969). The article described the Cuyahoga as “chocolatebrown, oily, bubbling with subsurface gases, it oozes rather than flows.” The problems with Lake Erie were that it had received nearly a century of industrial wastes from Detroit’s auto companies, Toledo’s steel mills, and Erie’s paper plants. In addition, the 120 municipalities surrounding the lake were dumping “inadequately treated wastes, including nitrates and phosphates” in the amount of 1.5 billion gallons a day. It was the nitrates and phosphates from household detergents that stimulated the growth of thick algal mats that sucked all of the oxygen out of the lake so that nothing else could survive. The nature of this problem was far more entrenched than the one in Santa Barbara. Even though the federal government provided funds to help upgrade the sewage treatment plans for the surrounding towns and to identify acceptable levels for the daily dumping of material from industrial sources, the problems now are the so-called nonpoint sources such as the fertilizer and insecticide runoff from the region’s homes and farms. The chemistry of toxicity can be reduced to two principles: the dose– response effect and the structure–activity relationship (Patnaik 1992). The principle of dose–response toxicity has been understood since Philippus Aureolus Paracelsus (1493–1541) became the father of toxicology when he said that everything was poison and that the poison was in the dose. He was also the fi rst physician to argue that diseases were caused by external factors and should be treated with cures obtained from nature. He used the elements sulfur, mercury, and antimony, which were favorites of the alchemists, to successfully cure diseases such as syphilis. Today, we also consider the duration of the exposure to the toxin. Acute toxicity is easier to study because it is the result of a single dose acquired over a few minutes to a few days. The usual types of acute effects studied in animals are death, tissue damage, birth defects, asphyxiation, and nerve problems. Chronic effects are the result of a small concentration for a long period of time, most often defi ned by 8-hour work shifts. Chronic effects have been possible to study only during the recent scientific era when there have been substantial funds to support such research. A typical chronic effect to study is cancer in animal models because there is a lag of years between the time of exposure and the development of most cancer. People may be more familiar with epidemiological studies that use known doses, responses, and durations in study animals or accidental and work-related human exposures to calculate the likelihood, or risk, for one person to get that disease or symptom. The principle of structure–activity relationships has emerged only during the twentieth century after studying particular effects in depth with hundreds of compounds. The rules of toxicity are still not concrete, but the structure–activity relationship is related to one of the most important principles of organic chemistry: the functional groups. A functional
108
ReAction! Chemistry in the Movies
Table 4.2. The relation between size and toxicity for selected functional groupsa Name
Structure
Effect
Toxicity Level
Smaller aliphatics Larger aliphatics Smaller alcohols Larger alcohols
R—CH 2—R⬘ R—CH 2—R⬘ CH 3OH, CH 3CH 2OH R—CH 2OH
Anesthetic None Blindness; depressant None
Low None High to moderate None
Tearing agents
Moderate to high
Irritation
Low
H
H C
Smaller carbonyls
H
O H3C
H C
O H2C
C CH
O
R C
Larger carbonyls
R
O
a
Levels and effects are collected from P. Patnaik, A Comprehensive Guide to the Hazardous Properties of Chemical Substances. New York: Van Nostrand Reinhold, 1992.
group is an atom or collection of atoms that confer chemical reactivity and physical properties to a molecule regardless of molecular context. The equivocating word in the defi nition is “similar” because molecular context does alter the degree of reactivity and properties. If it didn’t, there would be no need to make new compounds and test them because we would already know their properties. An examination of a few functional groups (table 4.2), out of the hundreds of different types, shows how the size of a molecule affects the toxicity level. Toxic compounds with four or five carbons are usually the most toxic. The exceptions to this carbon number rule are the carbonyls and alcohols, where the compounds with one, two, or three carbon compounds are the most toxic.
SILENT SPRING BY RACHEL CARSON, 1962 Rachel Carson’s 1962 nonfiction book about the widespread overapplication of DDT and other chemical insecticides during the 1950s has been ranked 5th (Modern Library), one of the top 25 (Discover Magazine), and the 78th (National Review) best nonfiction or science book of the twentieth century. In 1999, Time magazine identified Rachel Carson as being among the 100 most important people of the century (Matthiessen 1999). Her book helped create the environmental movement and influenced American public policy. Immediately after its publication, President Kennedy appointed a president’s advisory committee that validated her conclusions in May 1963. Carson’s book and the environmental movement it spawned are largely responsible for driving the establishment of the Environmental Protection Agency, which in turn led to the creation of the Clean Air Act, Federal Environmental Pesticide Control Act, and the Safe Water Drinking Act, among others. All of these outcomes are
Bad Company
109
highly laudable and occurred with considerable public input into scientific policy. With so many positive outcomes and such strong public engagement with scientific issues of national importance, it may be surprising to learn that most entomologists, several large DDT-manufacturing companies, and the American Medical Association all opposed her lop-sided message, which never mentioned that DDT was the best insecticide against mosquitoes. These groups believed that chemical insecticides did far more good than harm. By 1962, its use had already prevented millions of cases of malaria, a world scourge. Unfortunately, some of the chemical companies also responded with uncalled-for personal attacks against Carson’s integrity and qualifications, which backfi red and only enhanced the power of her message in the eyes of the public. Her critics responded weakly, if at all, to her powerful arguments about environmental persistence and the rapid rise of insecticide resistance. Because antibiotics and DDT had received very positive press during WWII, the American public was glad to have access to them when the war was over. The age of paints, plastics, fibers, antibiotics, herbicides, and fertilizers had just dawned and was getting better every day. DDT was and is the most potent, cost-effective, and nontoxic (to humans) insecticide ever discovered. Even in 1962, most entomologists, chemists, and physicians believed that DDT could be used to eradicate the world of its insect-borne diseases. In Carson’s nonfiction book, she did not argue that Americans should stop using chemical pesticides, but that we should not be using these powerful chemicals so indiscriminately and that we should also include biological insecticides in our repertoire. She documented numerous examples of how chemical insecticides killed more good insects than the intended bad and also that the bad insects were becoming resistant to these chemicals. She further showed that these chemicals accumulate in the environment because they are very stable, that they accumulate in the fatty tissues of those animals not killed immediately, and that DDT may be a human carcinogen. Of all of her claims, the carcinogen link was the weakest and remains unproven more than 40 years later, as described in the next section. The history of chemical insecticides was not presented in Carson’s book, but it helps put DDT in perspective. It began with Paris green pigment, which was used in paints and wallpaper (Shepard 1939; Whorton 1974). Paris green has the molecular formula of 3Cu(AsO2)2• Cu(CH3COO)2 and is green because of the copper ions, but its particular shade of green results from the associated arsenic dioxide anion. Paris green was fi rst used as an insecticide in the summer of 1867 by Michigan farmer Byron Markham. His claim to be the fi rst was reported in his letter to Insect Life many years later (Markham 1892–1893). Markham used it against the Colorado potato beetle, which by that time was a pest everywhere in the world that potatoes were grown. It was fi rst reported in central Nebraska in 1859 (Shepard 1939), had spread to the East Coast by 1870, and was in
110
ReAction! Chemistry in the Movies
Europe shortly thereafter. It is possible that Markham fi nished painting the trim on his house and decided to toss the remainder on the insects devouring his potato plants. It must have worked because, within three years, newspapers across the United States and Europe were touting it to control all sorts of agricultural insects. By 1890, it became clear that Paris green was not effective enough against the caterpillar of the gypsy moth Lymatria dispar (Whorton 1974). The gypsy moth is a type of tussock moth that feeds on the foliage of fruit trees and wood trees. Decades earlier, an eccentric individual in Massachusetts had imported this pest into the United States with the intent to breed them with silk moths. In 1869, some gypsy moths escaped and spread in the absence of natural predators. By 1890, the infestation had reached well beyond the borders of Massachusetts and was causing economic harm by destroying entire fruit tree orchards in New England. The cure was almost worse than the disease in that the high amount of Paris green one needed to spray on the trees was enough to damage its foliage. In 1892, lead arsenate was developed as an insecticide by chemist F. C. Moulton, who was working for the Massachusetts Board of Agriculture (Fernald 1898). Insecticide toxicologists seek to determine the lowest dose from which none of the individuals recover (Shepard 1939). Even though lead arsenate was more expensive than Paris green, it was gentler on the trees, and within 10 years it was the preferred insecticide for practically all insects. At fi rst, farmers purchased both lead acetate and sodium arsenate, made solutions of each, titrated the lead acetate into the other to form a white precipitate, and then sprayed the resultant slurry: Pb2+(aq) + HAsO42– (aq) o PbHAsO4(s) (Shepard 1939). Within only three years of the introduction of lead arsenate, there were reports that the residue in the soil was inhibiting the fruit yields and tree growth (Whorton 1974). Unfortunately, farmers were reluctant to stop using it because it was effective and there was no alternative. It was also important that the federal government lacked the power to set limits on the amount of lead arsenate residue on the fruit going to market. By 1920, there were confirmed reports of apples and pears with lethal doses of residue on them and of retarded plant growth. They had been sprayed particularly heavily because these commercially important pomes share the problem that even slight blemishes are undesirable to the consumer (Shepard 1939). The Bureau of Chemistry within the Department of Agriculture began educating farmers about the existence of a lead arsenate residue problem. In July 1927, a new federal department was created that was vested with regulatory powers named the Food, Drug, and Insecticide Administration. In 1930, it was renamed the Food and Drug Administration (FDA), the agency we know today. Even though it had regulatory powers and set limits for pesticide residue on farm products, the FDA of that era was vastly understaffed. The lead arsenate toxicity problems did not go away when DDT displaced it from the market in 1951. Many older orchards still have high levels of it in the soil, and efforts to determine permissible
Bad Company
111
concentrations continue (Creger and Peryea 1992; Lange 1994; Merwin et al. 1994; Peryea 1998). In 1939, chemist Paul Muller was searching for an organic compound that would be effective against the Colorado potato beetle and clothes moths (Whorton 1974; Dunlap 1981). He was working for the J. R. Geigy company of Switzerland, which sought a less toxic alternative to lead arsenate. Muller decided to focus his efforts on testing a variety of chlorinated compounds because these were known to be generally toxic. He soon discovered that DDT was a potent insecticide and practically nontoxic to plants and animals (Lauger et al. 1944). DDT was fi rst synthesized in 1874 by the German chemist Othmar Ziedler as part of a series of academic experiments (Ziedler 1874). It was named DDT because the naming rules of the time came up with p,p-dichlorodiphenyltrichloroethane. When we use today’s naming rules, it is 1,1⬘-(2,2,2-trichloroethylidene)bis(4-chlorobenzene). Its synthesis is remarkably simple: mix together chlorobenzene and trichloroacetaldehyde liquids in a 2:1 stoichiometric ratio (figure 4.2), add heat, watch the white precipitate form, pour the solution into water to dilute the unreacted reagents, fi lter off the white DDT precipitate, and use without further purification. Since the reagents are inexpensive and the synthesis simple, the product is considerably more affordable than lead arsenate. Three years before the Geigy company published its fi ndings in 1944 on DDT’s effectiveness (Lauger et al. 1944), it sent DDT samples to the U.S. Department of Agriculture. They determined its chemical composition and confi rmed that it was simple to synthesize, while the War Fund Administration tested its insecticidal efficiency and the Committee on Medical Research determined that its toxicity for humans was very low (Dunlap 1981). By 1944, American manufacturers were producing two million pounds of DDT per month for the military. Its fi rst success was in controlling a nascent typhus epidemic in Naples, Italy, in December 1943, just as the American soldiers were entering the area. Since the body louse is a carrier for typhus, the U.S. Army sprayed DDT powder on two million civilians and all of their clothing over a period of three months. That the DDT was able to squelch the epidemic in Naples can be contrasted with the 2.5 million Russians who died during the war from
+ Cl
Cl
Cl Cl
H
O
Cl Cl
Cl
+ H2O
+ H Cl
Cl
Cl
DDT
Figure 4.2. DDT synthesis occurs after mixing two compounds together in a 2:1 ratio.
112
ReAction! Chemistry in the Movies
typhus. Its next success was in the Pacific, where it was used to prevent transmission of mosquito-born malaria on a number of islands so that American troops could conquer them. After the war ended, penicillin and DDT were widely reported to have played a key role in winning the war. In 1948, Paul Muller was awarded the Nobel Prize in Medicine for having discovered this wonderful insecticide that could eradicate diseases. A new age of chemicals had dawned. After acknowledging this much-publicized military achievement of DDT, Carson noted in chapter 18 of her 1962 book that body lice around the world were resistant to DDT by 1957. She wrote that resistance to the new organic insecticides was experiencing a “meteoric rise that reached the alarming level of 137 species in 1960.” In 1962, the acquisition of resistance was only just beginning to be appreciated as a biological phenomenon. The mechanism by which resistance is acquired has taken many years to discover, but it has yielded to the pressure of research. On a biological level, the acquisition of resistance to an insecticide, herbicide, or antibiotic results from chronic sublethal exposure to the chemical. Even if an insecticide spraying program is able to eliminate 99.9% of its target, the survivors are much more prone to be partially resistant. In the absence of competitors, their offspring thrive because they are more fit in this new insecticide-rich environment. DDT entered the U.S. marketplace in 1945 and replaced lead arsenate as the preferred insecticide by 1951 (Dunlap 1981). Its low cost, high effectiveness on a broad range of insects, persistence on the crops and in the surrounding soil, and nearly complete lack of toxicity to plants and animals combined to give DDT almost ideal properties. These same properties meant that it could be and was used for public health programs to eradicate insect-borne diseases around the world (Whorton 1974). Its greatest successes have been against small insects such as mosquitoes. The World Health Organization credits DDT with saving 50 to 100 million lives by preventing mosquito-borne malaria. In India, 500,000 people died from malaria in 1960, but only about 1,000 in 1970 after an extensive DDT spraying campaign. Likewise, DDT was used to eradicate malaria from Italy and the southern United States. The 1950s was not the time to rock the government or business boat. America had barely emerged from its decade-long Great Depression when WWII started. Strong cooperation of many industries with the government during the war had vastly improved their facilities and efficiency. Now that the war had ended, people just wanted to get back to work and to be with their families. The industrial efficiency led to general prosperity and a baby boom. The cold war also began, bringing nuclear proliferation and the paranoia that the Soviet communists might blast the Americans off the face of the earth at any time. The push to conform was made palpable by the anti-Communist hunts of the bully Senator Joseph McCarthy. If citizens objected to some policy, did not attend church regularly, or did not participate in their community activities, they were
Bad Company
113
viewed as subversive. The result was that a whole range of political and social issues built up, bursting forth in the 1960s. DDT was on trial by the U.S. EPA between August 1971 and March 1972. Originally, the focus was on its role as a suspected carcinogen. By the time the trial ended, DDT was banned mostly because it threatened fi sh and wildlife (Dunlap 1981). After extensive testimony, the judge declared that “DDT is not a carcinogenic hazard to man.” He also declared that it is not a mutagen or a teratogen or a threat to freshwater fi sh, wild birds, or other wildlife. These latter claims were made with regard to DDT being an acute threat. They ignored the chronic environmental threat posed by DDT and other chemical insecticides that were well documented in Silent Spring. It would be incorrect to state that Rachel Carson’s Silent Spring was the sole reason that DDT was banned, but it certainly started the ball rolling at a time when the public was ready to receive its message. Perhaps the most important political development was the 1969 passage of the National Environmental Policy Act by Congress that created the precursor to the EPA. From the public’s perspective, it was probably equally important that Earth Day was created in 1970 by Senator Gaylord Nelson and Denis Hayes because it transformed environmental protest into an environmental celebration. Thus, Rachel Carson’s book in defense of the environment has the distinction of being one of the primary impetuses for the environmental movement that resulted in Earth Day in 1970, the banning of DDT use in the United States in 1972, and our current programs to “Think Green.”
IF IT SMELLS, IT’S CHEMISTRY; IF IT’S GREEN, IT’S BIOLOGY; SO WHAT IS GREEN CHEMISTRY? The principles of “green chemistry” ensure that the entire life cycle of a chemical product is considered, from natural resource acquisition to disposal, so that its environmental and economic impact are the lowest possible. Green chemistry was fi rst articulated as 12 principles in a book titled Green Chemistry: Theory and Practice by Paul Anastas and John Warner in 1998. They are officials from the U.S. Environmental Protection Agency, and the principles emerged from deep thinking about management of the Pollution Control Act of 1990, which charged manufacturers with far more responsibility for their waste and postconsumer disposal than ever before. Their theoretical stance is to completely reduce or eliminate side products and waste during the manufacture or to manufacture products that degrade quickly and benignly after the consumer has discarded the product. Even though their assumptions of perfect syntheses and fully degradable products are impossible, they acknowledge that the objectives will be realized only through “continual incremental improvements,” which doesn’t sound sexy enough for the movies. Many large chemical companies have voluntarily developed alternate manufacturing
114
ReAction! Chemistry in the Movies
schemes for their biggest-selling products that adhere to green chemistry principles. Also, there are a number of annual competitions to reward the most significant breakthroughs. The 12 principles are readily accessible on the internet at the EPA website, on Wikipedia, and on many other sites. Principle 7, to “maximize atom economy,” sums up many of the other principles nicely because it demands that the synthesis be designed so elegantly that the fi nal product contains the maximum proportion of starting atoms and the smallest proportion of wasted atoms. Some of the other challenging principles are principle 4, use renewable feedstocks such as those made from agricultural products rather than fossil fuels; 6, avoid chemical derivatives such as blocking or protecting agents because they must be removed later in the synthesis and become waste; 9, increase energy efficiency by running reactions at ambient temperature and pressure; and 10, design chemicals and products to degrade after use. The principles of reaction yield (the measure of atom economy), cost, safety, reaction conditions, and ease of workup and purification have been used to create EcoScale (ecoscale. k-supplier.com) (Van Aken et al. 2006), a web-based program that can be used to calculate and compare the relative “greenness” of synthetic procedures. In 2005, Nobelist Ryoji Noyori summarized three of the most significant chemical developments that demonstrate “practical elegance,” his phrase for “green chemistry” (Noyori 2005). He cited the use of supercritical carbon dioxide (CO2) as the reaction solvent, the use of hydrogen peroxide in water-based reactions as the preferred oxidant, and the use of hydrogen gas for asymmetric synthesis. Supercritical CO2 forms when carbon dioxide is heated above 31.0°C under pressures greater than 72.9 atmospheres. Supercritical fluids have properties intermediate between a liquid and a gas. They are useful for chemical reactions because they act as solvents but also allow the reactions to occur faster, as they do in the gas phase. Even though CO2 is a gas that contributes to the greenhouse effect and the consequent global warming, it is “green” for this purpose because it is fi rst captured from the atmosphere and then returned to it for a net zero change. As an aside, these very same properties of supercritical CO2 also make it the ideal replacement for perchloroethylene, the solvent used by dry cleaners. What chemical method does your dry cleaner use?
DAVID AND GOLIATH In Erin Brockovich (2000), Erin says to Ed Masry, “This is kind of like David and what’s his name,” to describe their law fi rm’s fight with corporate giant Pacific Gas & Electric (at 65:30 min). Indeed, the contest between David and Goliath is such a well-known Biblical tale that it’s hardly necessary to fi ll in the blank when it comes to pairing one name with the other.
Bad Company
115
The movies in this chapter featuring evil companies often carry a plot subtext of the David and Goliath story. As omnipotent symbols, these companies invite the expectation that overwhelming odds will need to be overcome to overthrow them. In these movies, over has overdone it and thus perversely attracts its opposite under into the fray. In the Bible legend (I Samuel 17:1–58), young David volunteers to meet the battle-hardened giant Goliath in combat. Against all probability, this inexperienced shepherd boy of small stature is victorious. David goes on to become one of the most celebrated underdogs in Western history. There is little sympathy for Goliath’s demise. How did David do it? How did he get over Goliath’s defenses and under his radar at the same time? He seems to have assessed the situation to take deadly aim at Goliath’s one unprotected spot, smiting him in the forehead with a single stone from his slingshot. The primary oppositions of the David and Goliath dynamic are large/small and top/bottom. These are the dynamics of power. One could also say that the visibility/invisibility theme discussed in chapter 2 has some resilience, for the people who lack influence are often “invisible” to those at the top, and Goliath is a top dog. He would have found it absurd to think that little David, unarmored and exposed, was an equal competitor. Like Goliath, the Konrads Company in the archetype movie One Man (1977) wears a suit of armor. It looks, however, like a business suit. Konrads provides welcome employment for regional workers, and all communities in capitalist societies desire economic growth and expansion. It is precisely this link with current and future prosperity that the elite at Konrads invoke to deflect any criticism of the impact of its operation on the public good. Their fearful workers, actively intimidated by management, feel the pressure to become silent collaborators since they don’t want to lose their jobs. It is hard to motivate community leaders to scrutinize a company that is positively viewed as an economic helpmate. We see how ingrained this perception is in an exchange between reporter Jason, the David-character, and Rod, his boss at the TV station. Jason asks Rod if he thinks the Konrads Company is poisoning people. Rod responds that it may be inadvertently possible, “but I don’t see some big, respectable company . . . ” Jason interrupts, saying incredulously, “Poisoning’s respectable?” (at 53:45). In One Man, and in other movie examples in this chapter, bigger has clearly crossed the line. Whistleblowers like Karen Silkwood are thrust into David and Goliath situations. Her task of outing the truth about the falsifying of plutonium rods takes her into life-and-death territory in Silkwood (1983). Despite legislative protection in the United States against retaliation for informing the government of a violation of specific federal laws, employer power is still little restricted on the federal and state levels. This means unless whistleblowers are unionized or working in the public sector—only 23.7% of the U.S. workforce—they can still be fired for seemingly arbitrary reasons, if race, gender, or another like category is not among them (Robin 2004).
116
ReAction! Chemistry in the Movies
To understand any underdog’s chance at success, we need to consider how others perceive them and how the top/bottom power dynamic can be effectively reversed. Additionally, there’s more underlying such a reversal than just a straightforward inversion, or exchange of positions. With it comes the sense that a healthy balance is being restored. University of Richmond psychologist Dr. Scott Allison is researching the underdog phenomenon and the conditions under which society’s support for the underdog can be strengthened. He has found that while most people intrinsically root for underdogs, defi ned by him as “competitively disadvantaged,” support for the underdog is “highly conditional and fragile” (Fitzgerald 2007). Three studies were conducted examining people’s willingness to lend emotional and behavioral support to underdogs. Study 1 found people giving the most admiration and support to underdogs who gave a 100% effort, with less support going to those exhibiting less effort. Study 2 experimentally divided underdogs into either ingroups or outgroups, defi ned, respectively, as either U.S. citizens or foreign citizens. The results revealed that people view top dogs as more ethical than underdogs, ingroup top dogs as more trustworthy, and ingroup underdogs as most deserving of support. Study 3 found underdogs, rather than top dogs, more likely to be offered behavioral assistance, and that the offer of assistance to a top dog is tempered by the costliness of that act. In conclusion, the research had two main fi ndings. First, an underdog’s perceived level of effort is crucial to gaining support from others (the heroic efforts of Erin Brockovich come to mind here). Second, an underdog’s group membership matters. Because people show bias toward their ingroups, the fi ndings demonstrate that underdogs who share our group memberships will likely receive our support. While the researchers characterize this latter fi nding as “discouraging” on the face of it, they point out that it can be a relatively simple task to change the aspect and fi nd a place of common ground where ingroup status can be shared (Devine and Allison 2006). The pressing problem of global warming gives all of us an incentive to do what we can to meet the goal of a sustainable planet. The odds of solving this problem may seem great, but we know that even small, individual actions count. For example, changing incandescent light bulbs to more energy-efficient light sources can have a David and Goliath kind of impact. Environmentalism has developed into a broad movement that links together issues from the economic, social, and political domains. Viewing ourselves as citizens of the world may be the change in aspect we need to create networks that “forge a solidarity” among different groups and movements (Schlosberg 2005). As we page our way closer to the book’s second half, it is useful to keep in mind that the bright side of chemistry in the movies will have its share of David and Goliath battles, too. This dynamic is closely associated with scientific discovery and the ability to put something new out into the
Bad Company
117
world, and to hold onto it once it’s there. Turning weakness into strength, trusting of instinct, nimbleness, seizing of opportunity; all that . . . and more to come. .
THE ARCHETYPE MOVIE: ONE MAN (1977) Production company: National Film Board of Canada Director: Robin Spry Screenwriters: Peter Pearson and Robin Spry Short summary: TV news reporter Jason Brady investigates Konrads Ltd., which is making deadly biacetylplumbane (BAP) Plot description: Jason Brady (Len Cariou) is a reporter for CKMC-TV Channel 11 news in Montreal. When he brings a gang member to the hospital to be treated for a gunshot wound, he meets social worker Marian Galbraith (Carole Lazare). She just brought a third sick child to the hospital but is reluctant to share the story. In the six-minute scene at 14:30, Brady and cameraman Ernie (August Schellenberg) are back at the children’s ward to learn more about acute BAP poisoning. The doctor says that BAP is like lead or mercury in that it is a cumulative poison. It attacks the liver, brain, and nervous system. Brady asks about the source of BAP, but the doctor says he doesn’t know. Acting on a tip, Brady and Ernie rush to Konrads Ltd. Canada to interview three men at the plant. Two of them want to talk while the third tries to end their discussion. One of them says that one worker died of BAP poisoning recently, and no one is allowed near the location of the leak. Back at the news office, Brady’s boss likes the story and gives him a week to develop it as children versus a big corporation. On the next news broadcast, it is noted that BAP is everywhere. Ninety percent of BAP comes from exhaust fumes. It is a gasoline additive, in paints, and so on. Galbraith hands Brady a Konrads report but won’t reveal her source. She tells him that it is now up to him. At that moment, some men break into her home and assault them. The thugs ask, “What are you trying to do . . . put 600 men out of work?” After they leave, Brady and Galbraith kiss and then the scene cuts to Brady barging into the office of Konrads President Campbell (Barry Morse). He hired the thugs and had them say those things so Brady would think they were union men. Campbell threatens Brady with more violence but then offers him money. Brady turns it down and returns to the news office to discover that the executives don’t want him to pursue the story further. The St. Xavier deaths have been ruled due to highway exhaust and not a Konrads leak. Then, they offer to promote him and he realizes they are now on Campbell’s side (figure 4.3). Commentary: As the credits roll at the end of the movie, a narrator reads statistics in a monotone: 93 died from mercury, 439 in Iran died
118
ReAction! Chemistry in the Movies
Figure 4.3. Jason Brady’s (Len Cariou) struggle with the Konrads Company reaches its climax. ONE MAN © 1977 National Film Board of Canada. Image courtesy of the National Film Board of Canada Stockshot Library.
from methylmercury on grain, and so forth. Finally, the endnote states that the characters and events are fictional and that “the fi rms, products, and incidents are also entirely imaginary.” BAP appears to be based on the real story of tetraethyl lead, which was the octane booster added to automobile gasoline from the 1920s until the 1970s, when it was fi nally banned. This fi lm has all of the complexities of a late-twentieth-century morality play with its layers of dilemmas. Even though Jason Brady is a journalist seeking important stories, “One Man” can be read as the “Everyman” confronted by knowledge of a wrong against society by a specific agent. He is forced to decide whether righting that wrong is more important than the cost to him and his family’s well-being. He is professionally ambitious but could lose his job if he doesn’t drop the story. Brady’s work keeps him from his wife and children, and his risktaking nature even places his life in danger. His professional conduct is questionable when he has sex with his source, a choice that elicits a strong negative response from his wife. The lives of his children are threatened by company thugs because he is trying to save the lives of hospitalized children who have been exposed to the company’s neurological toxin. Brady has to decide where his ethical allegiance lies while seeking proof that the company knew its product was toxic so that he can convince others of the problem.
Bad Company
119
MOVIES ABOUT BAD CHEMICAL COMPANIES The Constant Gardener (2005) Distribution company: Focus Features Films Director: Fernando Meirelles Screenwriter: Jeffrey Caine, from the same-titled 2001 novel by John Le Carré Short summary: After activist Tessa Quayle is murdered in Kenya, diplomat husband Justin searches for the reason MPAA rating: R Plot description: The movie is told in flashback after Justin Quayle (Ralph Fiennes) learns his wife Tessa (Rachel Weisz) and her hired truck driver were killed in remote Kenya. When the British High Commission sent Justin to Nairobi, Tessa asked to go with him so he married her. Upon arrival, Justin works on his backyard garden while Tessa travels with Dr. Bluhm (Hubert Koundé) to remote areas to help the poorest people. In the scene beginning at 25:00, Tessa comes upon a line of people waiting to be treated by medical personnel stationed at a table outdoors. Their sign proclaims they are giving away free Novampen pills so she asks about it. Dr. Bluhm tells her it is an anti-AIDS drug that is being tested in Kenya but isn’t reaching the people who really need it. Tessa is surprised a clinical testing company is giving free tests for HIV and tuberculosis because no drug company does anything for free. When Tessa arrives home, she becomes angry with Justin for using a weed killer. The audience sees that Tessa learns something important and is killed because of it but is in the dark as to what it is. As Justin mourns, he searches through her things and finds a package for Dypraxa, produced by the Threebees Company, which he later realizes is the company that produced his weed killer: “Threebees; Toxipest; Kills Pest and Weeds; New Improved Formula.” When Justin begins digging into Tessa’s activities, he learns that the people in line for the treatment have signed informed consent waivers but don’t really know what that means. Word quickly reaches the home office that he’s asking questions about things that aren’t his business, and he is placed on indefinite leave. Back in London, the officials retain his passport. Using a fake passport, he travels to Berlin and Kenya, where he learns that Dr. Lorbeer invented Dypraxa while working for KDH Pharmaceuticals in his search for a tuberculosis cure. Dr. Lorbeer quit his job when people began to die while being treated. The company decided to ignore the problem. Dr. Lorbeer has since devoted his life to treating the poor of southern Sudan. The British company KDH Pharmaceuticals continued to manufacture Dypraxa in its original overly toxic formulation because it would cost too much money and take too long to change it. If they didn’t market it now, they would no longer have a competitive edge. Instead, KDH bribed the Kenyan government to allow them to contract
120
ReAction! Chemistry in the Movies
with the Threebees Company to test it there. Threebees proved the drug was effective against tuberculosis and that it was possible to determine who would be most susceptible to its toxicity. Mysteriously, it remained deadly toxic to some people. So, Threebees suppresses knowledge of the deaths by bribing hospital officials to erase any pertinent records and to bury the dead in a mass unmarked grave. Commentary: Rachel Weisz’s performance earned her the 2005 Best Supporting Actress award from Academy of Motion Picture Arts and Science, Golden Globe, and Screen Actors Guild. Clinical trials are scientific studies in which humans willingly help determine whether a new therapy or diagnostic procedure is safe and effective. To participate, individuals must sign “informed consent” agreements indicating they understand and accept the potential benefits and side effects of the new drug or procedure. They also waive their rights in the event that something goes wrong, but they retain the right to withdraw from the study at any time. The researchers are ethically and legally responsible for explaining the purpose of the trial and its possible risks and benefits in ways that are understandable to the participants. The movie supposes that the cooperating Kenyans do not understand they might die as a result of their participation in this clinical trial of a tuberculosis treatment. In real life, very few people in any country would ever be part of a drug trial that caused the deaths of significant numbers of test animals. This would have been explained to the participants before they signed the informed consent waiver. Bhopal Express (2001) Distribution company: Phaedra Cinema, India Director: Mahesh Mathai Screenwriters: Piyush Pandey and Prasson Pandey Short summary: Events on the night of the Union Carbide factory explosion that released methyl isocyanate and hydrogen cyanide gases into the city of Bhopal, India Plot description: The movie begins from the point of view of the train driver, cuts to the side of the train, and fi nally to the front of the train as it moves toward us. The horn blows. The engineer sees a man run across the tracks desperately waving a white cloth. We soon learn that the man is Babu (Kay Kay), whose bloodied face is now calm because the train has stopped. As a second train passes, Babu sees that it is the Bhopal Express and his face contorts in terror because his wife Tara (Nethra Raghuraman) is on that train and it is racing to the station. The rest of the movie is told as a long flashback. As the credits roll, we see shots of the city of Bhopal as the sun rises. When the credits end, Tara and Babu wake to have breakfast and engage in loving banter. After Babu leaves for work, Tara performs a ritual in which she asks for long life for her husband. He is a supervising engineer
Bad Company
121
at the Union Carbide plant, which is only five years old but nonfunctioning and falling apart. Autorickshaw driver friend Bashir Miyam (Naseeruddin Shah) appears at the house to drive Tara to the train station. She’ll be gone to her mother’s for fifteen days as part of the ritual. Babu and Bashir spend the night drinking. The scenes alternate showing rumbling factory pipes and the two men enjoying themselves at the bar. Another Union Carbide worker at the bar says, “One day all of us will be killed by gas” and the pipes explode. People begin waking from the acrid smoke and one cries out that someone is cooking chilies. Soon there is a stampede and people yelling “Gas!” Babu drives some people to the hospital. Bashir helps people get cloth for facemasks. At the hospital, a doctor has called the Union Carbide plant to ask about the antidote and tells the others, “We’ll know about the medicine soon.” The factory doctor explains on the phone that the gas is just like tear gas. There is nothing to worry about, just wash their eyes with water. Commentary: The Bhopal insecticide factory was built in 1979 to tap into India’s growing economy. The plant synthesized tons of the insecticide methomyl (sold under the trade name Lannate) (figure 4.4). Methomyl is a broad-spectrum carbamate-type insecticide that was patented by Du Pont and has been on the market since 1966 (International Programme on Chemical Safety 1995). Its many favorable properties include degrading rapidly in water, where it has an environmental halflife of only 2–3 days. In dry soil, its half-life is 11–30 days. It is a white powder that a farmer would dissolve in water and spray onto crops. Like many insecticides, it is a strong neurotoxin for insects and a very weak one for animals. It is neither mutagenic nor teratogenic for animals. Droughts and floods reduced farm cash flow in India for five years in a row, and insecticide sales remained well below the early projections. Union Carbide shut down the plant for economic reasons, keeping a small crew on site to slowly dismantle the operation. With dire consequences for the people of Bhopal, the alarms were turned off because there were no longer any manufacturing workers on site, just a small maintenance crew. By the night of the deadly gas leak on December 2, 1984, more than 60 tons of deadly methyl isocyanate, a precursor to methomyl, was still being stored in three fail-safe storage tanks.
Carbamate part is contributed by O Methylisocyanate
H3C C H3C
S
N
O
C
NH CH 3
Methomyl (Lannate®) Figure 4.4. Methyl isocyanate is used to make the insecticide methomyl.
122
ReAction! Chemistry in the Movies
CH3 N
C
O
Methylisocyanate
+ H2O
CH3
NH2
Methylamine
+
CO2
Carbon Dioxide
Figure 4.5. One of many reactions that may have occurred when the water reacted with the methyl isocyanate.
On the night of the catastrophic gas leak, an employee was flushing water through a corroded pipe when multiple stopcocks failed and allowed the water to enter the largest methylisocyanate tank (figure 4.5). The resulting reaction was so powerful that the tank blew out of its concrete bunker and spewed 27 tons of a combination of methylisocyanate, hydrogen cyanide, and methylamine into the atmosphere (Lapierre and Moro 2002), although the exact composition is still unknown. The wind blew the gas south into the city but was strongest at the adjacent rail station. As many as 500,000 citizens were exposed and 3,800 died immediately. There are estimates that tens of thousands of survivors continue to suffer from ailments caused by the accident, including blindness, breathing problems, and reproductive problems. After the tragedy, the Indian Council of Medical Research, based in New Delhi, began monitoring more than 80,000 victims and more than 15,000 unaffected Bhopal residents. Unfortunately, they stopped their study in 1994 for reasons that were never publicly stated. After 10 years of legal and government wrangling, the council fi nally released a preliminary report of their pre-1994 fi ndings (Crabb 2004). They noted that the rates of death, miscarriages, and general morbidity were higher in areas exposed to the gas and that the most common long-term complications involved the eyes and lungs. In 1989, Union Carbide reached a settlement with the Indian government following approval by the Supreme Court of India (Union Carbide 2001). They paid $470 million in compensation, which gave $300–500 to each victim, enough to cover about five years’ worth of medical expenses. When Union Carbide sold its entire stake in its Bhopal plant to Eveready Industries in 1994, the $20 million proceeds were placed in a trust that funded the Bhopal Memorial Hospital and Research Centre to provide specialist care to the victims. In 1998, the state government of Madhya Pradesh assumed control of the site and cleanup efforts because Eveready was not making progress. Union Carbide folded as a company, and its assets were purchased by Dow Corporation in 2001. Erin Brockovich (2000) Distribution company: Universal Pictures and Columbia Pictures Director: Steven Soderbergh Screenwriter: Susannah Grant
Bad Company
123
Short summary: Amateur lawyer Erin Brockovich develops chromium case against Pacific Gas & Electric (PG&E) and wins MPAA rating: R Plot description: Erin Brockovich (Julia Roberts, who earned the Academy Award for Best Actress for her performance in this film) is an unemployed single mother with three children who gets a job as a file clerk in a law office. Her employer is attorney Edward L. Masry (Albert Finney). He had just lost her personal injury lawsuit against a doctor, who she claimed injured her in a car accident. While working with the files, Brockovich becomes interested in a pro bono case against PG&E. Masry allows Brockovich to carry out some research into the case since it was on the back burner. She learns that a number of people living in Hinkley, California, suffer from a variety of debilitating ailments ranging from nosebleeds to gastrointestinal problems to uterine cancer. A UCLA toxicologist tells her that hexavalent chromium is a potent toxin that can cause cancer and the other problems. Brockovich gets really angry when she discovers that agents from Hinkley’s PG&E plant had visited the townspeople to tell them hexavalent chromium had leaked into the water supply but that chromium was actually good for them. More digging reveals that PG&E used hexavalent chromium to prevent corrosion in their cooling towers but when it was used up it was placed in unlined pools, where it leached into the groundwater. After a great deal of time spent by Brockovich and money spent by Masry, PG&E decides to settle out of court for $333 million in 1996, the largest ever settlement in the United States. After Masry moves into his new office, we see his face on the cover of Los Angeles Lawyer magazine with the tag “Goliath Beware.” Commentary: Few nonchemists will know that hexavalent chromium refers to its oxidation state, but they will probably be swayed by this movie to believe it is capable of causing a wide variety of ailments. Its common species are chromium(VI) trioxide (CrO3), chromic acid (H2CrO4), bichromate (HCrO4 –), chromate ion (CrO42–), and dichromate ion (Cr2O72–). Chromate is commonly used as a bright yellow pigment and dichromate as a bright orange pigment. Chromium(VI) trioxide and chromic acid are strong oxidants that have useful anticorrosion properties. When other metals are treated with them, a thin coat of metal oxide forms on the surface, preventing the slower and more corrosive oxidation during prolonged use. During the surface oxidation process, the chromium(VI) is reduced to the blue-green–colored hexahydrated chromium(III) ion, written as Cr(H2O)63+. Green pools of chromium solution were discussed in the movie while showing an aerial photo of the PG&E site, indicating that it was the reduced form of chromium that formed the majority of products in the pools. As an aside, this trivalent chromium is an essential trace element that appears to help the body to metabolize sugars efficiently. Soil
124
ReAction! Chemistry in the Movies
consists of silicon dioxide crystal fragments called silicates that have a net negative charge. Silicates bind divalent and trivalent cations very strongly, and the chromium(III) ion would certainly not leach through the soil. The problem for PG&E was that its chromium waste also contained small amounts of unreacted chromate, bichromate, and dichromate. These are anionic, do not bind to the silicates, and would wash readily through the soil and into the groundwater, provided the pools were not lined. The U.S. EPA sets the maximum contaminant inhalation level for dry chromium(VI) at 0.1 parts per million (equal to 0.00001%), indicating that it is among the most toxic simple ions. This level was set much lower than the amounts that caused a number of chromate dye workers to succumb to a particular type of lung cancer due to chronic inhalation of dry powders. There are many gaps in understanding the toxicology of this effect, but several things are clear (O’Brien et al. 2003): Chromate is very water soluble and readily taken into cells by a nonspecific anion transporter; chromate plus common intracellular reducing agents such as glutathione, ascorbate, or cysteine are capable of damaging DNA, and this can result in DNA damage such as chromium adducts, DNA oxidation, and DNA cross-linking. Even though the movie narrative catalogs numerous health complaints among the Hinkley citizens and explains that PG&E knew it had contaminated the drinking water with hexavalent chromium, it remains unclear even today whether or how the ingested chromium could cause the reported health problems. The scientific questions that were not addressed in the movie are (1) whether the Hinkley citizens suffered illness at higher than normal rates, (2) whether it is possible for so many different diseases to have the same cause, and (3) whether there is supporting evidence that chromate can cause the specific diseases (Kolata 2000). There is currently no data linking hexavalent chromium in drinking water to any diseases. The movie indicates that PG&E settled out of court to avoid the negative publicity of having to defend the suit. Kids in the Hall: Brain Candy (1996) Distribution company: Paramount Pictures Director: Kelly Makin Screenwriters: Kevin McDonald, Scott Thompson, Mark McKinney, Bruce McCulloch, and Norm Hiscock Short summary: Roritor Pharmaceutical chemist Chris Connor develops antidepressant Gleemonex and company releases it before sufficient testing MPAA rating: R Plot description: To prepare us for what we are about to see, the opening credits contain numerous mathematical equations, chemical equations, atoms, and three-dimensional molecules. When the movie begins, the
Bad Company
125
camera drops deep beneath Earth’s surface to the “Depression Project” at Roritor Pharmaceuticals. At 8:00, scientists in white lab coats are working and eating food in the lab, a clear violation of the safety code. There is one caged spider monkey surrounded by bubbling chemical apparatus, another safety violation. Suddenly, lightning flashes and thunder booms as the camera focuses on a researcher who writes chemical formulas on a chalkboard. To complete the formula, he writes the word “happy” and drops his chalk in astonishment. The camera pulls out to a wide shot of the blackboard full of chemical equations. Don Roritor (Bruce McCulloch) founded the company when he developed Stummies, an antacid. After a researcher reports to the corporate board that Stummies has side effects, Roritor says, “That’s all right, as long as there aren’t any fl ipper babies,” to which the researcher says that there have been only a few. The researcher is forcefully carried away, presumably never to be seen again. Pharmaceutical chemist Chris Connor (Kevin McDonald) is called in to speak to the board next because his group has developed a powerful antidepressant called Gleemonex. The company decides to release the drug despite limited testing, although they do change its color from blue to orange. In actuality, the researchers had tested it on one middle-aged female called patient 57 (Scott Thompson), later identified as Mrs. Hurdicure. She was in a deep depression when she took the blue pill, which entered her stomach, turned its contents blue, and then electrified her brain so that she recalled a happy Christmas moment. It was actually a perfunctory visit from her son, but she smiles broadly in contentment. After “Elizabeth, Queen of England” signs an application form for the drug with its proper name listed as Duonoflouriximinimum 602, orange pills roll off an assembly line while happy music plays and trucks are loaded with boxes. Dr. Cooper makes an appearance on a Nina’s TV talk show, where the host asks the audience, “Doesn’t anyone want to know how the drug works chemically?” When she receives no response, she asks Dr. Connor to stand and wiggle his hips, which he does to thunderous applause. Soon, Gleemonex is outselling penicillin. Commentary: This edgy comedy makes many college-level chemical jokes about the development and marketing of an antidepressant. Even though the joke about “fl ipper babies” goes beyond the pale, it reveals that this Canadian comedy troupe was referencing the thalidomide disaster (briefly described in chapter 3). Safe (1995) Distribution company: Sony Pictures Classics Director: Todd Haynes Screenwriter: Todd Haynes Short summary: Carol White acquires multiple chemical sensitivity MPAA rating: R
126
ReAction! Chemistry in the Movies
Plot description: Carol White (Julianne Moore) has just married a young widower with a son and is moving into their house in San Fernando Valley, California. She and a friend have lunch but don’t really communicate. She exercises regularly but not enthusiastically. When she visits her doctor about her malaise, his only suggestion is to quit her milk and fruit diet. The next day she suffers a nosebleed while getting her hair permanently waved, but her unsympathetic doctor tells her she’s actually healthier than before. He recommends she visit a psychiatrist, which she does but to no effect. After a few more debilitating episodes, White and her husband join a group to watch a video of a man describing various symptoms. He concludes that if you have these, then you have environmental illness. Of the 60,000 chemicals in the environment, only 10% have been tested for toxicity. In the 3.75-minute scene from 53:00, White has an immunotoxicity test. Her only response is to 0.02% milk. Next, she attends a seminar about coping and listens to other people tell their stories. After more encounters with chemical fumes, White happens upon an infomercial for the Albuquerque Center for Chemical Detoxication. She enrolls to learn that detoxification is one part of the treatment and that learning to love herself is the other part. At the center, she complains about the smell of the fumes from the highway located miles away and is moved into a porcelain-lined igloo. The movie ends with White in her igloo looking into the mirror and saying “I love you.” Commentary: Multiple chemical sensitivity (MCS) is a mysterious disease that is physical and psychological. It is ironic and disturbing that Carol White is allergic to nothing except milk because it had been her primary nutritional source. In fact, many people with MCS show high milk toxicity, which could be related to their hypersensitivity because milk is used in so many products. When she has nosebleeds and fainting spells after encounters with permanent wave chemicals and pesticides at the dry cleaner’s, you wonder whether she is having an exaggerated psychosomatic response. In the movie’s narrative, though, White doesn’t seem intellectually engaged enough to have developed a negative gut response to such things. Instead, she seems to be someone who is particularly sensitive and is forced to learn why but can’t fi nd easy answers. Her response is also metaphorical in that her body forces her to take charge of her life. It is a credit to writer/director Todd Haynes that the cause of White’s MCS is unclarified. His interest lies in reticent Carol White’s response to her malady. It is particularly telling that she seems to admit self-love for the fi rst time only when she is alone in her porcelain igloo staring into the mirror at the end of the movie. One can see a poignant echo of the Jekyll and Hyde mirror scenes. The chemically constructed world she inhabits has left her divided from herself. Fractured and isolated, she must begin the process of making herself whole.
Bad Company
127
I Love Trouble (1994) Production company: Touchstone Pictures Director: Charles Shyer Screenwriters: Nancy Myers and Charles Shyer Short summary: Rival news reporters Peter Brackett and Sabrina Peterson compete for a story about a corrupt chemical company producing a carcinogen MPAA rating: PG
Plot description: Journalists Peter Brackett (Nick Nolte) from the Chicago Chronicle and Sabrina Peterson (Julia Roberts) from the Chicago Globe compete to cover the story of a train crash that killed 300 people. As they sleuth out the story and fall in love, they discover Chess Chemical Company is involved. Chess Chemical manufactured napalm and Agent Orange during the Vietnam War. When that war ended, the company entered the genetic revolution in the form of a Livestock Development Factor (LDF) milk production project. Brackett and Peterson learn that Drs. Gerald Beekman and Alexander Horby are somehow central to the story and were the codevelopers of LDF. Unfortunately, Beekman died one week after retiring, and Horby suffered a stroke, rendering him unable to communicate. Before Beekman died, however, he sent a pen to his son, a science teacher, that held microfi lm with incriminating evidence about the LDF project. While the son was returning home from his father’s funeral on the train, it crashed and he died along with 300 other innocent people. LDF could be the biggest moneymaker since NutraSweet except that it causes cancer in lab rats. In the short confrontation scene at 1:55:00, the company president, Wilson “Will” Chess (Dan Butler), exclaims, “I mortgaged this company to develop this product over ten years. So, it had to fly. Now that it has problems, we can’t just take it off the market because some rats died.” Commentary: Chess Chemical Company appears to be based on some combination of Monsanto and Dow, two of the largest chemical companies. For instance, Monsanto developed recombinant bovine somatotropin (BST) for use in milk production, Dow developed and supplied napalm-B to the U.S. government during the Vietnam War, and both supplied Agent Orange. In 2000, however, after this movie was released, Monsanto merged with and changed its name to Pharmacia. Two years later, the agricultural division was spun off to create the new Monsanto, with headquarters near St. Louis, Missouri. It is now an international giant among agriculture companies because it adds desirable traits using molecular biology. After a cow gives birth, it naturally produces BST, a polypeptide hormone, to stimulate production of its milk. In the 1980s, the Monsanto Chemical Company discovered that synthetic (or recombinant) BST
128
ReAction! Chemistry in the Movies
could be injected into cows to increase their milk production artificially. In 1993, the year before this movie was released, the U.S. FDA ruled that milk produced in this way was no different from natural milk and approved it for commercial use. Today, about 30% of milk in the United States is produced in this way. BST has not been without controversy and has never been approved for use in either Canada or the European Union. Even though this movie is a romantic comedy, it is more unnerving than humorous when the president is discovered to have killed his two most valued colleagues and a trainload of 300 people in an attempt to suppress some damaging information. Silkwood (1983) Distribution company: 20th Century Fox Director: Mike Nichols Screenwriters: Alice Arlen and Nora Ephron Short summary: Biography of whistleblower Karen Silkwood set in 1974, the year that she died or was killed on her way to meet a journalist about workplace safety violations; based on a true story MPAA rating: R Plot description: Karen Silkwood (Meryl Streep) processes plutonium at the Kerr-McGee Nuclear Plant in Oklahoma in the 1970s. A guide leads a tour through the facility and describes the process by which plutonium is obtained from uranium oxide. He tells the visitors, “The rays can’t hurt you if you’re careful.” The next scenes provide myriad examples of lax safety practice. Silkwood chews gum while working. At lunch, there is hushed discussion about a contaminated truck, and later Silkwood sees welders cutting it to pieces. Everyone ignores the safety test alarm because they never practice the procedures. One of the worker’s suits had a leak so she was exposed to 24 dpm (disintegrations per minute). The other workers don’t think that’s “too bad.” Silkwood becomes concerned about the possibility that this job could give her cancer and starts reading plutonium literature provided by the worker’s union. She is transferred to metallography, where she learns to cut sections, grind them smooth, and take a photograph. She watches her supervisor use a pen to touch up the imperfections. He says it is his job to make sure they are “correct.” Silkwood becomes convinced that the management hasn’t properly informed the workers about the dangers inherent in their jobs, so she volunteers to help with the union’s negotiating committee. In Washington, D.C., she meets with the commissioner of the union and the Atomic Energy Commission. After citing many issues, she casually mentions that her supervisor touched up the metallography negatives. The commissioner tells her that if you put defective rods in a breeder, you could melt down the entire state. He asks her to get proof so he can put her in touch with the New York Times.
Bad Company
129
As Silkwood begins snooping around the office, her coworkers treat her like a pariah because they don’t want to lose their jobs. Her boyfriend quits and urges her to do the same, but she won’t. One day, she walks past a monitor and sets off the alarm. She is brutally scrubbed down and a team of investigators discovers that her nasal sprayer at home has 45,000 dpm, which means that she’s internally contaminated. She is flown to Los Angeles, where the physician detects 6 nanocuries of americium, a breakdown product of plutonium. He reassures her that this is well under the 40 nCi ± 300%, the maximum permissible. After she returns home, she arranges to meet with a reporter from the New York Times. We learn in the afterword that Silkwood died when her car crashed under mysterious circumstances on the way to meet the journalist. Commentary: Streep gives a powerful performance as a worker who progresses from being just another member of the crew to an activist whistleblower. Americium-241 is a decay product of Pu-241, a contaminant produced when bombarding U-238 with neutrons to generate Pu-239. As noted in the commentary to Fat Man and Little Boy (1989) (see chapter 3), Pu-241 decays rapidly to Am-241, which is very long-lived and emits strong gamma rays. As the plutonium ages, the amount of Am-241 accumulates, making the material increasingly dangerous to handle. The Incredible Shrinking Woman (1981) Production company: Universal Pictures Director: Joel Schumacher Screenwriter: Jane Wagner, based on Richard Matheson’s 1956 novel and 1957 movie The Incredible Shrinking Man Short summary: Suburban housewife Pat Kramer starts shrinking after exposure to a mixture of “Sexpot” perfume and a new detergent Plot description: Pat Kramer (Lily Tomlin) is a happy suburban housewife taking care of her daughter, son, and a dog. She brings home all sorts of household products from the store, buys others from her friendly neighbor Judith Beasley (Lily Tomlin in a second role), and receives prototypes from her television-advertising executive husband Vance (Charles Grodin). In the 10-minute scene beginning at 4:30, Pat is driving home from the grocery store with her children when the dog passes gas and they all gag. Her daughter grabs a can of room deodorant but knocks off the nozzle to fi ll the car with fumes. They open the car windows for relief but a passing van backfi res fumes into their car. When Vance arrives home, he gives Pat a perfume sample called “Sexpot” that he accidentally spills on her blouse. She uses some new detergent to clean it up and lays the blouse on the counter, where it smolders as they make love. She wakes the next morning to fi nd that she’s a bit smaller. As the days pass, she continues to shrink. Once the word gets out, she becomes a minor celebrity but also less able to care for her
130
ReAction! Chemistry in the Movies
husband and children. When she is only a few inches high, a group led by a mad scientist kidnaps her for nefarious purposes. Commentary: This fi lm is a satire/homage to 1957’s The Incredible Shrinking Man, discussed in chapter 3. In both versions, a combination of unlikely agents is required to cause the miniaturization effect. In this version, the agents are household products such as a feminine hygiene spray called “Breathe Easy.” In fact, dangerous combinations of consumer products do exist. For instance, mixing ammonia and bleach produces a mixture of deadly toxic gases including monochloramine, dichloramine, and chlorine. Soylent Green (1973) Distribution company: Metro-Goldwyn-Mayer Director: Richard Fleischer Screenwriter: Stanley R. Greenberg, loosely based on the 1966 Harry Harrison novel Make Room! Make Room! set in 1999 Short summary: Ecodisaster dystopia about New York City Police Detective Robert Thorn, who investigates the murder of Soylent Corporation director William R. Simonson MPAA rating: PG Plot description: The images behind the credits show industrial plants with stacks belching smoke and spewing waste into the river, followed by polluted countrysides and rivers, and ending with masses of people. Robert Thorn (Charlton Heston) is a New York City police detective who investigates the murder of Soylent Corporation director William R. Simonson (Joseph Cotten). It is the year 2022, when natural foods such as strawberry jam, alcohol, and steak have been absent from the diet for so long that most citizens don’t know what they’ve missed. Even wilted lettuce is a delicacy. Most citizens eat crackers called soylent yellow and soylent red, the composition of which is never given. In 2022, the Soylent Corporation releases soylent green crackers, which they claim is made from plankton, a natural foodstuff. A riot ensues when the market can’t meet demand. While investigating Simonson’s murder, Thorn’s roommate and partner Sol Roth (Edward G. Robinson in his 100th and fi nal movie role) uncovers a conspiracy perpetuated by the Soylent Corporation. Roth is a former college professor and one of a small cadre of aging citizens who still know how to read and write. Commentary: The idea that populations will suffer as they outgrow their food supplies was fi rst put forward in 1798 by economist Robert Thomas Malthus in Essay on the Principle of Population. He concluded that the checks to overpopulation were wars, disease, and famines. In the absence of those undesirable factors, he suggested that people should limit the number of children they bear. As demonstrated by this movie, there are other limits to growth, such as environmental pollution and
Bad Company
131
limited resources. This fi lm doesn’t necessarily put the blame for the pollution on the industries that generated them but rather on the society that allowed them to pollute. Even so, equal blame has to be placed on the industries that ignored the global dangers of the pollution they generated in this fi lm. This was the second ecodisaster fi lm. It followed Silent Running (1971), which postulated a world in which the only remaining plants were in a space station circling Saturn because those on Earth had been destroyed during a nuclear apocalypse. Soylent Green presents a far more gradual and mundane end to the world as we know it—overpopulation coupled with an ecological disaster brought on by overconsumption and industrial pollution. Riders of the Whistling Pines (1949) Distribution company: Republic Pictures Director: John English Screenwriter: Jack Townley Short summary: Gene Autry supervises DDT spraying to prevent tussock moth forest infestation and then stockmen blame him for animal deaths Plot description: Recently retired forester Gene Autry (Gene Autry, who always played a character named Gene Autry) notices a tussock moth infestation as he rides his horse through a forest. In the five-minute scene starting at 18:15, he alerts his forester colleagues about the problem and then agrees to head a DDT-spraying project. First, they quarantine the forest, and then they build an airstrip. As the trucks labeled DDT roll past the county coroner’s office (figure 4.6), the townsfolk on the porch discuss how it might kill other animals, too. The scene ends with crop dusters spraying the forest. Unfortunately for Autry, the Mitchell Lumber Mill is run by some conniving lumbermen who believe they will be given a contract to cut down the infested wood. After we see a different plane fly over the forest repeatedly, a veterinarian announces that he just examined 20 dead animals. He says that Autry must have miscalculated the amount of DDT. Autry responds that it was tested and asks the veterinarian not to spread the information. Before he can do anything, one of the lumbermen spreads the rumor that the DDT killed the stock. Commentary: While making a rodeo appearance in Idaho in 1947, Autry learned that the local foresters were using planes to spray the forests in an attempt to eradicate a tussock moth infestation (Magers 2007). After he returned to Hollywood, he had screenwriter Jack Townley develop the story, and this movie was the result. Just after the war, DDT was being used for the fi rst time to control gypsy moths. There were virtually no reports that DDT was harmful. Nevertheless, when audiences
132
ReAction! Chemistry in the Movies
Figure 4.6. This screen grab shows townsfolk watching DDT trucks drive by the Coroner’s office. RIDERS OF THE WHISTLING PINES Image reprinted with permission from Autry Qualified Interest Trust © 1949, 2005.
today hear Autry say “It’s been tested,” they laugh because he doesn’t say whether the outcome of the test was good or bad. Despite the retrospectively amusing “DDT as hero” theme in this movie, Autry National Center curator Michael Duchemin noted that Autry was about 20 years ahead of his time in producing a story with an environmental theme (Duchemin 2007). In his introduction to a screening of this fi lm in 2007 at the Center, Duchemin also noted that this fi lm was released one year after Muller won the Nobel Prize in Medicine for his discovery of DDT. I Know Where I’m Going! (1947) Distribution company: General Film Distributors (United Kingdom) Director: Michael Powell and Emeric Pressburger Screenwriters: Michael Powell and Emeric Pressburger Short summary: Joan Webster travels to the Hebrides to marry the president of Consolidated Chemical Industries Plot description: In the five-minute scene beginning at 4:45, Joan Webster (Wendy Hiller) tells her father that she is about to marry the president
Bad Company
133
of CCI (Consolidated Chemical Industries). Her fiancé is Lord Robert Bellenger (whom we hear over the phone but never see), one of the richest men in England. He is living on the Island of Kiloran in the Scottish Hebrides, “where the war is a million miles away.” She will travel there according to a precise schedule that he laid out for her, and they will marry in a small ceremony. While sleeping on the train that night, she dreams that she marries “Consolidated Chemical Industries.” When Webster arrives at Port Erraig, there is no boat to take her to the small isle. The next day a fog sets in that prevents the journey, so she wishes for wind. A gale force wind prevents her journey on the next day. “Foster” (Roger Livesey) is on a brief leave from service in the war and also wants to reach Kiloran. He immediately falls in love with Webster and spends the rest of the movie subtly trying to make her fall in love with him. He helps her fi nd a room each night. This becomes awkward when she learns that he is actually Lord MacNeil of Kiloran, who has rented the isle to her fiancé Bellenger for the duration of the war. Commentary: Even though played for humor, the unseen but much discussed Lord Bellenger encapsulates so many distasteful features about chemical companies that this fi lm could have been used as the archetype fi lm for this chapter. The very rich Bellenger is getting richer by supplying chemical needs during a war that he is carefully avoiding. The movie also makes it clear that Bellenger doesn’t know how to live an authentic life, despite his money, showing that he is bankrupt on that score, as well.
5 A Master/Slave Narrative Drug Addiction and Psychoactives
EUPHORIA AND ADDICTION: DOPAMINE AND BRAIN CHEMISTRY All addictive chemicals are psychoactive, but not all psychoactive compounds are addictive although they can be abused. Notable nonaddictive psychoactive compounds are hallucinogens and the many mood-altering drugs marketed for depression, anxiety, epilepsy, and so forth. This chapter’s movies (table 5.1) were chosen for the variety of psychoactive substances used by characters. They all show either drug abuse or addiction. The fi rst few times that someone chooses to take a potentially addictive drug, they experience a rapid-onset pleasurable response followed by a slower onset, longer lasting dark side called withdrawal (Koob and Le Moal 2006; Grens 2007). Chronic use and bingeing lead to “tolerance” such that the euphoric effects are diminished even as the dose is increased. A person is said to be addicted to a drug when he or she seeks out the drug to avoid the dark side more than to induce its bright side. The bright and dark emotional and biochemical responses of euphoria, tolerance, and withdrawal are associated with a set of nerves called the “reward system” that lie at the top of the brainstem, buried deep within the human brain. We inherited these neurons from our earliest vertebrate ancestors. They are normally stimulated when we quench our thirst, stave off hunger, engage in sexual behavior, or participate in a host or other pleasurable activities. Repeated use of addictive drugs triggers the synthesis of proteins in the brain that cause anxiety or depression. Therefore, one promising line of research is to lessen the effects of withdrawal by fi nding other small molecules (therapeutic drugs) that bind to the receptors for the anxiety-producing or depression-inducing compounds. The “big three” legally addictive substances are nicotine, alcohol, and caffeine. About three-quarters (76%) of Americans over the age of 12 have smoked at least one cigarette in their lifetime, and 19% of Americans over the age of 12 smoke every day. From this we can calculate that 25% of users are addicted (=19%/76%), the highest addiction rate 134
A Master/Slave Narrative
135
Table 5.1. Movies about drug addiction and abuse Title (Year)
Psychoactive Substance
Thank You for Smoking (2005) The Salton Sea (2002) Formula 51 (2001) Trainspotting (1995) Altered States (1980) All That Jazz (1979) THX-1138 (1970) Easy Rider (1969) The Valley of the Dolls (1967) The Trip (1967) Seconds (1966)
Nicotine Methamphetamine POS 51 (MDMA) Heroin Atropine-like Amphetamine SP5 etrazene Marijuana (and LSD) Dexedrine and seconal LSD and Thorazine Pentothal and alcohol
Figure 5.1. The addictiveness of five psychoactive substances.
for any substance (figure 5.1). With regard to alcohol, 51% of Americans are regular, moderate drinkers, 23% are binge drinkers, and 7% are heavy drinkers. Nicotine and caffeine don’t cause substantial impairment during consumption but they do lead to anxiety and depression during withdrawal. For the most part, the health problems associated with the big three are related to their nonaddictive effects, such as lung cancer for cigarettes and cirrhosis of the liver for alcoholism, rather than to their psychoactive effects. The “little three” illegal addictive substances are heroin, cocaine, and marijuana. Even though only 2% of the population is addicted to
136
ReAction! Chemistry in the Movies OH
HO
NH2
OH
HO
NH2
H N
HO
CH3 HO
HO
Dopamine
Norepinephrine
HO
Epinephrine (or Adrenaline)
O H N
NH2 O
CH3
N O CH3
CH3 O
Cocaine
CH3
Amphetamine
CH3
Methamphetamine
Figure 5.2. Cocaine, amphetamine, and methamphetamine prevent the reuptake of dopamine and epinephrine in brain synapses so that the individual feels happier and more energetic, respectively. These compounds share significant chemical structural features.
these drugs, the percentage of occasional users who become addicted to them is high: 23% for heroin, 17% for cocaine, and 9% for marijuana. In addition, there is currently an epidemic of nonmedical oxycodone (sold as OxyContin) abuse in the United States that could easily overtake the little three in importance. It is difficult to predict who will become addicted to any given drug because there are so many factors at play. Initiation is likely to be due to access to the drug, temperament (associated with a strong genetic component), social development, frequent depressed moods (associated with strong genetic and life experience components), and protective factors (parental and social guidance), whereas abuse and addiction are most closely associated with genetic predisposition. As mentioned above, the short-term euphoric effects are centered on the reward system. This is composed of a group of neurons that travel from the ventral tegmental area (VTA) to the nucleus accumbans (NAc), where dopamine is rapidly released from the neurons to give a pleasurable feeling. Each drug raises the dopamine level in a different way. For instance, cocaine, amphetamine, and methamphetamine (figure 5.2) slow dopamine readsorption within the NAc, whereas morphine and other opioids stimulate initiation of many reward nerve impulses in the VTA. When the neurotransmitter dopamine binds to its receptor in the NAc, it initiates a cascade of biochemical events that results in the very long term elevation in the concentration of corticotropin-releasing factor (CRF) and neuropeptide Y (NPY) (Heilig et al. 1994; Funk and Koob 2007; Heilig and Koob 2007). The 41–amino-acid neuropeptide CRF leads to increased anxiety, while the 36–amino-acid neuropeptide NPY leads to lowered anxiety. When alcohol-dependent rats are treated with a small molecule that binds to the CRF receptor, they drink less during withdrawal. The same response was observed in rats with artificially
A Master/Slave Narrative
137
elevated levels of NPY. The search has already begun for a compound that is effective in humans for use in helping addicts during withdrawal.
ALBERT HOFMANN, HIS PROBLEM CHILD LSD, AND THE SEROTONIN RECEPTOR Pharmaceutical chemist Dr. Albert Hofmann discovered the compound named LSD during a systematic search for therapeutic compounds derived from the ergot fungus (Claviceps purpurea, or Secale cornutum) that grows on rye but discovered its hallucinogenic effects by accident (Hofmann 1980). Since 1929, he had been working with natural products at Sandoz in Basel, Switzerland, under the direction of Arthur Stoll. By 1938, Hofmann began his independent line of research. He saw that researchers in the United Kingdom and United States were beginning to characterize the structures of new compounds from ergot, and he didn’t want his company to lose its historical lead in ergot natural products. Notably, Stoll had isolated ergotamine in 1918 (see chapter 1). Hofmann’s work would eventually show that ergot contains a toolbox of therapeutic compounds. In popular culture, though, he is remembered for having discovered the world’s most hallucinogenic substance: d-lysergic acid diethylamide, or LSD (figure 5.3). Hofmann was soon able to devise an efficient semisynthetic route for an ergot alkaloid with strong effects on blood pressure and vasodilation. It was used to reduce the bleeding following childbirth and was marketed as ergobasine. Next, he prepared small amounts of several dozen amide derivatives of lysergic acid (lysergsäure in German) using Curtius’s synthesis. None of them had pharmacological activity that exceeded those already in use, so they were catalogued and shelved. Five years later, Hofmann had a hunch that the 25th compound, a diethylamide (LSD-25), must have had some special pharmacological properties because of its especially small modification. As he fi nished preparing a new batch, he began to feel dizzy so he left work early. At home, he entered into a dreamlike state where objects changed their shape, sounds turned into sights, and everything was bathed in a kaleidoscopic array of colors. Two hours later, his “trip” ended and he ate heartily. To determine whether the LSD was responsible, he knew he had to self-experiment because rats can’t tell you whether they are hallucinating. Three days later, he thought he was erring on the side of caution by taking 250 micrograms in the presence of an assistant. The effects came on so quickly that he felt overwhelmed by the trip this time. His assistant recorded no unusual changes to his blood pressure or any other vital signs despite Hofmann’s inability to communicate during his trip. Only as the LSD concentration in the bloodstream ebbed did he calm enough to enjoy the kaleidoscopic colors and shapes. When he woke the next morning, he was slightly tired but found that his breakfast tasted especially delicious.
138
ReAction! Chemistry in the Movies CH3 O
N
CH3
C
O
OH NH2
HN
CH3
CH3 N
N
CH3
CH3
HN
HN
5-MeO-DMT D-Lysergic Acid Diethylamide (5-Methoxy-Dimethyltryptamine) semi-synthetic hallucinogen from psychoactive toad
Serotonin brain neurotramsmitter
NH2
CH3
CH3 OPO3H-
H3C
O
N CH3
H3C
HN
Psilocybin from psilocybin mushrooms
NH
H3C O O
H2C O
O CH3
Mescaline from peyote mushrooms
MDMA (or Ecstacy) semi-synthetic hallucinogen
Figure 5.3. Many hallucinogens have structures resembling serotonin and bind strongly to the serotonin receptor.
Between 1943 and 1963, Sandoz provided LSD free of charge to researchers who sought to discover a therapeutic use. Psychologists, psychiatrists, and other interested professionals tested its effects on schizophrenia, depression, and alcoholism, but it was never quite successful for any of these. Instead, its hallucinogenic effects helped fuel the growing interest in the nature of reality, consciousness, and perception. For instance, in 1954, Aldous Huxley wrote The Doors of Perception about his self-experiments with mescaline, the psychoactive compound from peyote cactus. In that short book, Huxley postulated that the human brain fi lters reality to avoid being overwhelmed and that hallucinogens remove that fi lter to open the doors of pure perception. As an aside, the psychedelic rock band The Doors took their name from Huxley’s book. In The Trip (1967), the most successful independent fi lm made to that date, Paul Groves (Peter Fonda) wants to get in touch with his feelings so he decides to try LSD under the guidance of John (Bruce Dern). John explains that each pill contains 250 micrograms LSD and that Thorazine (figure 5.4) is available as an antidote in case he has a bad trip. Thorazine is credited with beginning the psychopharmacological revolution that we are still experiencing today (Lopez-Munoz et al. 2005). It is the trade name for chlorpromazine, which was fi rst synthesized
A Master/Slave Narrative
139
CH3 N
CH3 N CH3 N
S
Chlorpromazine (Thorazine®)
N N
Cl N H
S
CH3
Olanzapine
Figure 5.4. The antipsychotic chlorpromazine was marketed as Thorazine in the United States beginning in 1957 and has been superseded by the olanzapine-like compounds, first marketed in 1989.
by Rhône-Poulenc in 1950 and marketed as an antihistamine. Shortly thereafter, French anesthesiologist Henri Laborit began testing it as part of an antishock cocktail in patients being prepared for surgery. After he noticed that the patients were especially calm prior to surgery, he approached psychiatric clinics to test its effects on psychotic patients. It was nothing short of phenomenal in the way it revived so many delusional patients, even some schizophrenics who had been confi ned for decades. By 1957, Thorazine was marketed in the United States as an antipsychotic (Scheff 2007). The dominant theory about psychosis at that time was that too much dopamine was binding to the dopamine receptor. So, Thorazine was hypothesized to bind to dopamine receptors in a way that prevented dopamine from binding but without stimulating the receptors. It has since become clear that Thorazine also binds to serotonin receptors (to reduce anxiety) and to histamine receptors (to calm) for its overall effect. Olanzapine (figure 5.4) was introduced to the market in 1989 as an “atypical” antipsychotic, causing the therapeutic use of Thorazine to decrease. It is atypical in that it was developed to target both the dopamine and serotonin receptors to relieve the depression and cognitive defects that Thorazine didn’t treat. An important study in 2005 revealed that these fi rst- and second-generation antipsychotics are actually equally effective but that the more recent agents have fewer side effects (Lieberman et al. 2005; Weiden 2007). It is likely that elevated dopamine levels are a downstream consequence of some other event. In other words, the search is still on for better antipsychotics and for a better hypothesis regarding the cause of psychotic episodes.
140
ReAction! Chemistry in the Movies
MIXING IT UP As the DJ spins the latest Deep House Techno Trance mix, the dancers are gyrating in groups or pairs or alone. Some people who gather at discotheques take a combination of MDMA (called ecstacy on the street) and a stimulant such as an amphetamine with the intention of dancing the night away while sipping a vodka/energy drink cocktail. If they stay long enough, their euphoria and energy will subside by the time they return home. Just in case this doesn’t happen, one of the group often opts to abstain from the drugs and sticks to the vodka/energy drink cocktails. MDMA is a semisynthetic compound produced by bootleg chemists for sale on the street (Shulgin 1986). MDMA and the amphetamine both elicit euphoria because they have dopamine-like structures (figure 5.2). Despite its structure, MDMA has pharmacological effects closer to the psychedelics, indicating that it must interact with the serotonin receptors (figure 5.3). Users report that MDMA doesn’t induce sensory distortions like LSD but rather an inward self-absorption and clarity of thought (hence its street name ecstacy). An interesting aside is that a mixture of LSD and amphetamine was taken to heighten and prolong the psychedelic experience in the late 1960s, showing that mixing psychedelics with stimulants has a long history in the drug culture.
ADDICTION: SIGNED, SEALED, DELIVERED? In this chapter, the understanding of “chemical experiment” shifts, and the body becomes the surrogate for the lab bench. If it feels good, do it, and do it with chemicals. Mixtures, quantities, and trials are the constants of recreational drug experimentation. The desired outcomes are feelings of pleasure and escape. An alcoholic bacchanal scene illustrates the delight of abandonment in the movie Seconds (1966). In an ironic twist, in Formula 51 (2001) the blue POS 51 pills thrown to a crowd of youths at a dance club are placebos. Despite the deceit, the crowd starts dancing as if they took the real drug. As discussed above, the undesired outcome of many drugs is an addictive effect. Addiction is a diagnostic term for drug dependency that fell out of favor in the past two decades but is now being rediscovered. Its new appeal comes from its fit with current and emerging neurobiological understandings of brain pathology. The term “dependency” is more limited, connoting only the physiological features of tolerance and withdrawal. Addiction embraces a more primary understanding, that of the relationship of master/slave (Sellman 2007). Derived from Roman Law, addiction is defi ned as a formal giving over or delivery by court sentence. It is further understood to mean surrender or dedication of anyone to a master (Oxford English Dictionary online).
A Master/Slave Narrative
141
The addictive effect that drug use has on our old friend Dr. Jekyll is shown in several of the movie versions discussed in chapter 1. After a few drug-induced transformations, he begins to transform into Mr. Hyde without fi rst ingesting his chemical preparation. We see this happen in dramatic variations; he can be sitting on a park bench (1931 version) or sleeping in his bed (1920 version). The movie versions echo the autonomous transformations Jekyll undergoes in Stevenson’s novella. Scholar Daniel Wright counters the traditional literary interpretations that characterize Jekyll as a mere dualistic figure with a split or confl icted personality. He contends that Jekyll is actually a drug addict, presenting clear symptoms of denial toward his dependency, believing instead he can regulate the use and effect of his drug at will (Wright 1994). In Wright’s view, Stevenson has underscored the depth of Jekyll’s ravaged state by showing how he can be overcome by the “terrible potency” of his “fantastic substance” without even ingesting it. This serves to illustrate just how much self-control Jekyll has lost and the magnitude of his “involuntary subservience” to the drug (Wright 1994). The aura of Dr. Jekyll and Mr. Hyde persists. The movies about addictive drugs described in this chapter are often thematically self-centered, their story lines turning on tensions involving a character’s personal identity. Someone taking an addictive drug undergoes an alteration, becoming a noticeably different person. Prefacing his decision to end their troubled marriage, Mel says to his pill-popping wife Neely O’Hara in Valley of the Dolls (1967) at 59:15 min, “I keep remembering the old Neely; she was quite a girl.” In other words, Mel’s current perception of Neely and his memory of her no longer fit together in a reliable way. Does this mean that people can never change? In the physical dimension, our bodies change on a daily basis from infancy to old age. At this moment we no longer have the identical cellular makeup we had only moments ago. In the mental dimension, our experience grows as we come in contact with new knowledge, ideas, and situations. The self can be understood in common terms to mean “the subject of possible experience” (Robinson 2007). To philosopher and psychologist William James, who devoted much consideration to the concept of self, a subject’s experience is just what that subject chooses to pay attention to out of the “stream of consciousness” of life (James 1890). One key point defi ning James’s idea of self is that it will feel intimate and familiar through time. Both attention and selection play active roles in integrating experience into a self that is ours, which we feel we know and possess. Thus, the making of the self can be seen as a creative act. Mary Tod Gray, a nursing professor, fi nds healing potential in William James’s model of self. She contends that nurses who interact with addiction patients can enhance the “therapeutic conversation” if they understand the self as it is experienced by the drug user. For instance, the novelty and intensity of feeling produced in a drug-induced state cause a rupture in the continuity of self-as-usual. While the experience of
142
ReAction! Chemistry in the Movies
such new sensations can spark a connection to a future self, it also can lead to a realization of a sense of fluid boundaries between feeling and changing thought. In Gray’s estimation, this fluidity often produces feelings of insubstantiality and ambiguity that lead to fragmentation and fear. James’s model of self with its emphasis on selective agency of thought can help an addicted individual create a self in which “both intimate feelings and the power of choice can be equally embraced” (Gray 2005). Current brain chemistry research concerning addictive drugs (described above) suggests that concepts of self-control and will power have themselves been surrendered to a growing mental/mechanical split. It is clear that the user exercised choice to begin taking the drug, but what happens to their ability to choose afterward? Satel and Goodwin, psychiatrists and bioethicists, resist a brain-disease model that frames addiction in terms of an involuntary illness because “it leads us down a narrow clinical path.” They fi nd instead more “clinical utility” in thinking of addiction as a primary, though modifiable, behavioral phenomenon (Satel and Goodwin 1998). Addictive behavior is voluntary; the numbers bear this out. Many more drug users do not become addicted than do (see figure 5.1), and most drug users voluntarily quit for economic, health, legal, and personal reasons. More than 40 anticocaine pharmaceuticals have been tested, but none have proved minimally effective. Instead, they blunt general motivation and depress mood (O’Brien 1997). The biggest advantage of nonpharmaceutical treatments is that addicted individuals have a hand in shaping their own self-control.
THE ARCHETYPE MOVIE: THE TRIP (1967) Distribution company: American International Pictures Director: Roger Corman Screenwriter: Jack Nicholson Short summary: Television advertisement director Paul Groves is divorcing wife Sally and takes LSD to get in touch with his feelings Plot description: Before the movie begins, there is a text explanation that LSD is dangerous. Television advertisement director Paul Groves (Peter Fonda) is feeling depressed because he is divorcing his wife Sally (Susan Strasberg). In the scene beginning at 8:45, Groves wants to get in touch with his feelings, so he decides to try LSD under John’s guidance (Bruce Dern). John explains that each pill contains 250 micrograms LSD and that Thorazine is available as an antidote in case he has a bad trip. Groves takes the LSD but soon flees from John out of fear. As he roams around Los Angeles interacting with people, he experiences pulsating psychedelic colors.
A Master/Slave Narrative
143
Commentary: Thorazine is the trade name for the antipsychotic called chlorpromazine, discussed earlier in this chapter. This was the most successful independent fi lm at the box office.
MOVIES EMPHASIZING CHEMICAL ASPECTS OF ADDICTIVE DRUGS Thank You for Smoking (2005) Distribution company: Fox Searchlight Pictures Director: Jason Reitman Screenwriter: Jason Reitman from Christopher Buckley’s 1994 novel Thank You for Smoking Short summary: Academy of Tobacco Studies Vice President Nick Naylor spins support for cigarette smoking MPAA rating: R Plot description: Nick Naylor (Aaron Eckhart) is vice president of the Academy of Tobacco Studies, which is funded by cigarette companies to research the link between health and smoking. As expected, their head of research has yet to fi nd a single link to cancer. Naylor defends cigarettes on a talk show, meets with a movie producer to discuss product placement, and delivers silence money to an advertising icon now dying of lung cancer. Most of Naylor’s actions in the fi lm are in response to a threat by a U.S. senator to introduce a bill that would require cigarette companies to include a large skull and crossbones plus the word POISON on cigarette packs. In a short, disturbing scene at 59:00, Naylor is kidnapped and tied up, his clothes are cut off, and his exposed body parts are covered with nicotine patches. When he wakes up in the hospital, he is being treated for nicotine poisoning. The good news is that smoking saved his life. He survived only because he developed such a high tolerance as a smoker. The bad news is that he can never smoke again because he’s now too sensitized. Naylor says, “That’s OK, I’ve quit before.” Commentary: No cigarettes are smoked on screen throughout the entire movie. The only nicotine is from the weaponized nicotine patches that, of course, are sold as antismoking aids. There are three toxic components among the more than 4,000 chemicals in cigarette smoke: nicotine, carbon monoxide, and tar. Despite this toxicity, most smokers (81%) of cigars, pipes, and cigarettes enjoy smoking. The nicotine in the smoke produces mild euphoria, increases energy, and reduces stress, anxiety, and appetite. It binds to ligand-gated ion channels to initiate nerve signals between motor neurons in the same way as caffeine, scopolamine, and strychnine. This site of action also explains how these compounds act as potent neurotoxins at high doses. The carbon monoxide in smoke is produced by inefficient combustion. At low levels, it may contribute to smoking’s desirable properties, but at
144
ReAction! Chemistry in the Movies
high doses it is a deadly metabolic poison. Carbon monoxide enters the bloodstream to bind strongly to the oxygen-binding site of hemoglobin so that less oxygen is carried throughout the body. At low doses, therefore, it causes mild sedation. Tar is the residue that remains after the cigarette has been smoked and consists of a wide range of polyaromatic compounds, many of which are known carcinogens. Some of the tar enters the lungs, where it either initiates or stimulates the uncontrolled cell growth called lung cancer. Tobacco smoking is the leading avoidable cause of disease in the United States and is responsible for 440,000 deaths every year. There is an average delay of 25–30 years between the time a person starts smoking and when they begin to show symptoms of lung cancer. The Salton Sea (2002) Production company: Castle Rock Entertainment Director: D. J. Caruso Screenwriter: Tony Gayton Short summary: Tom Van Allen enters the meth world as Danny Parker seeking revenge against the masked men who killed his innocent wife MPAA rating: R Plot description: The movie begins with an overhead shot of a man (Val Kilmer) blowing a trumpet while flames lick the ceiling and consume the room’s contents and money. His voice-over says, “My name is Tom Van Allen . . . or Danny Parker. I don’t really know anymore.” The rest of the story is told as a flashback that begins with a 2.25-minute (starting at 2:45) rapid-fi re verbal and visual history of methamphetamine discovery and its modern bootleg production. We are told methedrine was synthesized by the Japanese before WWII and became popular with kamikaze pilots, and soon 2% of the Japanese population was addicted. In the 1950s, it was used as a stimulant by American housewives, and rumors circulated that it was used by President Kennedy. After it became a controlled substance in the 1960s, it was manufactured by bootleggers and “long-haired freaks” for sale on the street. This leads us to a home in which a contemporary “cook” uses drain cleaner, HCl, the red phosphorus from match heads, ether, and cold medicine that he buys at the drug store late at night. As the cook holds up a beakerful of something, the voice-over says, “He screws up from time to time,” and then his house explodes in flames, presumably from the ether. Without revealing too much of the story, Tom Van Allen and his wife stop at a stranger’s home in Salton Sea, California, to make an emergency phone call. While Van Allen uses the bathroom, two hooded men with guns enter the house and kill its residents, who happened to be drug dealers. Van Allen watches through bullet holes in the bathroom wall as the invaders shoot his wife and steal the drugs and money. Seeking revenge, he enters the meth world as Danny Parker.
A Master/Slave Narrative
145
Commentary: Methamphetamine is a stimulant that acts very similarly to cocaine, which is described in detail in chapter 2. Perhaps the biggest difference between the two is that methamphetamine can be synthesized from drugstore components whereas cocaine’s structure is too complex for household synthesis and must be isolated from its natural source. Formula 51 (2001) Production company: Focus Films Director: Ronny Yu Screenwriter: Stel Pavlou Short summary: Los Angeles bootleg chemist Elmo McElroy heads to Liverpool to sell his latest product, the most powerful drug ever created MPAA rating: R Plot description: After Elmo McElroy (Samuel L. Jackson) earned his medical degree in pharmacology from the University of California in 1971, he was busted for smoking marijuana. Flash forward 30 years and he is a bootleg chemist making street drugs for The Lizard (Meatloaf). After he decides to start a new life, he engineers the delayed explosion of a pink solution in a rotary evaporator that is supposed to kill The Lizard but doesn’t. In the meantime, McElroy heads to Liverpool and Manchester in the United Kingdom, where he tries to sell his latest creation. He calls it POS 51 because it is 51 times more powerful than cocaine, 51 times more hallucinogenic than LSD, and 51 times more explosive in the brain than ecstasy. A club owner is interested but asks for proof that the blue pills work. McElroy tosses them to the crowd on the dance floor, who take them and dance with even more energy. Later, in a short scene at 1:33:30, he reveals the compound’s formula (figure 5.5). Commentary: Most viewers will be amused by the trickster McElroy’s doublespeaking admission that POS 51 is a fake, but the chemically informed viewer will simultaneously be amused by the impossible structural formula. The even more chemically knowledgeable will realize that this impossible formula bears no relationship at all to the structures of cocaine, LSD, or ecstacy.
S C
NO2
N HN
⌽
H
NO2 H
NH3
Figure 5.5. The chemical formula for POS 51.The set designer who drew this structure used the shorthand Φ (Greek capital phi) to indicate a phenyl group.
146
ReAction! Chemistry in the Movies
Trainspotting (1995) Production company: Channel Four Films Director: Danny Boyle Screenwriter: John Hodge, from the 1993 novel by Irvine Welsh Short summary: Mark Renton is an Edinburgh heroin addict whose parents force him to quit; after the misery of withdrawal is over, he leaves for London to get away from junkie friends but they fi nd him anyway MPAA rating: R Plot description: The movie opens with Edinburgh heroin junkie Mark Renton (Ewan McGregor) running from his pursuers after he has shoplifted. He tells the viewer to “choose life” and that he chose something else—heroin. The next scenes depict the intimate life of a junkie in a humorously gritty but unromantic way. He shoots up with a group of junkie friends and talks of the euphoria being greater than 1,000 orgasms. He deals with the constipation that accompanies heroin use. He tells us of the reduced sex drive during drug use that returns as the euphoria ebbs. He loathes himself and the low place he believes the Scots have in the world. He steals all the prescription drugs he can get his hands on. His junkie friend Allison’s baby dies from neglect. After he is caught shoplifting, he is placed in a rehabilitation program that involves receiving three daily shots of methadone. When he takes just one more heroin hit for old time’s sake, he enters a coma and is revived at the hospital. His parents decide to force him to quit cold turkey and quarantine him to his bedroom. During withdrawal, he hallucinates about his miserable friends and the dead baby. When it’s over, they have him tested for the AIDS virus. He’s OK, but his friend eventually dies from it. After his girlfriend Diane (Kelly MacDonald) points out that the world is changing and that he needs to change, too, he decides to go to London where his friends can’t fi nd him. They fi nd him, however, and involve him in a big drug deal. Commentary: Heroin was fi rst marketed in 1898 by the Bayer Company as a nonaddictive cough suppressant. It is the doubly acetylated derivative of morphine (figure 5.6). Morphine is the psychoactive and addictive component that comprises 10% of the sticky resin called opium isolated from the flower of the opium poppy, Papaver somniferous. When the 1914 Harrison Narcotics Tax Act banned heroin’s sale because it was too addictive, Bayer developed oxycodone (figure 5.5), which is sold as OxyContin. Today, we know that heroin is the most addictive substance in part because its two acetyl groups make it more hydrophobic so that it passes through the blood–brain barrier more quickly than other opioids. Even so, heroin can’t bind to its brain receptor until at least one of its acetyls is removed, which occurs rather rapidly in bodily tissues.
A Master/Slave Narrative
147
HO
H3C
O O
O
O N
N
CH3
H
O
H
HO
H3C
Morphine active drug
H
O
Heroin inactive prodrug
HO
HO
O
O
N H
CH3
H
N
CH3
O
OH
HO
Oxycodone 150% as active as morphine
H3C
H
CH3 H
O
6-Monoacetylmorphine 100% as active as morphine
Figure 5.6. The structures of some the strongest known opioid analgesics.
The brain receptors that bind heroin are called opioid (opium-like) receptors because they were discovered using the compounds derived from opium. The presence of these receptors implies that humans have natural pain-relieving compounds in their brains, later identified as the enkephalins. They are human brain peptides composed of five amino acids, the fi rst of which is a tyrosine. Tyrosine has a phenol in its structure that resembles the part of the opioids that is buried deepest in the opioid receptor. Even though the enkephalins play an important role in natural pain relief, the opioids fit into the binding site better and are more effective. Altered States (1980) Distribution company: Warner Bros. Director: Ken Russell Screenwriter: Paddy Chayefsky using the pseudonym of Sidney Aaron from his own 1978 novel of the same title Short summary: Consciousness researcher Eddie Jessup engages in selfexperimentation with hallucinogens and a sensory deprivation tank; based on a true story MPAA rating: R
148
ReAction! Chemistry in the Movies
Figure 5.7. Dr. Eddie Jessup (William Hurt) floats in a sensory deprivation tank while under the influence of psychoactive drugs in this image from the VHS box cover. ALTERED STATES © Warner Bros. Inc. All Rights Reserved. Collection of the authors.
Plot description: Eddie Jessup (William Hurt) earns his doctorate in 1967 from Cornell Medical School. He and his anthropologist girlfriend Emily both land professorships at Harvard. Jessup uses a sensory deprivation tank in combination with psychoactive drugs to search for “the inner human, the beginning, the God in humans” (figure 5.7). Jessup soon travels to Mexico to join some elderly men gathered in a cave for a drug-induced but shared psychological experience. Jessup knows that the mushrooms contain belladonna-like alkaloids such as atropine and scopolamine. After he enters the cave, the shaman cuts Jessup’s hand so that his blood falls into the mushroom goulash. After taking a sip, Jessup has a hallucination that resembles a birth experience. Jessup returns to Boston, where his endocrinologist friend Mason
A Master/Slave Narrative
149
synthesizes the active ingredient but doesn’t give us its name. Mason tells him that this stuff heads straight to the brain and is a carcinogen. This does not deter Jessup from trying some. When nothing happens, he increases the dose to toxic levels. Mason offers to determine the half-life of the compound in rat brains before Jessup continues, but Jessup is impatient to get on with the experiment. He and his other colleague ready the deprivation tank. They clean and fi ll it with 10% magnesium sulfate for buoyancy. During one of his recorded sessions, he emerges from the tank with blood on his lips. They have an X-ray taken of his neck and a radiologist says it looks like a gorilla’s. During subsequent sessions, Jessup reverts deeper into his phylogeny until he fi nally becomes protoplasm. Commentary: Chayefsky called Altered States his modern Jekyll and Hyde story (Considine 1994). He based it on the two semiautobiographical, semifictional books written by John C. Lilly, who pioneered the field of sensory deprivation while under the influence of drugs such as LSD and ketamine. In his two books, Dyadic Cyclone (1976) and The Scientist (1978), Lilly describes his reversion to an apelike state during one of his sessions. Atropine and scopolamine have very similar chemical structures but yield different sets of physiological effects. Scopolamine induces a hypnotic state that makes it useful as an antiseasickness drug. It is also used to induce a wakeful state of general anesthesia that has been widely used on women during childbirth. As noted in the Thank You for Smoking commentary, scopolamine and nicotine bind to ligand-gated voltage channels to increase energy and produce mild euphoria. All That Jazz (1979) Distribution company: 20th Century Fox Director: Bob Fosse Screenwriter: Robert Alan Arthur and Bob Fosse, based on Bob Fosse’s life Short summary: Joe Gideon pops pep pills to maintain his creative flow until he meets the angel of death MPAA rating: R Plot description: The movie opens with Joe Gideon (Roy Scheider) taking eye drops, drinking Alka-Selzer water, showering, and popping two Dexedrine pills. He ends by saying, “It’s show time, folks” to the mirror. He works long days planning and choreographing his next show, editing a movie he directed about a comic, chain smoking cigarettes, and dealing with his young girlfriend (Ann Reinking). He is unpleasant to most people, including his ex-wife (Leland Palmer). The only exception is his sweetness with his daughter, Michelle (Erzsebet Foldi). He also fl irts with a white angel named Angelique (Jessica Lange). The morning
150
ReAction! Chemistry in the Movies
preparation routine is repeated four more times, but during the fi nal one he is coughing before he says “It’s show time, folks” and then he has a heart attack. Commentary: Dexedrine and Benzedrine are amphetamines that are prescribed as appetite suppressants, but both are among the most commonly abused prescription drugs. Each induces a low level of euphoria in addition to increasing mental and physical activity and wakefulness. People who take high doses will often stutter from the overstimulation. Chronic use is associated with heart failure. Bob Fosse cowrote and directed this semiautobiographical tale that famously predicted his own death from a heart attack 10 years later (Gottfried 1990). In real life, Fosse simultaneously edited Lenny and staged Chicago while popping pills. In 2001, the U.S. Library of Congress selected this fi lm for preservation in the National Film Registry. THX-1138 (1970) Distribution company: Warner Bros. Director: George Lucas Screenwriters: George Lucas and Walter Murch, based on George Lucas’s 1967 student fi lm Electronic Labyrinth: THX 1138 4EB at the University of Southern California Short summary: LUH 3417 disrupts THX 1138’s depressant SP5 intake MPAA rating: PG Plot description: THX 1138 (Thex; Robert Duvall) and LUH 3417 (Luh; Maggie McOmie) work all day using remote arms to build nuclearpowered devices. Everyone is bald and wears androgynous clothing. At work, they hear an announcement over the intercom say that improper sedation is a punishable offense. On the way home, Thex enters a “confessional” booth with an illuminated version of Albrecht Dürer’s Jesus face. He confides that his concentration hasn’t been very good and that his mate has been acting strangely. He doesn’t think his SP5 etrazene is strong enough. A voice comes through a speaker, responding generically to be thankful you have a job to fi ll. In fact, Luh hasn’t taken her medication for some time and has begun to give placebos to Thex. Soon, they embrace, kiss, and make love without realizing that their every move is being fi lmed and monitored. Commentary: The etrazene depressant in this fi lm resembles the drug “soma” in Aldous Huxley’s 1932 novel Brave New World. In the novel, a dystopia is imagined in which a dictatory, mechanized government controls people’s emotions by banning art and beauty but keeps everyone content by supplying them with “soma,” the “ideal pleasure drug.”
A Master/Slave Narrative
151
Easy Rider (1969) Distribution company: Columbia Pictures Director: Dennis Hopper Screenwriters: Peter Fonda, Dennis Hopper, and Terry Southern Short summary: Wyatt and Billy score big on a drug deal and then travel from Los Angeles to New Orleans in search of America, picking up George Hanson along the way MPAA rating: R Plot description: The movie opens with Wyatt “Captain America” (Peter Fonda) and Billy (Dennis Hopper) purchasing an unnamed white powder that they snort in a Mexican junkyard. They travel to Los Angeles, where they sell the powder to a man (Phil Spector) in a chauffeur-driven Rolls Royce for even more money. Steppenwolf’s “The Pusher” plays over the soundtrack as Wyatt wraps his money and stashes it into his chopper’s teardrop gas tank. As they hit the highway for Mardi Gras in New Orleans, Steppenwolf’s “Born to be Wild” plays. They stop at a hardscrabble commune for a few days before they are on their way again. In the next small town, they are thrown in jail for a minor offense, where they meet the alcoholic ACLU lawyer George Hanson (Jack Nicholson). Hanson helps them get out of jail so he can join them on their trip. He always wanted to visit a particular brothel in New Orleans. In the scene at 58:00, they are camping and Hanson smokes his fi rst marijuana cigarette. He’s reluctant because “it leads to harder stuff.” Commentary: In 1998, the U.S. Library of Congress selected this fi lm for preservation in the National Film Registry. Even though the fi lm begins with a cocaine deal and ends with an LSD trip, the most memorable drug scene occurs when Hanson smokes his fi rst marijuana cigarette. In fact, the fi lm is significant for the many other scenes in which the characters casually smoke marijuana without drawing attention to the fact. One-third of Americans have tried marijuana once in their lives, and 6% continue to smoke it at least once a month (Koob and Le Moal 2006). This makes it the most commonly used illicit drug in the United States, and coincidentally, it has a long history of portrayal in the movies (Starks 1982). Marijuana is the common name for the Asian herb Cannabis sativa. Its tough fibers, called hemp, have been used throughout history to produce cord and rope. Its flowers, leaves, and stems are covered in an oil that prevents water loss when the plant is grown in its naturally dry habitat. The sticky oil is richest on the female flowers and is called hashish after it is crudely extracted. Hashish can have as high as 20% tetrahydrocannabinol (THC; figure 5.8) (Gaoni and Mechoulam 1964). THC and other cannabinoids are also psychoactive compounds that bind to cannabinoid receptors. It is still unclear how THC stimulates the brain’s “reward system” to cause euphoria, but it is clear that when THC
152
ReAction! Chemistry in the Movies
OH
O
THC (² 9-6a,10a-trans-tetrahydrocannabinol) O
O NH CH2CH2OH
O
CH2OH
CH CH2OH
Anandamide (N-arachidonoylethanolamine)
2-AG (2-arachidonoylglycerol)
Figure 5.8. Brain cannabinoid receptors bind to natural anandamide and 2-AG, and the plant-derived psychoactive compound THC.
binds to cannabinoid receptors in the hippocampus, it interferes with memory, and when it binds to receptors in the cerebellum, it interferes with coordination and balance. The brain’s natural ligands for the cannabinoid receptors are called anandamide and 2-AG, but their physiological roles are not well established, even though these receptors are vastly more abundant in the brain than the very well-studied opioid receptors. The Valley of the Dolls (1967) Distribution company: 20th Century Fox Director: Mark Robson Screenwriter: Helen Deutsch, based on the 1966 same-titled novel by Jacqueline Susann Short summary: Three young women enter different parts of show business, begin using drugs, rise to the top, and then fall MPAA rating: R Plot description: Pill capsules are invisibly pulled apart, spilling their contents during the opening credits, and then the lives of three women starting their careers in the entertainment business are chronicled. Neely O’Hara (Patty Duke) is the one who rises the highest and who falls the furthest due to drug abuse. At 40:40, O’Hara accepts some pills from her dance trainer even though her manager husband Mel (Martin Mulner) shakes his head no. Her impulse is proved right when her next Broadway show is a success and generates a hit song. She moves to Hollywood to
A Master/Slave Narrative
153
make movies, but she soon begins showing up late for work. At 1:00:00, Mel complains that she takes pep pills to wake up from her sleeping pills. Their argument leads to a divorce, and she begins taking so many pills that she has a psychotic experience. At 1:29:00, she says she can’t get through the day without her “dolls,” because “they’re harder to quit than booze.” Her friends convince her to enter a sanitarium, where she suffers through withdrawal and recovers to reenter show business. Unfortunately, the pressure of opening night proves too much. Commentary: None of the pills are named in the movie, even though the characters use and abuse them for different reasons. This fosters a mood that there are pills for every need and that all pills are abusable. In the movie’s trailer, each character is associated with a different colored pill. For instance, O’Hara is said to be taking the red pills, which also happens to be the color of the Dexedrine pills she takes in the novel. Wikipedia lists Jacqueline Susann’s novel as one of top 30 best-selling books of all time. Its narrative differs significantly from the fi lm in a number of respects but not much with regard to O’Hara’s character development. In the novel, O’Hara’s friend gives her Dexedrine to lose weight, which is what they were commonly prescribed for during the 1960s and 1970s. Dexedrine has an amphetamine-like structure and effect, so the user feels energized. It is also mentioned in the novel that O’Hara takes Seconal to go to sleep. That was the market name for secobarbital, a member of the barbiturate class of compounds that induce varying degrees of the anesthesia, anticonvulsion, sedation, and hypnosis. Barbiturates are prescribed for a wide range of conditions and were also widely abused during the 1960s and 1970s. In fact, several movie stars used them to commit suicide, such as Marilyn Monroe in 1962, Judy Garland in 1969, and Charles Boyer in 1978. They also played a role in Jimi Hendrix’s accidental death in 1970. Seconds (1966) Distribution company: Paramount Pictures Director: John Frankenheimer Screenwriter: Lewis John Carlino, based on the 1963 same-titled novel by David Ely Short summary: Investment Banker Arthur Hamilton is transformed into painter Tony Wilson, who fi nally learns what life is all about but then drinks too much MPAA rating: R Plot description: Arthur Hamilton (John Randolph) is an investment banker with all of the trappings of success, including a devoted, beautiful wife and married daughter. Nevertheless, he isn’t happy, and any satisfaction that he might have felt for his life is destroyed when he receives
154
ReAction! Chemistry in the Movies
a phone call from his old college friend Charlie Evans (voice of Murray Hamilton), whom he thought was dead. Charlie tells Arthur that the “Company” can give him a chance for a second life. After Arthur agrees to sign over $30,000 of his insurance, his death is faked with another man’s corpse, and then his teeth, vocal chords, fi ngerprints, and nose are reshaped through surgery. At 40:00, he is told that his deepest desires were probed while he was under pentothal and sodium caffeinate. Between the two dream occupations he mentioned, the “Company” decides he’ll become established painter Tony Wilson (Rock Hudson), who lives in a studio on Malibu beach and who is cared for by John (Wesley Addy). After a period of seclusion, he meets Nora Marcus (Salome Jens) while walking on the beach. They become friends, and she invites him to a party in Santa Barbara. The party is actually a bacchanal with dancing revelers who stomp grapes in the nude. After being tossed into the grape vat, Tony struggles to get out but then has an epiphany and realizes that this is what life is all about. He is so happy that he hosts a cocktail party for his neighbors. Unfortunately, he drinks too much and reveals the truth about his transformation. Commentary: A barbiturate “truth serum” is used to discover what Arthur really wants out of life. After Tony drinks too much alcohol, he can’t stop himself from telling the truth. Depending on the dose, barbiturates can reduce inhibitions, induce anesthesia, or induce a state of unconsciousness. Like alcohol, barbiturates are general inhibitors of nerve transmission throughout the brain, which is why they are depressants. They bind to the GABA (γ-aminobutyric acid) receptor, which is naturally complexed with chloride ion channels (Leeb-Lundberg et al. 1980). The specific effects of sedation and hypnosis occur when barbiturates bind to GABA/chloride complexes located in the cerebral cortex, the cognitive part of the brain. Thiopental (trade name Pentothal or Sodium Pentothal) is a barbiturate that has been used in surgical anesthesia since 1934 (Robinson 1946). Its use as a “truth drug” was discovered accidentally but was soon used by psychiatrists at lower doses to help their patients discuss their thoughts (Winter 2003). Although there is no drug or combination of drugs that can cause people to tell the truth unfailingly, these “truth drugs” place people in a semiunconscious or hypnotic state to access thoughts normally buried. People under the influence also happen to be especially susceptible to suggestions, which is why barbiturates are also called “brainwashing” drugs.
6 Inventors and Their Often Wacky Chemical Inventions
INVENTION, DISCOVERY, AND SERENDIPITY Sidney Stratton (Alec Guinness in The Man in the White Suit) knows exactly what he wants to make. He just doesn’t know how to make it. So, he engages in a trial-and-error search for the right conditions to create his nonstaining fiber. Every time he makes a new trial, however, he sets off an explosion. As the Birnley Mills building crumbles around him, he tries, tries, and tries again. Like Stratton, most movie inventors create oxymoronic products such as rechargeable batteries, flexible glass, bulletproof tires, and water-repellent hairsprays (table 6.1). Movie inventors are very closely associated with the slapstick humor of the 1910s to 1930s, but ultimately they owe the strength of their fictional existence to Thomas Alva Edison. His inventions brought him worldwide fame in 1877, when he was 29 years old. After that, he regularly made front-page news until his death in 1931. His creation of the phonograph, commercialization of the light bulb, and 1,091 other inventions changed the way we live. Of all his inventions, the phonograph truly came out of nowhere, so much so that a journalist dubbed him the “Wizard of Menlo Park.” He followed that up with the electric light bulb and, more important, the electric power generation and delivery system. His most profound creation was the research laboratory, discussed in the next section, which he didn’t even patent. The iconic power of Edison is evident in the observation that inventors before The Absent-Minded Professor in 1961 create in the absence of theory, while those after 1961 rely on theory to make their products. Edison wanted to invent things that interested him. He didn’t care how they worked, just that they did. He hired men with advanced degrees for their theoretical expertise but relied on them more for their specialized technical abilities. In contrast, the industrial research laboratories that were founded on Edison’s example, such as General Electric Laboratories and Bell Laboratories, among many others, were and are staffed by large numbers of trained scientists, engineers, and technicians who rely on the free flow of ideas and expertise between theory and practicality to solve problems. 155
156
ReAction! Chemistry in the Movies
Table 6.1. Chemical inventions in the movies Title (Year)
Invention
Who Killed the Electric Car? (2006) Flubber (1997) Romy and Michele’s High School Reunion (1995) Back to the Future (1985) Caprice (1966) Lover Come Back (1961) The Absent-Minded Professor (1961) The Man in the White Suit (1951) Yellow Cab Man (1950) Edison, the Man (1940) Beauty for the Asking (1939) Mr. Edison at Work in His Chemical Laboratory (1897)
Rechargeable batteries Flying rubber Nonsticky glue Time travel machine Water-repellent hairspray Alcoholic mints Flying rubber Nonstaining fiber Elastiglass Incandescent light bulb Astringent cold cream Unnamed solution
In his day, Edison was derided for his reliance on trial and error, which he had to use in the absence of theory. He relied on the storehouse of knowledge in his head and the published literature. Most movie inventors rely on trial and error to create their products. Edison was more practical-minded than most inventors in the movies since he made sure there was a market for his products before he developed them. Movie inventors are either poor or independently wealthy. Edison built his Menlo Park laboratory and house a short distance from each other so he could travel between them as quickly as possible. Nearly all movie inventors work out of their garage or basement. The “good news” researchers in chapter 9 invent things that solve social problems, carry out their work in reasonably well-appointed laboratories, develop theories, and work methodically, sometimes in collaboration with teams of researchers. The main distinction between real inventors and scientists is that the former create new things while the latter discover new things systematically. This is a fuzzy distinction, though, because inventors have made many highly practical and important discoveries. The most famous are the lucky accidents, or serendipitous discoveries, that have often solved problems no one was pursuing. The surprising connection between the accident-prone but purposeful movie inventors is that their oxymoronic products are often based on actual inventions that resulted from happy accidents. Chemistry professor Royston Roberts wrote a 1989 book called Serendipity with many examples from the fields of pharmacology and chemistry plus a few others (Roberts 1989; Ball 2006). (See the movie commentaries in chapter 5 for examples such as LSD and Thorazine and this chapter for the phonograph and safety glass.) According to Roberts, there are two types of serendipitous discoveries. In one type, accidents occur while trying to solve
Inventors and Their Chemical Inventions
157
a problem that leads to a solution for that very problem. A researcher might mix reagents incorrectly, observe something unexpected but that was the natural outcome of the incorrect mixture, and then backtrack to figure out what happened. The other type of serendipity occurs when the researcher makes a discovery unrelated to the question being probed. There are many examples of both types. French chemist and microbiologist Louis Pasteur summed it up when he said, “Where observation is concerned, chance only favors the prepared mind” (Debre 1996). Pasteur is famous for the many microbiological problems he solved but also for his 1848 discovery of molecular asymmetry. That discovery was serendipitous for many reasons. During his doctoral thesis research in chemistry, he discovered that his seemingly pure compound formed two kinds of crystals that were mirror images of each other. He went on to prove that his compound was actually a mixture of two molecules that were asymmetric, a discovery that has had a profound impact on every other scientific field. Even though his discovery was serendipitous, it was inevitable that someone would soon make this discovery because crystallography was in its infancy, and other researchers were working in this same fertile area. He was also lucky he was working on industrially produced sodium ammonium tartrate during a cold February. Otherwise, he never would have obtained the chiral crystals (Kauffman and Myers 1975). In 1848, Pasteur was working on his doctoral thesis in chemistry at the École Normale in Paris. He had taken a course in crystallography from Gabriel Delafosse, where he learned how to measure crystal shapes (goniometry). Enigmatically, quite a few organic compounds formed the same crystal shapes. In his chemical work, Pasteur learned about tartrate and its long history of industrial production as a mordant for dyes and other uses. He also learned that racemic acid was fi rst observed by an Alsatian winemaker and then discovered to have the same elemental composition as tartaric acid but without its optical activity (the Latin racemus means “bunch of grapes”). Pasteur decided to study the crystals formed by these compounds. He began his project by trying to repeat some published research but noticed something the previous researcher had not: Under the microscope, his crystals of sodium ammonium “racemic acid” were present as two types that were mirror images of each other. Pasteur separated the crystals using tweezers and soon found that one of his chiral crystals had the same shape as crystals of tartaric acid. When he dissolved that crystal, it had the same optical activity as a tartaric acid solution. This proved “racemic acid” was an equal mixture of two compounds, one of which was the same as biologically produced tartaric acid. The other type of crystal had the opposite optical activity when it was dissolved. He famously explained this result by concluding that the molecules in the crystals have different shapes (he used the word “dissymmetric”).
158
ReAction! Chemistry in the Movies
There are two serendipitous aspects to Pasteur’s fi nding (Kauffman and Myers 1975), neither of which diminishes the importance or validity of his discovery. Of all organic compounds known at that time, only sodium ammonium tartrate (the one he was studying) and sodium potassium tartrate separate into chiral crystals. He was also lucky to be working in February and to have placed his solutions on the windowsill to crystallize overnight. Tartrates only crystallize chirally at temperatures below 26°C. At warmer temperatures, the two enantiomers crystallize together to form an altogether different crystal structure. Today, the word “racemic” means an equal mixture of two enantiomeric compounds (see chapter 1).
THOMAS ALVA EDISON, AMERICAN INVENTOR In the 1940 movie Edison, the Man, Edison builds a house and research laboratory in Menlo Park, New Jersey, with the money he earns from selling a patent for a ticker tape machine to an investor he doesn’t like. His small staff of eight or so employees travel with him from the Western Union research laboratory, where he had workspace but was not employed. In Menlo Park, he and his team soon invent the electric pen and a few other things. After two years, the income is less than what they owe, so the sheriff announces foreclosure in a month. By chance one day, the telegraph recording machine produces a patterned noise when one of the workers turns up the power. Edison rewinds it, replays it, and it reproduces the same noise pattern. He develops a drawing and hands it to a worker who builds it, and the phonograph works. It is a great success, and they are back out of debt. More patents emerge. The cycle repeats itself, and they are soon in such fi nancial straits that Edison gives them severance pay. None of them leaves, and they invent the light bulb, the central story of the movie. In real life, Edison’s research and fi nancial independence were a bit more complex. He was already moderately wealthy when he built the Menlo Park facilities in spring 1876. Menlo Park is 25 miles southwest of New York City and 25 miles south of Newark. He decided to leave Newark because he believed his landlord was overcharging him (Jehl 1936 and 1938). Still working for Western Union but seeking greater independence, he happily sold his manufacturing facility in Newark to generate the funds to build the house and laboratory he designed. The electric pen was indeed one of the earliest devices produced at Menlo Park. It consisted of a motor that drove a shaft rapidly up and down so that it perforated the page. When this master page was mounted and inked, it could be used to generate many copies. Mimeography was immediately adopted by organizations that wanted to produce hundreds of messages cheaply. The movie’s depiction of the serendipitous inspiration for the phonograph is based on the story recorded by Francis Jehl in his 1936 book
Inventors and Their Chemical Inventions
159
Menlo Park Reminiscences (Jehl 1936 and 1938). Jehl was 18 years old in November 1878 when he joined Edison’s team, one year after the phonograph had been patented and one year before the development of the light bulb. Jehl would later help set up the English, French, and Italian Edison companies and then return to the United States to help Henry Ford’s team set up the Edison Institute in Dearborn, Michigan. While John Kruesi, head of the shop operations, was building the prototype for the phonograph, he had no idea what Edison had in mind. Until this time, nearly all of their devices were electric, but this one was hand-cranked. It consisted of a piece of tin foil wrapped around a cylinder, a crank on one end of the cylinder that turned a screw to pull the cylinder to one side, and a cone attached to a stylus that touched the tin foil. When the workers saw it, they were astounded by its ability to play back Edison’s recitation of “Mary had a little lamb.” In December 1877, Edison exhibited his cylinder phonograph to the editors of Scientific American, who also believed it was something truly new (Anonymous 1877). This invention made Edison world famous. He marketed the phonograph as a dictating machine for businesses, but it was never really successful for that purpose. Other entrepreneurs began to make money with recorded music, and still others encroached on his patent by making substantial improvements to his design. He responded by selling recorded music and making even more improvements to his phonograph. He replaced the tin foil with hard brown wax cylinders and later celluloid cylinders. The industrial and federal research laboratories that dominate our current invention landscape are Edison’s greatest legacy (Vanderbilt 1971). His team’s improvement of the light bulb provides the perfect demonstration of Edison’s method. In fall 1878, Edison formed the Edison Electric Light Company to attract venture capital. This business would not only manufacture a useful product but also deliver the electricity to power it. Even though he didn’t have a prototype for the bulb to show investors, he could say that other men had invented various light-emitting devices that he knew he could improve to a point where they could be used in the home less expensively than gas or oil. Platinum was the most popular choice by others, so he set his chemist, the American Dr. Alfred Haid, to the task of its purification. Dr. Haid used the periodic law of elements to guide his choices for other metals to purify. The periodic law had been independently discovered by Mendeleev and Meyer only a few years earlier in 1869. Few of these metals proved satisfactory because they were too brittle or melted or combusted in response to the high temperatures generated by the current. Assuming that combustion was the problem, they next tested metal oxides, and then everything again in gas-evacuated glass bulbs. This was a daring choice because the original conception was that consumers would be able to reuse their bulbs when the wire burned out. The vacuum considerably increased the lifetime and luminosity, but not enough. Prior efforts by others and his own team’s prior work with the electrical conduction of carbon led them to test many
160
ReAction! Chemistry in the Movies
organic fibers, including John Krusei’s beard and J. U. Mackenzie’s side burns. The chemistry is further described in the commentary section for this movie below. Edison still holds the record for the number of U.S. patents given to an individual. Most of his 1,093 patents involve electrical devices or improvements, which is also true of the average U.S. patent. Whether the inventor is an individual or an organization, a patent is a right granted by a government to make, sell, import, or use something exclusively. In the United States, patents are administered by the U.S. Patent and Trademark Office (www.uspto.gov). To be eligible for a patent, an invention must be new, useful, original, and not easily discovered or created. In the United States, the monopoly granted by the patent begins on the date the patent is issued and expires 20 years after the patent application was fi led. An inventor may sell all or part of the rights given by a patent for a flat fee or for royalties, which are payments based on sales. The Edison laboratory created other forms of intellectual property, namely, copyrights. A copyright is the right granted by a government to disseminate intellectual property such as literary, artistic, or musical work. Some of the fi rst copyrights claimed by the Thomas A. Edison Company were for their phonograph sound recordings. When the company began producing kinetoscope movies (see Mr. Edison at Work in His Chemical Laboratory [1897], below), it copyrighted the collection of images that constituted the movie by printing them in book form and copyrighting the book. The earliest Edison kinetoscope fi lms still exist and are reproducible for this reason. When a patent or copyright expires, the intellectual property falls into the public domain and anyone can use, make, or sell it without paying royalties or becoming liable in a lawsuit. In the United States, copyrights remain in force for the author’s lifetime plus 70 years, or 95 years from publication if the author was hired by someone else to produce the work. The U.S. Library of Congress has a website called “Inventing Entertainment” (lcweb2.10c.gov/ammem/edhtml/edhome.html) with examples of the earliest American movies and sounding recordings, all of them now in the public domain.
INSIDE OUT OF THE BOX The plot of Yellow Cab Man (1950) revolves around elastiglass, a form of safety glass invented by Red Purdy (Red Skelton). In a scene at 9:20, characters with suspicious motives ask him whether he has protected his idea. Red says he’s had difficulty patenting it but, not to worry, his head holds the secret to the way it’s made. Gesturing to his head, he says, “Greatest safe deposit box ever invented; nobody can open it and nobody can get a thing out of it.” Red Skelton, Lucille Ball, Rock Hudson, Alec Guinness, and Doris Day (to name a few) may have it, but creativity is the real star power in
Inventors and Their Chemical Inventions
161
chapter 6. Words like discovery, intelligence, innovation, imagination, originality and genius orbit around it. The red carpet welcomes its accomplishments from science, engineering, the arts, education, architecture, design, music, and entertainment. Wannabes from other professions such as business, fi nance, law and health care cluster at the velvet ropes hoping for a glimpse, a smile, a touch, or that rare kiss. Everybody, it seems, wants a piece of creativity. But just what is it? Defi nitions are elusive. Often called “divergent thinking” or “thinking out of the box,” creativity has a lot to do with deviating from the norm and breaking familiar thought patterns associated with materials and ideas. To evolutionary biologist E. O. Wilson, the creative process is an “opaque mix” that involves knowledge, obsession, and daring (Wilson 1998). About the thrill of discovery, Wilson speaks for all scientists when he says, “There is no feeling more pleasant, no drug more addictive.” Discovering and making are inherently pleasurable and seem to be rewarding in themselves. Writer Diane Ackerman calls this unique human behavior “deep play” and views creativity as the “ideal playground” for this absorbing, transcendent activity (Ackerman 1999). She notes that play is widespread in the animal world and is fundamental to evolution. Playful behavior involves problem solving, setting limits, and developing strategies. Deep play is closely associated with rapturous and ecstatic states of being where one feels both uplifted and positioned outside of oneself at the same time; it also imparts a feeling of wholeness. To Ackerman, deep play inspires many elements of culture, including exploration and discovery. Deep play and psychology professor Mihaly Csikszentmihalyi’s concept of “flow” have much in common. He coined this term to defi ne an optimal experience that comes about when someone is highly focused on the kind of work or activity that stretches that person’s capacity and involves elements of novelty or discovery (Csikszentmihalyi 1996). Along with scientific pursuits, painting a picture or rock climbing are examples of flow-producing activities. Time and body are forgotten, and the activity feels enjoyable, automatic, and effortless. Csikszentmihalyi explores flow’s connection to the creative process in his 1996 book Creativity: Flow and the Psychology of Discovery and Invention. For those who aren’t getting it, he suggests cultivating flow in everyday life by developing a sense of curiosity and surprise. This can lead to the “self-sustaining chain reaction of creativity.” Csikszentmihalyi’s detailed study of the creative process follows nearly 100 people whose public lifetime achievements have left a mark on the culture. In addition to cultivating flow in their various pursuits, he found that creative people have complex personalities capable of holding contrasting traits in dialectical tension. They can experience the extremes of each polarity fully without inner confl icts arising; they are able to access “tendencies of thought and action that in most people are segregated.” Csikszentmihalyi has identified 10 pairs of contrasting traits that are present in creative people. Creatives are physically energetic, yet capable of
162
ReAction! Chemistry in the Movies
quiet and rest. They are smart and naive, playful and disciplined. They use imagination yet are rooted in reality. They are extroverted and introverted, humble and proud. They can be masculine and feminine. Here, he notes this relates to possessing psychological androgyny, in the sense of resisting rigid socially constructed gender roles. Creative people are also traditional and conservative when paired with rebellious and iconoclastic. They have a passionate connection to their work yet are also objective. They are capable of experiencing suffering and pain and also great enjoyment from what they do. Csikszentmihalyi suggests that this ability to work with an expanded repertoire of traits gives creative people an edge in fi nding a problem to solve. New ideas need to be recognized and developed. Those who can do both are said to be creative. It seems appropriate at this point to fi nd ourselves thinking back to chapter 1 and Dr. Jekyll’s destructive act of separating the poles of his personality. We may wonder what that gifted chemist could have done if he had instead played both sides off of one another. Today, the creative types belong to an influential economic class all their own, as economist Richard Florida describes in The Rise of the Creative Class (Florida 2002). The core of this creative class is composed of the disciplines making their way down our red carpet. Those clustering at the velvet ropes are members of the broader creative professionals. Comprising more than 30% of the workforce, the economic function of the creative class is to come up with new ideas, new technologies, and new creative content. Unlike the working class and service class, those in the creative class are paid to create. Thomas Edison in Edison, the Man (1940) would have appreciated the autonomy and flexibility this new class enjoys. In a scene at 50:00, Edison is out of money for his projects, so he pays a visit to a potential venture capitalist. He asks for $50,000 and, to his surprise, is offered $100,000. There is a catch, however. The investor says he will direct which projects Edison pursues. Edison says, “I’m an inventor; I can’t be told what to do. I work with ideas, visionary things. Nobody, not even I, know whether they’ll be profitable unless I work them out in my own way.” Florida contends that the “human creative faculty” is the true source of economic value held by the creative class. The key point is that the “useful new knowledge” found in intellectual properties such as computer programs, patents, and formulas originate in the minds of people.
IT ISN’T ALL ABOUT HIM As we take our fi rst bold steps into the bright side of chemistry in the movies, we may hesitate a bit as we ask this question: How do you recognize a chemist when you see one in the movies? The hesitation is justified, for our feet have just landed on the thin surface of stereotype. Consider that the richest cluster of women chemists in the movies are inventors of
Inventors and Their Chemical Inventions
163
face creams or perfumes, the most mentionable female-specific products. This can be understood to reflect the movie reality that science is fi ltered through an emotive and personal lens. You can show the femininity of the female inventor by having her develop a face cream. Since her product will be sold primarily to women, her prowess is perceived as less threatening to her male suitors because she isn’t competing in their “masculine” domain. On the other hand, this seeming cliché also reflects the historical reality that the cosmetics industry has always been a fertile field for women inventors and entrepreneurs. Kathy Peiss, in her 1998 book Hope in a Jar, notes that until recently, most of America’s richest women entrepreneurs came from the beauty industry. She also describes the transformation of the industry over the past century. It was originally local and service driven and dominated by women. After WWI, it became increasingly mass-produced, massmarketed, and multinational and run mostly by men. Even so, women are still found in greater abundance in this industry in the advertising and marketing departments than in other types of corporations. The beauty business was created largely by women between the 1890s and 1920s, coinciding with the increase in women living and working in cities, where a “public face” became more important. In her conclusion, Peiss writes, “Women still perceive beautifying as a domain of socializing, creativity, and play.” Many of the great originators of this industry were immigrants. One example is Helena Rubinstein, who started selling a facial cream in Australia created by her Hungarian relative Jacob Kusky, a chemist (Peiss 1998). By 1908, she had opened a beauty salon in London, and by 1912, in Paris. The start of WWI caused her to relocate to Fifth Avenue in New York City, where she competed with Elizabeth Arden, innovating for decades. Perhaps more important for a book on chemistry in the movies is the story of Max Factor. He fled to America in 1904 from his very successful position as beautician to the Russian royal family. By 1908, he was running a barbershop, wig business, and makeup studio in Los Angeles, where he catered to the stage and screen actors who were just beginning to populate that city. In 1914, he invented a “flexible greasepaint” for movie actors that he sold in a squeeze tube. It looked natural and didn’t crack under the hot lights of the movie set. Local women began to purchase it for home use as a foundation, and by 1920 he was marketing it as Pan-Cake in his Society Makeup line of cosmetics that also included the eyebrow pencil. His children combed fi rst the West Coast and then the East Coast to place the product in drugstores. By 1927, his products were being distributed across the country, and he began his fi rst advertisements in movie and romance magazines, all of them featuring screen stars. Among Hollywood movies, the earliest example of cosmetics invention may be The Blooming Angel (1920), in which Claire Floss (Madge Kennedy,
164
ReAction! Chemistry in the Movies
a popular actress at that time) invents “Angel Bloom” cream. She and her struggling student husband promote it. In the light drama Beauty for the Asking (1939), described further below, Jean Russell (Lucille Ball) develops an astringent facial cream so successful she becomes the president of her own company. In 1958, Myrna Duchane (Doris Dowling) is an assistant to perfume chemist Mr. Atterbury (Jonathan Kidd) in the British comedy Wink of an Eye. In the spy spoof Caprice (1966), also described below, Madame Piasco (Lilia Skala) plays a small but critical role as the inventor of a water-repellent hairspray and other cosmetics. Her son-inlaw claims them as his own in his role as head of research development for a large cosmetics company. In the comedy Death Becomes Her (1992), Lisle von Rohman (Isabella Rosselini) sells an immortality formula in return for her customer’s souls. Most recently, Patience Phillips (Halle Berry) in Catwoman (2004) works in the advertising department of a cosmetics company. She learns their antiaging cream is addictive and deadly when its use is discontinued but is assassinated before she can blow the whistle. She survives to become the Catwoman. There are also at least three women presidents of companies that produce personal care products in the movies. In the delightful Trouble in Paradise (1932), Mariette Colet (Kay Francis) runs Colet Parfumerie. In the B movie Wasp Woman (1960), Janice Starlin (Susan Cabot) runs Starlin Enterprises, a cosmetics fi rm with an all-male research team that develops a youth formula from wasp jelly. In The Main Event (1979), Hillary Kramer (Barbara Streisand) discovers that her accountant stole everything she owned except a boxer purchased as a tax write-off. The common features of these movies is that the women know how to create their inventions, have some understanding of the market for their product and are either in love or fall in love, and their product drives the narrative of the movie. These inventors are closest to the “lonely heroine” type of women scientist in the movies (Flicker 2003) developed by Dr. Eva Flicker, University of Vienna professor of sociology. According to Flicker, this type is the most competent specialist in her area, and her sexual relationships are secondary to scientific relationships. This defi nition provides another insight into the differences between inventors and scientists in the movies. Inventors can be represented as “normal” people who dabble in their spare time; these characters are more sympathetic to the mostly nonscientist audience. In contrast, scientists have to know facts and act in accordance with them; these characters make judgments that aren’t fi ltered through their emotions, making them a matter of curiosity but not identification for most viewers. In a number of movies, the female scientists are given male or genderneutral names. In some, they are mistaken for men until they make their appearance. In one way, this genderlessness signifies that their scientific credentials are of primary importance. This practice may originate from the fast-talking comedies of the 1930s but, with regard to chemists in the movies, begins with two women forensic chemists. In Kid Glove Killer
Inventors and Their Chemical Inventions
165
(1942), Jane Mitchell has a master’s degree in chemistry from Montana and has just joined the Chatsberg police forensic laboratory (see chapter 7). She knows how to use an atomic emission spectrometer. Her supervisor and only coworker, Gordon McKay, calls her “Mitchell” in part to respect her position, but also in part to irk her because they are actually falling in love. In the scientific horror fi lm Undying Monster (1942), the forensic laboratory assistant Miss Christopher grows bored if their latest case doesn’t make her skin crawl. Everyone calls her “Chris.” She uses an atomic emission spectrometer to prove a killer is a werewolf. In the 1950s science fiction B movies (Noonan 2005), male or genderneutral names were given to female scientists for 6 of 50 major women’s roles in 44 movies (data gathered from the fi lmography in Noonan’s book). Five of these characters are scientists, and one is a physician. The 50 women portrayed geologists, physicists, marine biologists, rocket fuel scientists, among many others. Within this genre, graduate student Stephanie Clayton in Tarantula (1955) just had a paper published on the nutritional aspects of populations. She travels alone to a remote Arizona town to begin work on a radioactive nutrient with a renowned nutritional biologist. When she arrives, she tells everyone to call her “Steve,” which they do. Today, the battle of the sexes becomes overt. In Medicine Man (1992), “Dr. Crane” is revealed to the audience to be a woman only after she travels a great distance by many conveyances to a remote Amazon village. Just as the audience takes this in, the botanist she meets is disgusted and says, “They sent a girl?” Her response is, “I’m published, and more extensively than Dr. Sealove. I hold degrees from CCNY, Berkeley, and Cambridge. I’m the recipient of the Thurman Award in ’82 and ’86, the fi rst and only time it’s ever been given to the same person twice.” In that statement, she more than establishes her scientific competency. Soon we learn she is a biochemist and that she knows how to use a gas chromatograph.
THE ARCHETYPE MOVIE: THE MAN IN THE WHITE SUIT (1951) Distribution company: GFD, UK; made in Ealing Studios Director: Alexander Mackendrick Screenwriter: Roger MacDougall, with John Dighton and Alexander Mackendrick, from one of MacDougall’s unproduced plays about the invention of a cleaning fluid Short summary: Chemist Sidney Stratton develops a fabric that can’t be stained, get dirty, or wear out Plot description: Sydney Stratton (Alec Guinness) lost his Cambridge University scholarship for mysterious reasons. We soon suspect it was due to the explosions that accompany his polymerization process, which requires “heavy hydrogen” and “radioactive thorium.” After losing his
166
ReAction! Chemistry in the Movies
eighth mill job due to lack of credentials and overspending, he is accidentally asked to train researchers to use the new electron microscope at Birnley Mill Works. In the 4-minute scene beginning at 19:15, the mill owner’s daughter Daphne Birnley (Joan Greenwood) discovers Sydney’s deception and intends to alert her father. Stratton catches her as she is about to race away in her sports car and explains his new fibers. They are very long copolymers of amino acids and carbohydrates with ionic cross-links for extra strength plus the ability to repel dirt. In plain terms, the new fiber doesn’t stain, dye, or tear. She becomes his ally and even convinces her father to give him extra help. After practically destroying the millworks, Stratton has his white solution isolated in a round-bottomed flask that he proudly shows to his supervisor. We watch the industrial production of fibers change into material and then into the tailoring of a white suit for Stratton. The fiber is so strong a blowtorch is used to cut the pattern from the cloth. When the significance that the fiber won’t stain or get dirty dawns on the millworkers, they spread the word that it will drive them out of their jobs. Simultaneously, the mill owners also become concerned. When they meet, they realize Stratton is naïve, so they offer to buy the rights to manufacture his fiber even though they never intend to do so. Stratton is wearing his white suit when he meets the mill owners but, ultimately, decides not to sign. As he runs away, the millworkers chase him until it rains and his suit dissolves in clumps. His fiber wasn’t perfect after all. As he walks away from the factory with his head hanging low in defeat, the intensity of the bubbling apparatus musical theme increases until he raises his head in triumph. He has a new idea and he is electrified to begin experimenting again. Commentary: This fi lm and Dr. Ehrlich’s Magic Bullet (1940; see chapter 9) have among the highest quotient of chemical research per screen time of any feature fi lm. This one relates invention to funding, employment, safety, and commercialization better than any other fi lm. And, it is loosely based on the true story of the introduction of nylon! Stratton’s “white suit” represents purity, innocence, and disinterested science. As soon as he puts it on, however, the black-suited corporate heads unite with the dirt-covered union workers to suppress his fi nding. Stratton was defeated because he didn’t consider the social implications of his discovery. Even by the end of the movie, he is only interested in his idea. The pun of this movie is that polymerization chain reactions are confused with nuclear chain reactions. Polymers such as nylon are formed from monomers in a sequential (or linear) series of events that take place at a roughly constant speed. After certain isotopes of uranium and plutonium have been formed, they decay quickly to release more than one neutron, which then activate adjacent nuclei exponentially. Stratton’s fiber is tough like cellulose, a linear-chain polysaccharide with special hydrogen bonds that immobilize and stiffen sugars within the same chain (figure 6.1). His fiber has amino acids, the building blocks of proteins.
Inventors and Their Chemical Inventions
167
HO H
H H
H O O HO
H
H
O
OH
H
O OH
H
O H O
H HO
H
Glucose Dimeric Unit within Cellulose
O
H N
N H
O
O
O
H N
H
N
N
H O Two Adjacent Polymers of Nylon 6,6
Figure 6.1. The special hydrogen bonds in both cellulose and nylon are shown as dashed lines.
Wool and silk are fibrous proteins and are related to nylon in that their backbones are polymers of amide bonds. The regularly recurring amides in nylon form special hydrogen bonds that hold adjacent nylon chains together for flexibility along with strength (figure 6.1). Nylon was originally sold in the 1930s as a substitute for silk, which was in short supply and expensive. It doesn’t dye well and is reasonably easy to clean.
INVENTOR CHEMIST MOVIES Who Killed the Electric Car? (2006) Production company: Sony Pictures Classics Director: Chris Paine Screenwriter: Chris Paine Short summary: Stan and Iris Ovshinsky’s nickel metal hydride battery is the only suspect not guilty of killing the electric car MPAA rating: PG
168
ReAction! Chemistry in the Movies
Plot description: Martin Sheen narrates this documentary about the creation and then destruction of a fleet of electric cars leased between 1996 and 2004 in California and Arizona. At 7:15, there is a diagram of the drive train and battery, but it is short. In 1990, California Air Resources Board passed the zero emissions bill, which called for increasingly greater emissions standards from automakers for the average car they sell in California. In response, several automakers produced electric cars, and the State of California built recharging stations. We learn the most about the EV1 automobile produced, marketed, and leased by General Motors, but at 41:00, we watch a segment from the California Green television program during which a nearly new Honda EV is shredded. During the second half of the documentary, starting at 47:00, the suspected killers are summarized. In the 3-minute segment starting at 48:00, battery technology is summarized, and we are introduced to Stan and Iris Ovshinsky. Delco’s lead acetate batteries were used in the original electric cars and provided a 60-mile range. Since the average daily commute was 29 miles, this was barely sufficient for most needs. These batteries were replaced by the Ovshinsky’s unnamed batteries, which lasted significantly longer. Commentary: Stanford Ovshinsky is a true descendent of the Thomas Edison school of invention. He didn’t graduate from high school or college. Instead, he learned about science from reading books in public libraries, loves to invent things relating to electricity, personally eschews theory but hires Ph.D.s who know it, and makes public pronouncements that have placed him at odds with the engineering and scientific communities (Anonymous 2006). For instance, he sketched out a hydrogen power life cycle more than 50 years ago. No one else had such a vision at the time, and the company he cofounded with his wife Iris has since created the technology needed for a hydrogen-powered economy. More than anyone, Ovshinsky has utilized disordered amorphous materials, which are now called ovonics after the name of his company. The company also creates solar cells, fuel cells, and the nickel metal hydride batteries that power everything from laptop computers to electric cars. He is a visionary and in 2000 was honored as one of the “Heroes of Chemistry” by the Industrial Chemistry Division of the American Chemical Society for “advances in electrochemical, energy storage, and energy generation.” Flubber (1997) Production company: Walt Disney Pictures Director: Les Mayfield Screenwriter: John Hughes, adapted from the 1961 Disney screenplay for The Absent-Minded Professor MPAA rating: PG Short summary: Professor Phillip Brainard creates a flying rubber called flubber
Inventors and Their Chemical Inventions
169
Plot description: Phil Brainard (Robin Williams) is a chemistry professor at Medfield College. His rival is Wilson Croft (Christopher McDonald), the chemistry professor from nearby Rutland College who has stolen all of Brainard’s ideas over the years. Croft is also Brainard’s rival for the affections of Dr. Sara Jean Reynolds (Marcia Gay Harden), the president of Medfield College. Brainard is so absent-minded that he has already missed his own wedding twice. If he misses the third ceremony, Reynolds threatens to break off their relationship. While preparing for his wedding, Brainard realizes that he needs to change the mathematic formula that describes his reaction so that it is COLD rather than HOT (figure 6.2). He says that this will allow the Cooper pairs to form a conductive polymer and complete the metastable sphere. While he says this, the computer cursor moves an ethylene to complete a C60 fullerene (figure 6.2). He heads to his basement to put his new theory into practice. After fl ipping switches and turning dials, he plucks a hair from his head to add as an “organic catalyst,” and the solution bubbles violently. Then, he clamps power cables to a large bucketlike container and pushes the activator handle but nothing happens. From outside the house, we watch the house shake violently. After Brainard recovers, he opens the frozen bucket to fi nd some green gel that interacts with him. He says it is an elastomer, can phase shift, is foldable, is ticklish, is ductile, and is elastic. In fact, it is so elastic that it bounces through five or six houses before coming back. He names it flying rubber, or flubber. Cut to a scene in which rival Croft leads Reynolds home from the failed wedding. After Brainard learns how to use a radioactive isotope in a capsule labeled “Gamma Radiation—Metastable” to control the degree of levitation, he realizes that he missed his wedding again. Chester Hoenicker (Raymond J. Barry) has loaned Medfield College enough money to keep it afloat but intends to call in the loan at the end of the year. He loaned money to the school for the sole reason that his son would get straight As so that he could go to Harvard Business School. That’s why Hoenicker gets angry when his son Bennett (Wil Wheaton) tells him that he’s been kicked off the basketball team because he flunked Brainard’s chemistry exam. Hoenicker sends his thugs Smith (Clancy Brown) and Wesson (Ted Levine) to Brainard’s house to change the grade. They never complete their mission, but they do return with bruises caused by something that Brainard was working on. Hoenicker decides to investigate further.
E = kT2 COLD dT N,V
Figure 6.2. Mathematic formula and chemical structure on Brainard’s computer.
170
ReAction! Chemistry in the Movies
Back in the lab, Brainard is working to harness the energy of the viscoelastic flubber with the gamma radiation. He surmises that it picks up energy from every collision so that it bounces continually harder. He places the flubber and radiation source into the engine compartment of a 1963 Thunderbird convertible and then takes the car for a flying trip. His last test occurs on the basketball court where he places some flubber on the soles of his shoes (figure 6.3). Commentary: The opening credits are fi lled with real and imaginary molecular structures interspersed with people’s names that look like mathematical formulas. This is a signal to fans of chemistry that this movie is about physical chemistry, the branch that relates molecular or bulk energy to molecular or atomic properties. After the movie begins, Brainard gives a lecture on gravity, which prepares us to think about the forces that will be overcome when flubber levitates. That’s right—flubber doesn’t fly; it levitates. Levitation happens to be one of the most amazing properties of superconducting materials, and Brainard tells us that flubber is a superconducting material. The fi rst superconductor that was functional above 77 Kelvin (the temperature mentioned in the movie and, not coincidentally, the boiling temperature of liquid nitrogen) was discovered in 1987 (Wu et al. 1987). Also according to the movie, flubber is a polymer created from a metastable fullerene. Fullerenes were discovered in 1985 (Kroto et al. 1985), and
Figure 6.3. Professor Brainard (Robin Williams) gives flubber a test run. FLUBBER © Disney Enterprises, Inc. Photo courtesy of the Mary Riepma Ross Media Arts Center.
Inventors and Their Chemical Inventions
171
superconducting fullerene polymers created from metastable fullerenes were discovered soon thereafter (figure 6.4) (Buntar and Weber 1996). During the 4-minute wedding preparation/experimentation scene that starts at 19:00, we see the critical mathematical formula that somehow controls flubber synthesis. In words, the equation reads: energy E is equal to the Boltzmann constant k times temperature T times the “heat” (represented as HOT or COLD) as a function of the change in temperature dT at constant volume V and constant numbers of particles N. It is the use of the Boltzmann constant that indicates Brainard is working on either the molecular or quantum level because it relates temperature and energy for individual molecules, atoms, or other particles. The Cooper pairs Brainard mentions were discovered by Leon Cooper in 1956 at the University of Illinois in Urbana. They are electrons that form pairs at temperatures near absolute zero due to a special type of long-range attraction (Cooper 1956) and are now called Cooper pairs. The next year, Cooper published a paper with Bardeen and Schriefer in which they developed a theory of superconductivity now called the BCS theory (Bardeen et al. 1957). The theory says that when certain materials are cooled below their critical temperature, Tc, so many Cooper pairs form that they no longer collide with one another to resist the flow of their partners. Since it was the random collisions that caused electrical resistance, the flow of the few remaining unpaired electrons are unimpeded. The BCS theory of superconductivity earned the three researchers the 1972 Nobel prize in Physics. The equation that Brainard uses in the movie is the band gap energy equation from the BCS theory, one of its most important predictions. It predicts that the band gap energy Eg is related to the critical temperature Tc according to the relationship Eg ~ 3.5 kTc. As the temperature is lowered through the Tc, the conducting voltage jumps, and this jump is called the “band gap.” The approximate value of 3.5 in the equation relates to the thermal dependence of the heat capacity, which can be written as (dC/dT)N,V, and that was fi rst demonstrated in 1956 (Corak et al. 1956). When the material is above the Tc, or HOT, dC/dT is linear and regularly conductive. When the material is below the Tc, or COLD, dC/ dT is exponential and superconductive.
3
C60 Fullerene
Figure 6.4. Fullerene polymerization reaction.
Fullerene Polymer
172
ReAction! Chemistry in the Movies
A bit of trivia is that the gamma rays used to energize the flubber were supposedly from element 106 (Balderfroid 1997; Walt Disney Pictures 1997). Element 106 is seaborgium and was named in honor of Glenn Seaborg from the University of California at Berkeley, who shared the 1951 Nobel prize in chemistry for his team’s artificial creation of five elements heavier than uranium. It seems that Seaborg visited the Flubber movie set to meet Robin Williams, the producers, and the crew on the day that they fi lmed the use of element 106. They jokingly apologized to him for using his element to break the laws of physics. Another connection to the Berkeley chemistry department is that postdoctoral fellow Jeff Cruzan served as the movie’s “science advisor” and “technical advisor” (Walt Disney Pictures 1997; Internet Movie Database 2008). Cruzan was paid to make sure that all of the glassware and instruments were hooked up correctly and that Williams’s ad-libs were scientifically and technically correct. Romy and Michele’s High School Reunion (1995) Distribution company: Touchstone Pictures Director: David Mirken Screenwriter: Robin Schiff from her 1988 stage play Ladies Room Short summary: Romy White and Michele Weinberger claim to have invented Post-it Notes MPAA rating: R Plot description: The movie opens with an aerial shot of Venice Beach, California, and then zooms east into the fourth floor window of an old building. Romy White (Mira Sorvino) and Michele Weinberger (Lisa Kudrow) are watching Pretty Woman over again because they love it so much. The next day, a customer who happens to be an old classmate tells Romy that their tenth high school reunion is coming up. That night, Romy and Michele realize they haven’t accomplished much in 10 years. They identify their biggest problems as losing weight, fi nding better jobs, and getting boyfriends. Over the next few days, they fail at all three. Romy decides they should tell everyone they are businesswomen. They happily purchase new business suits to look the part and drive to Tucson in a cool car. At the reunion, a woman says: “Wow, we have a whole class of inventors!” after she learns that Heather Mooney (Janeane Garofalo) invented the quick-burning paper in Lady Fair cigarettes: “Twice the Taste in Half the Time for the Gal on the Go.” While at Tucson’s Sage Brush High School 10 years earlier, Heather realized there was a market for a cigarette that “you can smoke all the way through between classes.” The woman then tells Heather that Sandy Frink (Alan Cumming) “invented some special kind of rubber that’s used in every tennis shoe in North America.” Both Sandy and Heather had been loners at Sage Brush and both were
Inventors and Their Chemical Inventions
173
good at science. In fact, one flashback scene shows them together at their science fair poster. On their road trip, Romy decides they’ll tell everyone she invented Post-it Notes and that Michele chose the color. Michele is insulted by her proposed contribution, so they decide to break up. In the next 2.75minute scene starting at 58:15, the famous dream sequence, Michele convincingly describes the special glue she invented for Post-its to former head cheerleader Christy Masters (Julia Campbell) and the other former cheerleaders. Commentary: This movie is held together by stories of inventors. Two of the high school loners invent utilitarian items that bring them great wealth. Since most inventors aren’t well known outside their own circles, choosing to claim invention of Post-it Notes would seem to be a perfect cover. There was no market for sticky notes (Post-it is the 3M trade name) when they were released nationwide in 1980. By 1990, they were hailed as the top consumer product of the decade. In 2004, they were among the 122 everyday items in the “Humble Masterpieces” exhibit at the Museum of Modern Art. The sticky notes story began in 1968 when 3M chemist Spencer Silver was preparing a series of acrylate copolymers in his search for a new adhesive. One of his preparations formed viscoelastic microspheres (U.S. Patent 3,691,140, 1970). When sprayed onto paper, the spheres enter the pores and allow the paper to stick to other surfaces through van der Waals contact. The spheres remain in the paper, making it removable and restickable. Silver spent the next six years trying to develop a commercial product. In 1974, 3M chemical engineer Art Fry learned about the glue at one of Silver’s presentations. He realized he could make a sticky but removable bookmark for his hymnal. Since 3M allows developers to spend 15% of their time on special projects, Silver and Fry worked together on the product, making sure the sticky bookmarks didn’t cost too much, were recyclable, and so forth. While doing so, Fry realized that you could also write on them, and they became the sticky notes we use today. Back to the Future (1985) Distribution company: Universal Pictures Director: Robert Zemeckis Screenwriters: Robert Zemeckis and Bob Gale, who thought of the idea after reading his father’s high school yearbook Short summary: At 1:20 A.M. on Saturday, October 26, 1985, Marty McFly travels back 30 years in Dr. Emmett L. Brown’s plutoniumpowered DeLorean DMC-12 automobile time machine MPAA rating: PG Plot description: Seventeen-year-old Hill Valley, California, high school student Marty McFly (Michael J. Fox) spends most of his afternoons with
174
ReAction! Chemistry in the Movies
Emmett “Doc” Brown (Christopher Lloyd). One day, he fi nds a note asking him to meet Doc that night in the mall parking lot. On the way home, a woman asks him for a donation to repair the town clock that hasn’t run since it was struck by lightning 30 years ago. At home, his mother Lorraine (Lea Thompson) retells how she fell in love with Marty’s father George (Crispin Glover) after her father ran into him with their car. Downtown, Doc shows Marty the DeLorean DMC-12 sports car he modified into a time machine. Its 1.21-gigawatt per trip (oddly pronounced as jigowatt) flux capacitor is powered by what appears to be 10 milliliters of red plutonium solution, which, we learned earlier from radio announcements, was stolen from Libyan terrorists. Doc says the car must travel at 88 miles per hour to be activated. After he shows Marty how to program the machine by setting it to November 5, 1955, the Libyans appear in a van and shoot Doc, who falls to the ground. Marty jumps in the DeLorean and speeds away until he reaches 88 mph and disappears, leaving behind flaming tire tracks. Commentary: As a young man in 1955, the independently wealthy and reclusive Doc had already designed the flux capacitor that would make time travel possible. He just didn’t have a power source strong enough to run it. A watt is the rate of using or producing energy, and one gigawatt is the same as one million kilowatts, the unit used by the energy community. Imagine the intensity and heat produced by a 100-watt light bulb and then consider that Doc needs the power of 10 million light bulbs in a volume smaller than one light bulb. How much plutonium does he need? Plutonium-239 is the plutonium isotope used as the power source for nuclear power and nuclear bombs. Even though its specific power from spontaneous decay is massive at 1.92 watts per gram (Stout and Jones 1947), you can do the calculation to fi nd out that the DeLorean would have to use most of its energy to carry the plutonium. So, he must be using the plutonium for nuclear fi ssion, where E = mc2. In 2007, the U.S. Library of Congress selected this fi lm for preservation in the National Film Registry. Caprice (1966) Production company: Twentieth Century Fox Director: Frank Tashlin Screenwriter: Jay Jayson from Martin Hale’s story Short summary: Double agent Patricia Fowler’s search for the waterrepellent hairspray formula leads her to inventor Madame Piasco Plot description: Patricia Fowler (Doris Day) works for Femina Products in Paris but arranges to be caught selling secrets so that rival May Fortune Company in the United States will hire her. She intends to steal their secret for a water-repellent hairspray that will revolutionize the market. May Fortune employee Christopher White (Richard Harris) brings her
Inventors and Their Chemical Inventions
175
to the United States and later reveals he is also a spy for Femina Products. May Fortune president Matt Cutter (Jack Kruschen) knows Fowler is a spy and has given White the task to obtain proof so he can put Femina out of business. Fowler meets with May Fortune’s Dr. Stuart Clancy (Ray Walston) and learns about his face cream, eye cream, lipstick, and moustache depilatory. Fowler sees that his secretary Su Ling (Irene Tsu) is the one who tested the hairspray, but when she visits Ling’s apartment, Ling is lying unconscious on the floor. As Fowler looks around, she discovers the hairspray in the medicine cabinet and then collects some black residue from a face powder jar on the table next to Ling. She sends it out for analysis. In the meantime, White has discovered that Fowler’s father was on the trail of a narcotics ring and was killed just after discovering that a woman ran the operation. Fowler learns that Dr. Clancy could have been a championship skier and that his children are living in Switzerland with their grandmother. Fowler travels to Switzerland and fi nds a shop selling beauty products, including the hairspray. At 1:05:15, in the backroom laboratory, she meets Madame Piasco (Lilia Skala), who is working on a new formula for eye shadow. She learned how to invent cosmetics from her grandmother, who invented a cream for children’s faces from goat milk. The water-repellent hairspray is called Caprice. She developed it for herself because she does a lot of skiing. After more plot development, the face powder is described as “not injurious to the skin but when burned it becomes a powerful hallucinogen.” It is the May Company’s biggest seller. Commentary: Practically every character in this fi lm is a double agent, including the face cream. The exception is the cosmetics inventor, who is actually a kindly grandmother and not a skiing assassin. Lover Come Back (1961) Production company: Universal International Director: Delbert Mann Screenwriters: Stanley Shapiro and Paul Henning Short summary: Dr. Linus Tyler develops the alcoholic mint VIP while advertising executives Carol Templeton and Jerry Webster compete for new accounts while engaging in a battle of the sexes Plot description: Carol Templeton (Doris Day) is the new advertising executive at the Brackett Agency in New York City. She is hard working, plays by the rules, and is naive, all of which is summed up by noting she is from Nebraska. Her rival is Jerry Webster (Rock Hudson), a boozing womanizer who works for Ramsey and Company across the street. The Miller Wax Company account is up for grabs, and they both want it. Templeton has her team study the market and develop an attractive new can in preparation for her meeting the next day. Since Webster knows that
176
ReAction! Chemistry in the Movies
Wax Company president, J. Paxton Miller (Jack Oakie), is from Virginia, his strategy is to learn more about the Civil War and whether Miller’s wife will accompany him. Cut to Webster and Miller at the Bunny Club, where the lead dancer is a Southerner named Rebel Davis (Edie Adams). Her confederate flag-patterned brassiere drives Miller wild. The next morning, Templeton arrives at Miller’s hotel room just as the all-night party is ending only to discover that Miller signed with Webster. Templeton is furious and brings Webster before the Ad Council for this unethical behavior. She is able to convince Rebel Davis to testify against him because he lied to her. Webster turns the tables by offering Davis the chance to appear in advertisements for a new product named VIP. The product doesn’t actually exist. He invented the name from a newspaper headline about V.I.P.s. During the series of commercials Davis makes for the mysterious VIP, she is sexy in the fi rst, coy in the second, pregnant in the third, and holds a baby in the fourth. Her line is that VIP got her where she is today. When the advertisements are accidentally broadcast, the public demands to know where they can buy the product. Webster decides to hire Dr. Linus Tyler (Jack Kruschen), a Phi Beta Kappa, Nobel Prize–winning chemist, to make a deodorant, shampoo, or toothpaste they’ll sell as VIP. At 1:01:15, there is an explosion fi lling Tyler’s laboratory with white smoke. Tyler tells Webster he’s made a slight mistake somewhere and then replaces the “3” on the top methyl with a “4” in the structure on the blackboard (figure 6.5). There are three more short scenes of explosions, each with different colored smoke: orange, red, and fi nally purple. After the second explosion, Tyler erases the entire right side of his compound. Webster’s boss says that Tyler “isn’t a chemist, he’s a munitions maker.” Commentary: This seems to be the only movie in which the chemist interacts with his targeted chemical structures. The explosions accompanying each synthetic attempt are the cinematic equivalent of the eureka moment of discovery, but also of failure. The chemist has learned one more thing that doesn’t work. Indeed, the molecules on the blackboard would certainly be unstable and probably explosively reactive, provided
CH3 HC CH3
C
C
C
CH4 CH C
HC CH3
CH3
H
On blackboard at 1:01:15
C
C
C
CH C
H
at 1:03:15
HC CH3
CH3
C
C
C
CH C
CH3
H
at 1:05:15
Figure 6.5. The first, second, and third compounds on Dr. Tyler’s blackboard. He erases the circled part after the third explosion.
Inventors and Their Chemical Inventions
177
they could be synthesized at all. The fi rst compound is the closest to being a real compound, except it is missing three double bonds of 1,3,5trimethylbenzene. There are many ways to create colored smoke on a movie set or stage (McCarthy 1992). It is most likely, however, they used SpectraSmoke in this 1961 movie. It came in many colors (white, red, yellow, green, blue, gray, violet, off-white, orange, and pink) and was one of the many chemical special effects created by Ira Katz, who ran Tri-Ess Sciences Inc. in Los Angeles from 1950 until his death in 2005. The Absent-Minded Professor (1961) Production company: Walt Disney Pictures Director: Robert Stevenson Screenwriter: Bill Walsh, from Samuel W. Taylor’s short story A Situation of Gravity Short summary: Professor Ned Brainard creates a flying rubber called flubber Plot description: Ned Brainard (Fred MacMurray) is a chemistry professor at Medfield College of Technology, Betsy Carlisle (Nancy Olson) is his fiancée and secretary to the president of Medfield, and Shelby Ashton (Elliott Reid) is his rival for Carlisle’s affections and an English professor from rival Rutland College. Brainard is so absent-minded he has already missed his own wedding twice. If he misses the third ceremony, Carlisle threatens to break off their relationship. In a 5-minute scene beginning at 5:00, the professor is in his garage with his gurgling apparatus. After the maid reminds him of his wedding, he begins to take off his lab smock but then takes one more notice of his lab notebook. He realizes that the mathematical sign was wrong; it was negative but should be positive (figure 6.6). He says, “Any child should know that,” adjusts his apparatus to take the sign change into account, and anxiously starts it up. After some amusing sound effects, we observe a large explosion from outside the garage. When he wakes up, he kicks a can in anger at the mess. The can slowly rises while making bubbling noises. He rolls some of the stuff into a ball, bounces it to fi nd that it rebounds with greater energy than he gave to it. He tells his dog, “We’ve discovered a new type of energy.” He rigs up an apparatus that will deliver gamma rays to the flying rubber, or flubber, so that it will levitate. In a short scene beginning at 29:00, Brainard explains to his now ex-fiancée Betsy Carlisle that he has always been thinking of magnetic
E = H + P —C=C—C=C—
Figure 6.6. Mathematical formula on blackboard and chemical structure in Brainard’s notebook.
178
ReAction! Chemistry in the Movies
energy but he should have been thinking of repulsive energy. He’s certain flubber can solve all of the college’s fi nancial problems. Alonzo P. Hawk (Keenan Wynn) is a local real estate mogul trying to buy Medfield College so he can close it. He gets angry when his son Biff (Tommy Kirk) tells him he can’t be on the basketball team because he flunked Brainard’s chemistry exam. So, Hawk sends his thugs to rough up the professor. Before they can, they accidentally discover the flubber and report back. In the meantime, Brainard realizes he can demonstrate the utility of flubber by helping the Medfield Squirrels basketball team win their big game against the nearly unbeaten Rutland team. Commentary: Flubber is an elastic plastic that picks up kinetic energy with every bounce. This is a violation of the First Law of Thermodynamics, which states that energy is conserved. Professor Brainard’s notebook contains the shorthand formula for butadiene, the earliest synthetic rubber (figure 6.7). In the early 1930s, Germans developed a butadiene and styrene synthetic rubber copolymer called Buna S that remains the most abundantly produced synthetic rubber even today. The United States didn’t enter the synthetic rubber industry until the Japanese took control of 90% of the world’s rubber tree plantations in 1942 during WWII. A great variety of rubber types were produced in a short period of time. Today, more rubber is synthetic rather than natural due to the greater selectivity for properties and lower cost. Brainard’s theoretical breakthrough involved changing a minus to a plus in a simple mathematical equation: H = E + P. The only other clue on the blackboard is a more a complex equation: √E = CDg. These equations are difficult to decipher without knowing the meaning of the terms. The most obvious guess is that E is for total energy, H is enthalpic heat (or thermal energy), P is pressure, and the subscript g is for gravity. The meanings of C and D are obscure although C often means constant. The problem with these assignments is that pressure doesn’t concur with Brainard’s statements about repulsive and attractive energy, the two types of energy involved in the formation of a covalent bond. The attractive
H
n
H C
H
C
H
H C
H
C
R
H C
C C
H H
Butadiene (synthetic rubber monomer)
H
H
C
R H
n
Polybutadiene (one type of unit in the polymer)
Figure 6.7. Structural equation for butadiene polymerization.
Inventors and Their Chemical Inventions
179
force is between the electrons from one atom and the nucleus of its bonding partner. The repulsive force is between the nuclei of both atoms and between the electrons of both atoms. The use of this language provides a chemically based metaphor for flubber’s ability to pick up energy with every bounce in that it seems to be repulsed by every encounter. The phrase “absent-minded professor” is still being used today to describe professors from all academic disciplines. Even though the term was coined long ago, this movie made such an impression (it was the fourth highest grossing fi lm of 1961; the sixth was Lover Come Back) that it is now closely associated with it. It turns out that the character of professor Ned Brainard (probably based on brain + nerd; the word “nerd” was coined in the 1950s) was inspired by the real-life Princeton University chemistry professor Hubert Alyea (pronounced All-yay) (Harmon 1996; Shakhashiri 1997). Walt Disney was in the audience when Alyea was performing demonstrations at the science pavilion of the Brussels World’s Fair in 1959. Disney liked Alyea’s fast pace and zany persona so much that he decided to create a movie featuring a character who acted like him. Alyea was invited to Hollywood and performed his chemical demonstrations for Fred MacMurray, who then used what he observed as the model for his character. The theory behind flubber within the movie’s scenario was developed by El Camino College physics professor Dr. Julius Sumner Miller of Torrance, California. Miller also supervised the lab sequences in this fi lm and in its sequel Son of Flubber (1963). This led to his formal relationship with Disney Studios when he became Professor Wonderful on The Mickey Mouse Club television show in 1964 and 1965. He would later host a television show in 1960s Australia and in 1970s Canada that featured physics demonstrations followed by comprehensible explanations. Yellow Cab Man (1950) Distribution company: Metro-Goldwyn-Mayer Director: Jack Donohue Screenwriters: Albert Beich and Devery Freeman Short summary: Accident-prone Augustus “Red” Purdy invents an unbreakable elastic glass Plot description: Augustus “Red” Purdy (Red Skelton) is an accidentprone inventor of safety products. In the opening sequence, Purdy is hit by a Yellow Cab but won’t press charges because he knows it was his fault. He was holding an enormous cuckoo clock and didn’t look when he stepped onto the street. Claims Adjuster Ellen (Gloria DeHaven) arrives from the Yellow Cab Co. to settle the account with him. She wants him to sign a form, but he begins talking about his inventions when lawyer Creevy (Edward Arnold) enters to say that Purdy shouldn’t sign a thing. In the short scene at 7:30, Purdy says the cab company would benefit
180
ReAction! Chemistry in the Movies
from making their windshields with his elastiglass. He bounces a flask made with elastiglass against his door and it bounces back. Creevy asks whether it has been patented. Purdy says no, but it’s safe in his head. Ellen says she’ll arrange for Purdy to demonstrate his product to the company president if only he’ll sign the release form. In the 4.25-minute scene at 10:15, Purdy demonstrates the elastiglass for president Pearson Hendricks (Paul Harvey), but Creevy convinced one of the employees to switch it with real glass. With the president behind the wheel of a car, Purdy pitches a baseball straight at him that breaks the windshield and beans him. Red speculates that sometimes he bakes it too fast and the glass gets brittle. The injured president pummels Purdy so badly he ends up in the hospital, and he demotes Ellen to the Lost and Found. As the plot develops, Purdy becomes a cab driver of car 1313, his lucky number. Lawyer Creevy schemes with other men to get the formula from Purdy. In the 5-minute scene at 58:30, they use narcosynthesis to get to the truth. He is told to revert to the fi rst time he thought of elastiglass formula. In his flashback, he is a baby sharing a crib with his baby sister, who konks him on the head with a bottle. “Later, I read an excellent book on chemistry.” He is dressed as a little boy and pours solutions together. He wanted to be as famous as Edison, Pasteur, Madame Curie, and even the Wizard of Oz. “Years later, I moved into a larger basement laboratory. I mixed silica [SiO2], lime [CaO], potash [K 2CO3], and lead oxide.” He pours the solution into the flask, and it explodes to create blown glass out of the flask. “Later, I fused the silica with alkali and made elastiglass.” Commentary: There are three glass stories that relate to this movie. First, Purdy correctly described the manufacture of leaded glass in the movie. When inexpensive silica sand (silicon dioxide), soda ash (sodium carbonate), and limestone (calcium carbonate) are heated, the components melt. As the molten material is slowly cooled in a process called annealing, the ions arrange themselves to form a transparent brittle ceramic called glass. When some of the calcium oxide is replaced by lead oxide, it is called leaded glass. The second story is the serendipitous discovery of safety glass in 1903. It began when French chemist Edouard Benedictus dropped a flask on the floor of his lab and it shattered but retained its shape (Roberts 1989). He was curious, so he asked his technician and found it was the flask that had been fi lled with nitrocellulose solution. The water had evaporated and left a fi lm inside that held the glass shards together. He recorded his observations in his notebook and forgot about it. Within the next few weeks, he read two news reports about people getting cut by windshield glass in automobile accidents. After the second one, he rushed to his lab to do the fi rst purposeful experiment. Over the next six years, he perfected the process and patented it. Safety glass consists of a sheet of nitrocellulose between two glass sheets that are heated until they form a clear unit. It became standard equipment on the now faster American cars of the 1920s. In 1933, cellulose acetate replaced nitrocellulose that
Inventors and Their Chemical Inventions
181
yellowed in the sun. In 1939, poly(vinylbutyral) replaced cellulose acetate that became opaque as it aged. The fi nal story is about Plexiglas, which appears to be the material Purdy holds in the movie. Plexiglas was fi rst marketed in 1933 by the Rohm and Haas Company and was still a novelty in 1950. Today, acrylic plastics are used as substitutes for glass in eyeglass lenses, aircraft windows, automobile tail lights, and lighting fi xtures. They don’t shatter like glass but are more easily scratched. Edison, the Man (1940) Distribution company: Metro-Goldwyn-Mayer Director: Clarence Brown Screenwriters: Bradbury Foote and Hugo Butler Short summary: Biography of Thomas Edison that builds up to the invention of the light bulb Plot description: The fi lm begins in 1929 at the 50th anniversary celebration of the invention of the light bulb, with Thomas Edison (Spencer Tracy) as the honoree. A speaker tells the story of Edison’s life as a flashback, which begins before 1879 when the light bulb was invented. Edison arrives in Boston as a young man and learns how to operate and repair telegraphs. In his spare time, he reads about Faraday’s experiments in electricity and chemistry. Edison’s fi rst break occurs when he sells his ticker machine patent for $40,000 to General Powell (Charles Coburn) and Mr. Taggart (Gene Lockhart). He uses the funds to marry, build a house and laboratory in New Jersey, and fi le a number of patents that are named but not described, such as the electric pen. He and his wife soon have a daughter and son, Dot and Dash. After five years, Edison is broke so he fi nally decides to work on the light bulb to make some quick money. After four days of continuous work alone, he gets nowhere. Powell is near death so he approaches Taggart for money. He offers him $100,000 and asks for only one thing in return, to be able to direct Edison’s attention in the most profitable directions. Edison refuses to accept this condition and returns home empty-handed. The next day, a machine runs fast and makes a noise that causes Edison to ponder. In short order, he invents the phonograph and it is a success. All too soon, though, the money runs out again. This time he gives the men their severance pay and sends them on their way. The next day, he fi nds them working in the lab when he arrives, and they develop the light bulb as a team. In the 1.5-minute scene at 1:14:00, Edison and his workers search for the perfect material to illuminate their light bulb. They test dozens of materials, such as tin, copper, iron, iridium, brass, and nickel. As the narrated list trails off, the scene of workers at the bench is double exposed with a printed list of materials in a notebook being crossed off with pencil,
182
ReAction! Chemistry in the Movies
such as manganese, sulfur, lead, mercury, antimony, bismuth, birch, ash, and oak. Then, all the workers gather round as platinum is tested. It gives a brief glow but burns out like the rest. At home that night, he speaks to his wife of a friend’s stupidity. Edison describes him as having “a vacuum in his head” and then exclaims, “Too much oxygen!” and races back to the lab. They borrow a mercury vacuum pump from Princeton, but it doesn’t solve the problem completely. They’ve now tested 9,000 things that don’t work. The next day, Edison decides to knead some carbon putty into sewing thread and have it heated to create a carbonized fi lament. It glows more than 40 hours in the vacuum bulb, and their search is over. The next task is to create a power-generating station in New York City so they can light up an entire block. Commentary: More than 125 years after its discovery, the electric carbon fi lament light bulb is still the symbol of bright ideas. Its creation by Edison and his research team provides an excellent primer on Edison’s methods (Livesay 2007). First, Edison and the senior members learned everything they could about previous attempts to create incandescent light. They shared what they discovered with each other and then brainstormed improvements and new ideas. Next, they built working models of the most likely possibilities so they could examine and compare them. After settling on the use of wire that glowed on electrification, it became clear that few materials besides platinum had been tested, so they gathered and tested many materials in roughly the following order: metals, metal oxides, and cellulosic materials such as bark, cotton, silk, facial hair, cardboard, and bamboo. The cellulosic material was kneaded with a carbon putty (ground carbon black plus tar) and then heated to drive off hydrogen and oxygen, leaving behind carbonized fi laments (Jehl 1936 and 1938). Bamboo proved to be the best material because it has especially long fibers. In 1878, Edison sent William Moore to Japan to fi nd the widest variety of bamboo to test from among the 200 known species. It took Moore two years to visit the four main islands of Japan to gather the samples he mailed home. The best one was discovered in a grove near the Iwashimizu-Hachiman-Gu Shinto Shrine in Yawata, Kyoto. The Edison group called it Moore Bamboo. Edison arranged to have it farmed for use in the light bulb. Even though this bamboo was never bettered, the search never ended. Edison sent other men on bamboo expeditions through the years to Cuba, Brazil, South America, and then on a world tour of tropical zones. Although this is supposedly a biography of Thomas Edison, very few of its biographical details are based on reality. All characters have fictional names except Edison, his wife, and their two children. Most characters in the movie, including his wife (he remarried after his fi rst wife died), are composites or fictional. The moviemakers must have decided it was best to present an engaging story rather than attempt to present the known facts engagingly. The two “technical advisors” listed in the opening credits were probably responsible only for showing the actors how to work the
Inventors and Their Chemical Inventions
183
machinery: They were William A. Simonds from the Edison Institute in Dearborn, Michigan, and Norman R. Speiden, Director of Historical Research at Thomas A. Edison, Inc. in West Orange, New Jersey. Beauty for the Asking (1939) Distribution company: RKO Radio Pictures Director: Glenn Tryon Screenwriters: Doris Anderson and Paul Jarrico, from a story by Edmund Hartman Short summary: After Jean Russell loses her boyfriend and job, she decides to market the face cream she invented Plot description: Cosmetician Jean Russell (Lucille Ball) develops an astringent cream in her apartment after work with the help of her cosmetician friend and roommate Gwen Morrison (Inez Courtney). In the one-minute scene at 5:15, Russell explains her facial cream formula to Morrison: boil lanolin and add a dash of rose water. While they fi ll jars with the cream and dream of getting rich, Russell says she feels like an alchemist: “This stuff is magic—it will turn everything to gold for Denny and me.” She spoke too soon because her cosmetics salesman boyfriend Denny (Patric Knowles) dumps her the next day to marry millionairess Flora Barton-Williams (Frieda Inescort). Then, she and her roommate lose their jobs. Morrison convinces Russell to focus on her face cream, so she visits the largest ad agency in town. The young president, Mr. Jeff Martin (Donald Woods), says it will cost a quarter of a million to market the cream but decides to help her anyway because he’s attracted to her. First, he has a sculptor create a fancy jar for the cream because the one she was using “is about as attractive as a can of zinc oxide.” Then, he fi shes for investors by sending samples to 12 leading ladies of New York society. Finally, he creates a public image for Russell as Contessa Jeanne de la Varelle. BartonWilliams takes the bait because it will be perfect for her new husband, the former cosmetics salesman. At their meeting, everything turns out well even after Russell reveals she isn’t really a contessa. The next scene shows a magazine advertisement from the Jeanne Varelle Company. Commentary: An “astringent cream” is an oxymoronic joke most women would understand but many men would not. Moisturizing cream is a semisolid applied to soothe the skin after one has cleaned it and added an astringent to close the skin pores. The moisturizer traps moisture and incidentally causes the pores to open. In this movie, the moisturizing cream is lanolin and the astringent is proposed to be rose water. Lanolin is the fat obtained by scraping sheep’s wool and has long served as the base for cold cream preparations. Rose water is the vapor distilled by heating the flowering parts of roses in water, a procedure that was presented very dramatically in the recent movie Perfume: The
184
ReAction! Chemistry in the Movies
Story of a Murder (2006). It is normally used not as an astringent but as the base for fragrances. There are quite a number of parallels between Jean Russell (Lucille Ball) in Beauty for the Asking and the real life story of Harriet Hubbard Ayer (Peiss 1998). Ayer was born in 1849 to a prosperous Chicago family and married the son of a prosperous iron dealer when she was 16 years old. They divorced after 21 years of marriage, causing Ayer to take various jobs before moving to New York with her youngest children. Once there, she began manufacturing a face cream named Madame Recamier, after a famous Napoleonic beauty. With endorsements from her society friends, it became an immediate success and her house was soon full of women helping her to produce it as fast as possible. Mr. Edison at Work in His Chemical Laboratory (1897) Distribution company: Edison Studios Producer: James White Photographer: William Heise Plot description: The camera is fi xed in one position as Edison moves from one side of the screen to the other. He weighs something, adds it to a flask, and pours the solution into another flask. Commentary: This was the 344th fi lm produced by the Edison Studios and, according to the 1898 catalog entry to sell the fi lm, “The scene is an actual one, showing Mr. Edison, in working dress, engaged in an interesting chemical experiment in his great laboratory” (Musser 1991, 1997). It is a 20-second kinetoscopic movie consisting of a series of photos printed onto cards attached to a large-diameter cylinder. The viewer paid a nickel to peer through a peephole in the top of the machine while turning the crank on the side. As the cards fl ipped past a light bulb, the viewer perceived motion. Even though the advertisements for this movie and its title suggest Edison was fi lmed in his own laboratory, he was actually in Black Maria studio. It was built in 1893 under the direction of photographer William Dickson and was the fi rst studio built specifically for making movies. Edison’s bench and supplies were brought from the adjacent laboratory building for this historic recording. The fi rst successful motion photographs were made by Eadweard Muybridge in 1877, a British photographer working in California. His solution was to take a series of photographs of an active animal or person using a row of cameras with strings attached to their shutters. In his fi rst series, a running horse broke each string in succession, tripping the shutters, and capturing motion in a simple, accessible way. His photos inspired showmen and inventors to create the field of motion photography. Edison’s vision was to produce a movie machine designed like his phonograph. He called it the kinetoscope, and beginning in 1890, his company produced hundreds of 20- to 90-second movies with a wide
Inventors and Their Chemical Inventions
185
variety of themes. These novelty items were soon replaced by projection units that generated wall-sized images that could be viewed by many people at once. Beginning in about 1896, Edison’s company aggressively developed and manufactured moviemaking and projecting apparatus, even some for which it did not hold a patent (Bowser 1990). In 1908, Edison cajoled the other major movie producers to pool their patents and form the Motion Picture Patents Company (MPPC). This allowed them to control production, distribution, and exhibition in the United States by charging fees. In 1917, the U.S. Supreme Court ruled the MPPC was a monopoly and had to be dismantled, causing Edison and the others to leave the movie business altogether. On the positive side, the MPPC created an industry standard for fi lm size and projection apparatus that brought coherence to this market.
7 Hard Science = Hard Evidence Forensic Chemistry and Chemical Detectives
EVERY CONTACT LEAVES A TRACE Someone killed the mayor of Chatsberg using a makeshift pipe bomb in Kid Glove Killer (1942). It was connected with a wire to the electrical system of the mayor’s automobile and exploded when he turned the ignition key. The two members of the Chatsberg police forensic team, supervisor Gordon McKay (Van Hefl in) and his assistant Jane “Mitchell” Mitchell (Marsha Hunt), visit the crime scene to collect clues such as a coat fiber caught in the garage door, the exploded bomb fragments with attached wire, and the burlap bag beneath the automobile. Their hypothesis is never stated, but it is clear from their actions they are using Locard’s exchange principle: “Every contact leaves a trace.” McKay and Mitchell focus their subsequent efforts on fi nding physical evidence on suspects that connects them with the material collected at the crime scene. In forensic science, crime scene reconstruction is the “thought experiment,” the documented crime scene is the “effect,” and the evidence is used to establish “cause.” In the case of murder, “cause” equates to who did what to whom. An examination of the chemistry in detective and spy movies (table 7.1) shows they fall into the two main categories of elemental analysis or qualitative analysis. Most qualitative analyses fall under the category of forensic toxicology, the identification and quantification of drugs and poisons. Toxicology also happens to be one of the oldest branches of forensic chemistry. None of the movies shows the creation or even improvement of a chemical procedure. Instead, chemistry plays an infallible supporting role in solving a crime or mystery. For instance, in Kid Glove Killer, elemental analysis determines that vanadium was present as a tracer in the gunpowder but is not under the fi ngernails of the prime suspect, thereby decisively eliminating him as a suspect in the eyes of the forensic experts. The actual proof linking the killer to the crime scene does not involve chemistry but, rather, repetitive routine testing coupled with good guesswork. Even though the ability to detect arsenic in body tissues in 1815 gave chemical forensics its start (see “Limits of Detection,” below), the next advances in forensics were philosophical. Body identification and the 186
Hard Science = Hard Evidence
187
Table 7.1. A list of selected movies featuring forensic chemistry or chemical detectives Title (Year)
Chemical Procedure
The League of Extraordinary Gentlemen (2003) Merci pour le Chocolat (2000) Backdraft (1991) Batman (1989) The Name of the Rose (1986) Moonraker (1979) The Seven-Per-Cent Solution (1976) The Andromeda Strain (1971) Our Man Flint (1966) D.O.A. (1949) Kid Glove Killer (1942) Exploits of Elaine (1914)
Qualitative analysis Toxicology Qualitative analysis Toxicology Toxicology Toxicology Qualitative analysis Elemental analysis Toxicology Toxicology Elemental analysis Gadgets
exchange principle are simple in concept but have grown ever more powerful as technology has lowered the limit of detection to single molecules. Specifically, criminal and victim identification began with body measurements and then quickly moved to fi ngerprints, where it remained for a century. It has now moved inward to DNA sequence analysis with its presumably immutable chemical features. DNA sequence analysis is able to detect single molecules of DNA in hair, blood, sweat, semen, and sloughed skin. At the crime scene, there are investigators who specialize in photographing and documenting the layout of crime scenes, and others who specialize in collecting evidence. In the criminalistics lab, chemists trained as special investigators analyze the material. Using Locard’s exchange principle, each piece of evidence is examined for traces, or clues, that might link it to the victim, the suspected perpetrator, or someone else. Investigators have to imagine the crime event that plausibly took place to create the crime scene as they consider each piece of linked evidence and each witness statement (Chisum and Turvey 2000). All evidentiary links and corroborating statements have to make logical sense during the reconstruction. The reconstruction can reveal evidentiary holes, unaccounted time gaps, and alternate hypotheses. These in turn might suggest new suspects or investigative avenues. In the language of detective fiction, crime scene reconstruction is the search for the clue that doesn’t fit, and only the prepared mind is capable of making that discovery. Forensic science began in the early 1800s when the fi rst police forces were formed in Europe’s two largest cities, London (population 959,000 in 1801) and Paris (population 546,000 in 1801). Imagine a time before driver’s licenses, social security numbers, house addresses with street numbers, credit loans, and cell phone numbers that all establish your identity. This was a time when local militia enforced the laws. The
188
ReAction! Chemistry in the Movies
militia usually had close family and political ties to the area’s wealthiest inhabitants. They often kept order and protected property through brutal suppression of a region’s entire working class rather than by punishment of individual wrongdoers. Police forces developed in cities because there was no single strong local authority but rather many business and property owners. City life was subject to civil law and allowed greater individual freedom. This freedom meant wrongdoers could easily make themselves invisible by moving to a part of the city where no one in the neighborhood knew them. Most crimes were solved not by detection but through reliance on informers and undercover police. This changed only after Alphonse Bertillon created the fi rst system of criminal identification in the 1880s. By proving the value of a logical identification system, Bertillon opened the door for many others to develop the methods to link physical evidence with suspects and to develop the logic of crime scene reconstruction. Alphonse Bertillon, the so-called father of scientific detection, was born in Paris in 1853 (Rhodes 1968). His father was a physician with a strong interest in physical anthropology, and his mother was the daughter of the leading anthropologist. A few years prior to Bertillon’s birth, his father had published a paper showing it was possible to distinguish groups of people by simply comparing their heights. By age three, Bertillon was using his father’s calipers to measure ribbons and other things in their house. From this, he gained an instinctive feel for statistical measurements, but he never understood them theoretically or mathematically. He was never a good student and, in fact, especially disliked history and mathematics. In 1879, at age 26, Bertillon’s father found him a job hand-copying forms for the Paris Police. He soon learned, however, that no one used the forms after they were fi lled out because heights were often listed as “average,” looks as “ordinary,” and the photographs looked as though they were taken in a dark basement. Over the next eight months, Bertillon worked out a system of 11 body measurements he believed could distinguish any two individuals. The measurements included length of the head, breadth of the head, length from the left elbow to the end of the left middle fi nger, and so on. When the prefect of police didn’t appreciate the value of his system, the dismayed Bertillon showed his proposal to his very concerned father. After a quick glance, his father could see it was brilliant, so he suggested Bertillon wait for the installation of the next prefect. In the meantime, his father shared it with his anthropology friends, who received it quite favorably. In 1881, Bertillon was promoted and given the chance to prove his system was useful. He measured every arrested suspect, attached their photos, and fi led the cards according to one of 81 groupings. The groups were fi rst separated by head length as large, medium, or small, then head width the same way, followed by left middle fi nger length, and left little fi nger length. He defi ned large, medium, and small for each measurement
Hard Science = Hard Evidence
189
so that equal numbers of cards were in each group. After two months of making measurements and fi ling cards, he measured a suspect whose measurements were already in the system but under a different name. Sure enough, the photo on the card matched the suspect, who then confessed to using a false name—both times. By the end of six months, Bertillon caught seven more recidivists. As the number of cards increased, the frequency of recidivism became even more apparent (Bertillon 1893). By 1888, his method was called Bertillonage and was being used by police departments around the world. When he became chief of the Judicial Identity Service, he used his new position to improve the photographs so that they used the same lighting, the same distance to the camera, and the same forward and side head angles. Of all his innovations, his mug shot procedure has endured. Next, he developed a metric photographic system for use at crime sites. In 1896, he reconstructed the exact method by which two criminals had broken into a desk. He wrote down all of his reasoning and conclusions, no matter how obvious. His reconstruction matched one of the confessions exactly. By 1900, he was analyzing fi ngerprints found at crime scenes. Many others had suggested their utility in crime detection and even developed methods to collect and categorize them (Cole 2001), but Bertillon was the fi rst to use them to solve an actual case in 1902. He brought some broken glass back to his laboratory and found a way to photograph and enlarge the fi ngerprints so they could be studied and compared. Methods to dust the prints or to visualize them with reactive chemicals hadn’t yet been standardized. Even though Bertillon didn’t get along with many people, his closest associates were the forensic science duo in Lyons, Dr. Jean-Alexandre Lacassagne and Dr. Edmond Locard. Lacassagne had a medical degree and was 10 years older than Bertillon. After training in Paris a few years ahead of Bertillon, Lacassagne became the fi rst director of Laboratory of Police in Lyons in 1884. He was gregarious and trained many men in the methods of Bertillonage and other forensic technology. Edmond Locard was considerably younger than the other two men and had earned degrees in medicine and law at the University of Lyons. He became Lacassagne’s assistant, and then director of the Lyons forensic laboratory in 1912. His major research areas were fi ngerprints, dust, and ashes. During WWI, his dust detection methods were used by the French Secret Service to determine where soldiers and prisoners had passed. A bit of sleuthing reveals that Reginald Morrish is credited with coining the popular version of Locard’s exchange principle in 1940 (Bisbing 2004). Morrish stated the principle succinctly: “Every contact leaves a trace.” Locard’s original undogmatic version was more verbose, allowing for less misinterpretation. In Locard’s second book, Manuel de Technique Policiére [Manual of Police Techniques] (1923), he began chapter 3, titled “Traces,” with this sentence: “It is impossible for the lawbreaker to act, above all with the intensity that a criminal act presupposes, without leaving traces of having been on the scene.” Seven years later, in a journal
190
ReAction! Chemistry in the Movies
article on dust traces, he wrote it differently: “For the microscopic debris that cover our clothes and bodies are the mute witnesses, sure and faithful, of all our movements and of all our encounters” (Locard 1930). He was inspired by two writers to study dust: judge Hans Gross and novelist Sir Arthur Conan Doyle (Locard 1930). Gross was a judge in Graz, Austria, who wrote the 1893 book Criminal Investigations, which was reprinted and translated many times. Gross argued for systematic investigation and analysis of evidence. He disliked the heavy reliance that juries placed on often unreliable eyewitnesses. About Sherlock Holmes, Locard wrote: “I hold that a police expert, or an examining magistrate, would not fi nd it a waste of his time to read Conan Doyle’s novels.” In fact, Holmes used specks of mud to detect the visitor’s route in Conan Doyle’s fi rst story, A Study in Scarlet (1887). Locard’s studies of fi ngerprints, handprints, footprints, and so on, and the many types of dust and ashes influenced an entire generation of detectives. His exchange principle remains the central principle of crime scene reconstruction.
THE SIGN OF THREE: DUPIN, HOLMES, KENNEDY The majority of movies described in this chapter were adapted from published detective stories, the most popular form of literature. They have suspense, mystery, crime, action, and drama. Some of these stories are designed so that the reader tries to ferret out the most important clue ahead of the fictional detective. In this competition, readers compare the preparedness of their minds with the sleuth’s. This genre is also known for its character development and narrative digressions. The crime is always less interesting than the interpretation used to solve it. These rules were present in the first detective story, written by Edgar Allen Poe in 1841. His Parisian detective, C. Auguste Dupin, used his comprehensive knowledge and razor-sharp thinking to solve crimes in three short stories. Sir Arthur Conan Doyle has probably done more than any other author to portray forensic science as a valuable tool through his detective fiction (Snyder 2004). In his first novel, Sherlock Holmes developed a new chemical method for blood detection. In Conan Doyle’s second novel, Holmes’s addiction to cocaine bookends the story. He solved most of his crimes with his unparalleled ability to parse clues for their proper meaning. Arthur B. Reeve completes the trio. His scientific detective Craig Kennedy was called the American Sherlock Holmes and played an important role in the history of the movies. Kennedy was a chemistry professor at the University in New York City and relied on his gadgets to solve crimes.
Edgar Allan Poe, C. Auguste Dupin, and the Unnamed Narrator Edgar Allan Poe wrote the fi rst detective story (Haycraft 1984; Silverman 1991). It is a curiosity that someone better known for the macabre created
Hard Science = Hard Evidence
191
the fi rst hyperrational detective. The best explanation seems to be that he wrote his detective stories when he was at the height of his career and his personal life was at its most stable. In 1827, at age 18, he anonymously published his fi rst collection of poems and embarked on a career as one of America’s fi rst professional writers. At age 26, he married his cousin, who was half his age. They enjoyed seven years of marital bliss living with her mother/his aunt before his wife contracted tuberculosis and slowly died of consumption during their last five years together. The height of Poe’s career occurred when he became editor of Graham’s Magazine in 1840. His poems, short stories, and literary criticism had earned him a very good reputation. Under his hand and with his contributions, its circulation rose in just a few months from 5,000 to 40,000, the largest of its time. In that magazine in April 1841, he published The Murders in the Rue Morgue, the world’s fi rst detective story. According to detective fiction critic Howard Haycraft (Haycraft 1984), detective stories are about the professional detection of crime and could only have appeared after the creation of the real thing. Poe loved everything French and Gothic and appears to have used the descriptions of Paris streets and the police force from the 1829 published memoirs of François Eugéne Vidocq, the fi rst head of the Sûreté Nationale (National Security), the investigative branch of the French police force. According to Haycraft, the Dupin stories (two novellas and a short story) exemplify the three types of detective fiction. The resolution of the fi rst story depends on physical evidence. The second, The Mystery of Marie Rogêt (1842–1843), involves purely mental reasoning. The third, The Purloined Letter (1844), is balanced in its use of both reasoning and physical evidence to solve the crime. The Murders in the Rue Morgue (1841) begins with a discussion of the mental faculties engaged while playing chess, checkers, or whist. The reader is then asked to read the story’s narrative “in the light of a commentary upon the propositions just advanced.” The unnamed narrator begins by describing how he met C. Auguste Dupin, who was from a once-prominent Parisian family but was now living on a small stipend. “Our fi rst meeting was at an obscure library in the Rue Montmartre, where the accident of our both being in search of the same very rare and very remarkable volume, brought us into closer communion.” The narrator realizes Dupin is likely to be a singularly interesting individual and proposes to pay for their common lodgings. After living together for a while, they read an extraordinary tale of double murder in the newspapers that has the police baffled. A mother and daughter were murdered in their own apartment, but the door and windows were locked from the inside, and the chimney was too small even for a cat. The published interviews with neighbors agree they heard a gruff French voice and a shrill voice inside the apartment. The interviews do not agree on the language spoken by the second voice—the Englishman was sure it was a Spaniard even though he doesn’t speak Spanish himself, the Spaniard
192
ReAction! Chemistry in the Movies
thinks it was English but has no knowledge of English, and so forth. They visit the scene of the crime, where Dupin spends half the day examining the premises. The next day, he explains the true facts that can be gleaned from the newspaper reports, but before he goes further, he says, “It is only left for us to prove that these apparent ‘impossibilities’ are, in reality, not such.” He then describes his fi ndings at the crime scene and interpolates the only possible reconstruction of the crime. The story ends after they meet with the only person who observed the crime, who provides the otherwise unknowable specifics of the case. The Mystery of Marie Rogêt (1842–1843) was serialized in three issues of Snowden’s Ladies’ Companion. The story concerns a “perfumery girl,” Marie Rogêt, who is found dead floating in the Seine River in Paris. Dupin and the unnamed narrator read the sensational newspaper accounts like everyone else, but unlike everyone else, Dupin reasons his way to the truth. First, he describes the false assumptions and incorrect interpretations put forward in the papers. Then, he uses the remaining facts to reconstruct the “supposed scene of the assassination.” He ends with suggestions for further detective work. The footnotes to the republished version of the story, in Poe’s Tales (1845), indicate the narrative was based on a real crime that took place in New York City in 1841. When Mary Rogers’s body was found floating in the Hudson River, the newspapers covered the case daily and referred to her as the “Beautiful Cigar Girl.” Poe believed the authorities bungled the case and wrote this story to prove his point. The Purloined Letter (1844) begins with a visit from the prefect of the French police. He describes the theft of an incriminating letter from the queen by one of her ministers. The police have every reason to believe it is in his apartment and not on his person. A thorough search of every hidden inch of the apartment, however, did not reveal the letter. When the prefect visits him one month later, Dupin hands him the purloined letter in exchange for a portion of the reward. Dupin then tells the unnamed narrator how he reasoned the letter must be in plain sight. So, he wore a disguise while visiting the minister in his apartment, spied the location of the letter, and returned to exchange it for a facsimile.
Arthur Conan Doyle, Sherlock Holmes, and Dr. James Watson Arthur Conan Doyle loved sports and read a great deal of fiction when he was a boy in Edinburgh (Hingham 1976); he especially enjoyed reading tales of chivalry and adventure but was bowled over by Poe’s The Gold Bug and The Murders in the Rue Morgue. While earning his medical degree at the University of Edinburgh, he became an assistant to a noted local surgeon, Dr. Joseph Bell, who gave occasional demonstrations at the university. As patients entered the stage for examination, Bell would observe them briefly, deduce their occupation and so forth, and then describe to the class how he came to his conclusions. It was during this time that
Hard Science = Hard Evidence
193
Conan Doyle began writing adventure novels. After his graduation in 1882, he set up a medical practice in a suburb of Portsmouth in southern England. He married, had a modest practice, and spent his spare time writing more adventure novels. After three years of medical practice, he wrote his fi rst detective story, A Study in Scarlet (1887), featuring Dr. John Watson and Mr. Sherlock Holmes. It was published in the premiere issue of Beeton’s Christmas Annual and enthusiastically received by the public. Keep in mind that Robert Louis Stevenson’s gothic detective tale about Dr. Jekyll had been published the previous year. They were still talking about it because Jack the Ripper was in the midst of his murder spree in London, and many felt there were uncanny similarities between the two. To read about a detective who could decipher the motives behind bizarre crimes would have been reassuring. In total, Conan Doyle wrote four novels and 56 short stories featuring Holmes and Watson, plus many other fictional and nonfictional adventure stories. A Study in Scarlet opens with Dr. John Watson giving a thumbnail sketch of his service in Afghanistan as a medical doctor. He received a bullet wound to the shoulder and then contracted malaria. He returned to London but had virtually no income, so he began searching for cheaper lodgings in 1881. He ran into an old friend, who told him someone named Sherlock Holmes had found suitable lodgings to rent but couldn’t fi nd anyone with which to share them. When Watson is introduced, Holmes is holding a test tube. He joyously explains the new chemical test he developed to identify human blood, the scarlet of the title. (Dye chemist and Holmes scholar Samuel Gerber [1983] proposed several possibilities for Holmes’s test. He also noted, however, that the species of the blood’s origin couldn’t be identified until Dr. Karl Landsteiner discovered the A, B, AB, and O blood types.) Watson and Holmes agree to share the lodgings at 221B Baker Street. Instead of asking about Holmes’s occupation, Watson tries to figure it out by cataloguing Holmes’s knowledge base. For instance, he writes that Holmes has “profound” knowledge of chemistry but “nil” of astronomy. Of botany, he has extensive knowledge of poisons but none of practical gardening. Watson can’t figure it out until Holmes explains that he is a consulting detective. After that, the game is afoot in this strange tale of murderous polygamist Mormons. Conan Doyle’s second novel, The Sign of Four (1890), opens with Holmes readying to inject a seven-per-cent solution of cocaine as Watson asks him, “Why should you, for a mere passing pleasure, risk the loss of those great powers with which you have been endowed?” Before he is able to answer, a client enters with a case. It is a locked-room story like Poe’s The Murders in the Rue Morgue. When Holmes realizes the room was locked, he rhetorically asks Watson, “How often have I said to you that when you have eliminated the impossible, whatever remains, however improbable, must be the truth?” After the case has been solved and the novel ends, Holmes prepares to inject himself with the cocaine solution.
194
ReAction! Chemistry in the Movies
In The Hound of the Baskervilles (1901), the client says Holmes is the “second highest expert in Europe.” Holmes is surprised and asks who is fi rst. The client says, “Bertillon.” The hound of the title glows in the moonlight because he is covered in phosphorus, which is known for its chemiluminescent properties but which would probably not glow under the conditions described. It glows only when combusting, which would harm the hound.
Arthur B. Reeve, Craig Kennedy, and Walter Jameson Arthur Benjamin Reeve is one of the most important American pulp fiction writers because he created America’s fi rst scientific detective, Craig Kennedy (Steinbrunner and Penzler 1976; Pitts 1979). In 1951, detective story theorist Howard Haycraft wrote that Reeve’s “Kennedy stories are ‘dated’ and nearly forgotten today, making it easy to underestimate their contemporary importance” (Haycraft 1984). Reeve earned his bachelor’s degree at Princeton in 1903 and a law degree at New York Law School but became a journalist. After writing an article on scientific methods of crime detection, he began writing detective fiction. His detective was inspired by Dr. Otto Schulze, medical adviser to the New York district attorney. Reeve published 12 short stories, beginning with the December 1910 issue of Cosmopolitan, that featured Craig Kennedy, a chemistry professor at the university in the “Heights” area of New York City. In the stories, Kennedy builds electronic devices and performs his own chemical analyses to solve crimes. These collected stories were published as a book titled The Silent Bullet (1912). For instance, in chapter 2, “The Marksman,” Kennedy builds a copy of Bertillon’s dynamometer with Bertillon’s permission. It is a device to measure the amount of energy required to pull open doors, and so forth, that actually was invented by Alphonse Bertillon in the 1890s. In chapter 7, “The Azure Ring,” a variety of poisons are discussed, including potassium cyanide and curare. Kennedy’s roommate and chronicler was Walter Jameson, a reporter for the New York Star. Jameson neglected his professional duties while helping Kennedy solve crimes but wrote them for publication afterward. Even though Craig Kennedy is no longer part of our popular culture, his spirit remains in the characters of Q and Bruce Wayne/Batman. Q for Quartermaster was created by Ian Fleming and is the fictional head of Britain’s MI6 secret service research and development division. Bruce Wayne/Batman was created by artist Bob Kane and writer Bill Finger. Wayne is a wealthy and clever but otherwise ordinary man who fights crime while wearing a Batman disguise. He is the only superhero whose power is scientific and technological prowess, and who has created a gadget for every purpose. All of these characters foreshadowed our current reliance on expensive instrumentation, because chemical analysis has meant instrumental analysis only since 1950, when the fi rst commercial machines became available. A major shift took place between 1920 and
Hard Science = Hard Evidence
195
WWII in which physical methods of mixture separation and compound analysis were developed that were exploited only after WWII. In 1912, Craig Kennedy was way ahead of the curve. “As recently noted, the [instrument] revolution in analytical chemistry is important because with it we have widespread recognition that building a new instrument can teach us about the world just as devising a new theory can” (Baird 1993). In 1914, the Pathé Film Company hired Reeve to write a serial based on his scientific detective Craig Kennedy. Their requirement was that a prominent role be created for actress Pearl White because they wanted to build on her recent success as the charming heroine of the actionpacked The Perils of Pauline (1913). They reasoned he had published in Cosmopolitan and would be known to their mostly female audience. The 14-episode serial was called The Exploits of Elaine and centered on Elaine Dodge, who “was both the ingénue and the athlete—the thoroughly modern type of girl—equally at home with tennis and tango, table talk and tea.” Every episode ended with a cliffhanger, often with Dodge in peril. The serial appeared in theaters across the nation, and every week’s 20-minute episode was serialized in subscribing newspapers. The printed version gave more details and clarified important plot twists. In an example of great prescience, in chapter 1, titled “The Clutching Hand,” Kennedy says, “Every criminal leaves a trace. If it hasn’t been found, then it must be because no one has ever looked for it in the right way.” Perhaps Edmond Locard read this story or watched this serial when it appeared in France. When the movie serial was complete, the written version was published in novel form as The Exploits of Elaine (1915). The villainous Clutching Hand used thermite to burn holes in metal, poison gases, hypnotic drugs such as scopolamine, curare poison darts, and chemical time bombs. Only a chemical detective would know how to identify and defuse such nefarious chemical weapons, especially since he had encountered most of them in his earlier exploits. Before The Exploits of Elaine had fi nished its run, Pathé contracted with Reeve to write scenarios for 10 more episodes. When the box office receipts didn’t slow, they asked for a fi nal 12 episodes. These sequels appeared on the screen and in newsprint in 1915 as The New Exploits of Elaine and The Romance of Elaine. These two serials were published in the form of a novel as The Romance of Elaine (1916). Since both movie serials are presumed to be lost, this book provides an important source for reconstructing their narratives.
THE LIMITS OF DETECTION U.S. Secretary of Defense Donald Rumsfeld characterized risk as “unknown unknowns” (Rumsfeld 2002) but it is possible to defi ne that and his other terms more scientifically. “Known knowns” are facts about physical objects and situations. “Known unknowns” are things we know
196
ReAction! Chemistry in the Movies
we don’t know and encompass statistical probabilities, random deviations from the norm, and risks. Something will happen—we just don’t know its specific time, place, or quantity. “Unknown knowns” are things we don’t know now but could if we began searching for them—either we haven’t wanted to know yet or haven’t had the tools to fi nd out. When these become known, they are called discoveries. “Unknown unknowns” are things we don’t know exist and may not exist. They may be beyond the limits of detection, or they may occur with such low frequency it is not possible to predict a specific outcome. These are uncertainties, imperfections, catastrophes, and mutations. Fictional detectives process new known knowns faster than anyone else and have an uncanny ability to decide which unknown known to search for next. Real detectives have to link two known knowns together to prove contact between them beyond a reasonable doubt. They lower the risk of misdetection due to statistical uncertainties associated with examining small sample sizes by recognizing their method’s limits of detection. The story of arsenic detection provides an illustrative example (Webster 1947). Forensic toxicology was established in 1813 when Dr. Matthieu Orfi la published his fi rst book Traité des Poisons tires Regnes Mineral, Vegetal, et Animal; ou, Toxicologie Generale [Treatise on the Poisons Drawn from the Mineral, Vegetable, and Animal Kingdoms; or, General Toxicology]. It was the commanding text in the field for half a century. It reflects both his medical training, in its description of poisoning symptoms and its microscopic examination of poisoned tissues, and his analytical chemistry skills, in its collection of protocols for poison detection, a number of which he improved. For arsenic traces in human organs (actually arsenic trioxide, As2O3), he used the test developed by German chemistry teacher Valentine Rose in 1806. Rose treated the stomach contents of a poison suspect with potassium carbonate to hydrolyze the organic material, and then calcium oxide to precipitate that material; he then treated the resulting solution to Metzger’s test. In 1787, Johann Metzger discovered that heating arsenic trioxide with charcoal (a source of carbon, which acts as a reductant) caused a shiny black material (elemental arsenic) to deposit onto cool surfaces (2 As2O3 + 3 C o 3 CO2 + 4 As). The next advance in arsenic detection was the more sensitive Marsh test, published in 1836 by English chemist James Marsh (Marsh 1836; Webster 1947). Marsh added zinc metal and sulfuric acid to the test sample to create arsine gas (AsH3) as a product (As2O3 + 6 Zn + 6 H2SO4 o 2 AsH3 + 6 ZnSO4 + 3 H2O). The gas was ignited to combust the hydrogen and release elemental arsenic, which formed Metzger’s shiny black deposit on a cool surface. The assay could be used quantitatively by determining the mass of the cool object before and after the reaction. Orfi la used this assay in many cases, including one that received a great deal of press in 1840. He also discovered that everyone has arsenic in their tissues because traces of arsenic trioxide are present in the soil and our
Hard Science = Hard Evidence
197
food. Its structure and properties resemble phosphate, so the two polyanions are found in the same tissues and bone. Because the Marsh test is so sensitive (it can detect amounts as small as 20 µg of arsenic trioxide, a small mass but actually containing 61 x 1015 arsenic atoms), it could no longer be used qualitatively (is arsenic present or not?) but had to be used quantitatively (how much arsenic is present?) in comparison with control samples (is it present in higher amounts than normal?). After all, the poison is in the dose. The issues of toxic doses (ingestion of 100 mg of arsenic trioxide is usually lethal), chronic versus acute toxicity, and individual metabolic responses are briefly addressed in chapters 4 and 5. Legal facts are much more inclusive than scientific facts. In legal terms, statements are either facts or interpretations. In this sense, almost any information can be a fact. Scientifically, facts are reproducible observations. Anyone can watch the movies described in this book, see the same images, hear the same sounds, but interpret their meanings differently. Scientists avoid the possibility of alternate interpretations by naming and describing their observations in the most detailed and least ambiguous way. If you can’t name it or quantify it, it doesn’t exist. In chemistry, the descriptions often concern physical properties such as color, mass, volume, temperature, and pressure. Each of these can be described subjectively and ambiguously, with terms such as heavy, small, and room temperature. Or, they can be described objectively in a way that anyone else could reproduce, such as 552.2 nm emission, 50.0 mg, and 101.3 kPa. In the latter examples, each measurement consists of a number and a unit. Each unit is defi ned according to international agreement as the Systéme Internationale, or SI for short. The most precisely defi ned unit is the “second,” the basic measure of time (Coel 1988). In 1967, the second became the “duration of 9,192,631,770 periods of the radiation corresponding to the transition between the two hyperfi ne levels of the ground state of the cesium-133 atom.” This effect doesn’t involve nuclear decay but is the result of shining radio waves on cesium atoms to cause the outermost electron’s spin to fl ip. The number of cycles per second was determined during a 2.7-year-long study in the late 1950s that reported a measurement error of only ±20 periods. Mathematically, that means the first seven figures in the number 9,192,631,770 are “significant” but that there is uncertainty about the last “7” in the number. The cesium second replaced the astronomical second, which had been the standard since 1820. Specifically, it had been defi ned as the 86,400th part of the mean solar day, which equaled 365.25 parts of a mean solar year between two vernal equinoxes. The astronomical second varied from year to year because Earth’s rotation around the sun speeds and slows in response to the tides, because Earth wobbles on its axis in response to its sloshing molten core, and because Earth’s speed is progressively slowing. The problem had become so acute by 1972 that the fi rst two leap seconds were added to the calendar to prevent the two types of seconds from getting out of synchrony by more than 0.7 seconds. Since
198
ReAction! Chemistry in the Movies
the second is now so precisely defi ned, phenomena can be measured in the subnanosecond time scale with very high precision. The kilogram (about 2.2 pounds) is the only scientific unit still based on an actual physical object (Girard 1994). In 1889, the kilogram was defi ned as the mass of the International Prototype Kilogram (IPK), kept by the Bureau International des Poids et Mesures in a Paris suburb. There is no error associated with its measurement, but there are a number of identical copies kept around the world that have masses that can differ. The masses of these “prototypes” are determined every 40–50 years relative to the IPK. The IPK and three identical 90% platinum/10% iridium alloy cylinders were constructed in 1879, and 40 more replica “prototypes” were constructed in 1884 (two of these were given to the United States, where they are now kept by the National Institute of Standards and Technology). The iridium was added to increase the object’s hardness and resistance to oxidation as well as to lower its magnetic susceptibility. The fi rst measurement in 1889 found them to have identical masses, but the 1946, 1989, and 2007 reports found the prototypes had diverged from the IPK by an average of about 50 µg. The variance creates problems when measuring other things relative to the prototypes when the prototype masses keep changing. To put the mass differences in perspective, the IPK mass is 1 kg exactly, whereas the average prototype mass is 1.00000005 kg. That is, the average prototype differs from the IPK by 50 parts per billion (ppb). The reasons for the divergence in mass of the IPK from the prototypes, or vice versa, is unclear, but the ability to measure such a small differences with such high precision is a testament to the storage, cleaning, and measuring methods that are employed. The drive to detect lower and lower amounts of chemicals has followed on the heels of the drive to defi ne the units of measurement ever more precisely. A case in point is the detection of dioxin. The toxic effects of dioxin were fi rst noted in the 1950s and 1960s after accidents involving workers at various chemical factories in the United States, United Kingdom, and Germany who were producing the herbicide 2,4,5-T (later a component of Agent Orange; see chapter 3) (Crummett 2002). The toxic effect was not death but a reversible form of acne, called chloracne, induced by many chlorinated organic compounds. The dioxin story is complicated because it induces chloracne and general malaise in humans, which can’t be studied except in cases of accidental exposure. Nevertheless, dioxin is useful in the present context because it was the most public example of increasingly lower limits of detection. One of the largest 2,4,5-T manufacturers in the United States was Dow Chemical. Analytical chemist Warren Crummett had a career at Dow, where he became a senior manager (Crummett 2002) and was responsible for developing dioxin detection methods during the 1960s, before the public became interested in the topic. He was the person who met with federal officials to communicate that dioxin was present at 27 ± 8 ppm in the 2,4,5-T of one company’s samples. The events and regulations that transpired quickly took on a life of their own, and Dow was increasingly
Hard Science = Hard Evidence
199
called on to explain how much dioxin was present in their products. One issue that Crummett and others soon encountered was the demand for “zero tolerance” limits. Technically, it doesn’t make sense to call for zero amounts of something unless you defi ne the method of detection, since detectable amounts are always limited by the instrumentation and method. It is not possible, or economically feasible, to detect all things at the level of single molecules, ions, or atoms. To help him communicate the difficulty of fi nding trace concentrations and the continually improving limits of detection, Crummett developed a chart of trace equivalents in 1976, a small portion of which is reproduced in table 7.2. Risk is a topic intimately related to limits of detection. To help others manage risk (Crouch and Wilson 1982; Wilson 1984), Richard Wilson from Harvard University had a different take on small quantities and small probabilities. He created a risk index that stated the likelihood of increasing the death risk by 1 ppm (the same as reducing your life expectancy by 8 minutes). His 1 ppm risk index included smoking 1.4 cigarettes, having one X-ray taken at a hospital, eating 100 charbroiled steaks, and living 150 years within 30 miles of a nuclear power plant. The public is aware of and willing to accept certain voluntary risks, such as automobile deaths and lung cancer deaths due to cigarette smoking, but not others, such as living near a nuclear power plant. On longer reflection, these comparisons cause the reader to realize that cigarettes are worse than nuclear power plants. On the other hand, the fear of things with low probability is higher today with the advent of terrorism campaigns even as the actual death rates keep dropping. Even though Americans are far more likely to die from lung cancer or heart disease than from acts of terrorism, people are aware of more risks (known unknowns) than ever before. If they are convinced it can happen to them, it doesn’t matter that the statistical probability is low. Every measurement is associated with an error for that measurement. When the mass of an object is determined using an instrument close to its limit of detection, you don’t always get the same mass twice. Instead, you get a distribution of masses that are represented by an average mass and the standard deviation, for instance, 105 ± 5 mg. The
Table 7.2. Visualizing and communicating trace concentration units Unit
1 ppm
1 ppb
Text Scientific Weight Quality Time Money
1 part per million 1 microgram/gram 1 ounce salt /31 tons potato chips 1 bad apple/2,000 barrels 1 minute/2 years 1¢/$10 thousand
1 part per billion 1 nanogram/gram 1 pinch salt/10 tons potato chips 1 bad apple/2 million barrels 1 second/32 years 1¢/$10 million
The units of weight, quality, time, and money were collected from W. B. Crummett, Decades of Dioxin: Limelight on a Molecule. Philadelphia: Xlibris, 2002.
200
ReAction! Chemistry in the Movies
standard deviation describes the certainty with which the average value was determined. By statistical defi nition, the standard deviation (from 95 to 110 mg in the example) actually indicates that 64% of the measurements (six in the example since there were 10 measurements) fall within that deviation. The remaining 36% of the measurements (four in this example) fall outside that range. (Readers wishing to brush up on probability distributions and estimations of statistical parameters may wish to consult a basic statistics text or the reasonably intelligible entries on wikipedia.org.) The measurement deviations are caused by errors from the user, instrument, and sample. User error can be minimized by training and familiarity. Instrument error can be minimized through user training but is ultimately dependent on instrument design. Sampling error plays a role when the sample does not have a uniform composition, such as in the case of many forensic samples, but can be reduced by user training. For comparison’s sake, undergraduates in a chemistry lab are pleased when their samples show differences of only 5%, or 5 parts per 100. Dr. Gunter Zweig defi ned the ability to measure ever smaller amounts as the “vanishing zero” effect (Zweig 1970, 1978). When an airplane disappears from view in the sky, you can use binoculars and then a telescope to watch it disappear again and again. Each observation method carries with it the ability to estimate the size of the plane but with less certainty the farther it is from view. The “limit of determination” (or “limit of quantitation”) is the point at which the plane’s size cannot be reliably be measured, and the “limit of detection” is the point past which the plane can be observed. Zweig tabulated the lowest reported detection limits from various analytical procedures (Zweig 1978). Crummett added the dioxin detection limits (Crummett 2002). At a concentration of 0.01 ppt, dioxin would be present at about 10 million molecules in a pool of 1 gram of other material (table 7.3).
Table 7.3. Lowest limits of detection in ppt (or picograms/gram) Year
All Methods
1958 1960 1965 1970 1975 1980 1983
1,000,000 500,000 1,000 100 1 — —
Dioxin Detection — — 1,000,000 50,000 10 0.2 0.01
Data collected from W. B. Crummett, Decades of Dioxin: Limelight on a Molecule. (Philadelphia: Xlibris, 2002), which includes Crummett’s (2002) data for dioxin detection and Zweig’s (1970) for the lowest detection limit of all analytical methods.
Hard Science = Hard Evidence
201
With the above thoughts in mind, let us conclude with a mathematical consideration of the limits of detection and determination. In 1968, in the midst of numerous studies claiming to have achieved the lowest possible limits of detection, Lloyd Currie from what is now the U.S. National Institute of Standards and Technology defi ned those terms precisely (Currie 1968). Terms such as “limits of detection,” “detection sensitivity,” and “minimum detectable activity” were being used indiscriminately and could mean either the “minimum amount detected” or the “minimum amount that could be determined relative to some standard deviation.” The essence of Currie’s argument is that the ability to detect (qualitative analysis) or determine (quantitative analysis) depends entirely on the chosen limits of confidence in the measurements, which in turn depends on the standard deviation of the measurements. In his examples, he chose a high confidence limit of 95% (meaning there are 19 chances in 20 to get the correct answer and only a 1 in 20 chance of “false positives” and “false negatives”) to arrive at the reasonable proposal that something can be reliably detected when it is present at 3.29 times the standard deviation (SD) of the measurement. That is, the limit of detection (LOD) is 3.29 x SD, a useful relationship that can be remembered as the “rule of three times SD”). He further proposed that an amount can be reliably quantitated when the error of its measurement is less than 10% of the sample’s concentration, or 10 times greater than the standard deviation. Using Currie’s defi nitions and the example of 27 ± 8 ppm dioxin in 2,4,5-T, we can be confident that dioxin has been detected if it is present at more than 3.29 times 8 ppm, or 26 ppm. Since the measurement was 27 ppm, it is present at an amount just barely above the detectable limit. It is nowhere near the 80 ppm at which we could reliably determine its concentration. In conclusion, any value that does not provide information about the confidence of its measurement is not scientifically useful, but when provided with the standard deviation of the measurement, it is possible to reconstruct its detectability by using the “rule of three times SD.”
CHECKMATE Investigator and fi refighter Brian McCaffrey (William Baldwin) is questioning imprisoned arsonist Ronald Bartel (Donald Sutherland) for leads to help solve a series of arson/murder crimes in Backdraft (1991). In the process, he fi nds out just how closely his own irrational fascination with fi re reflects the criminal’s pyromania. In a three-minute scene starting at 1:40:30, Bartel says, “Did it look at you? Did the fire look at you?” McCaffrey’s eyes betray his controlled demeanor. Delighted, Bartel says, “So, our worlds aren’t that far apart after all.” The criminal/detective duality has elements of playful sparring, but catching a clever criminal is no trivial pursuit. Let’s call on our crack
202
ReAction! Chemistry in the Movies
companion, the mirror, to shine some reflective light on the nature of the game these opponents are playing. Detective fiction is the literary source for many of the movies in this chapter. In this section we consider the representation of analytical detective aspects of the genre, or the use of logic to reach conclusions that ultimately solve crimes. Drawing on John T. Irwin’s prize-winning study of the topic in The Mystery to a Solution: Poe, Borges, and the Analytic Detective Story we examine how the criminal/detective duality fi nds its mental expression in the geometrical pattern and numerical configuration of mirror imagery (Irwin 1994). We also consider how the arrangement of material in a book about chemistry in the movies has an affi nity to Irwin’s conception of the “specular double.” The analytical detective story has a mirror structure at its core that is revealed in the “battle of wits” taking place between detective and criminal. Edgar Allan Poe inaugurates this specialized plot architecture, now the genre’s “mainstay,” in the Dupin stories, described above. Dupin, in effect, solves the crime by identifying with the criminal. His gaming strategy is to double the criminal by putting himself into the criminal’s mindset. Once there, he looks with intimacy at the criminal’s motivations and methods in order to anticipate the next move and assume the advantageous position, or get “one jump ahead,” as Irwin puts it. Dupin’s own mind thus becomes as “other” to himself; it splits in two, taking the form of a mirror image. Poe uses this device to bring readers with him into what Irwin describes as the “dizzying, selfdissolving effect of thought about thought.” The twist is that the subject is a reflecting self facing an object that is the self under observation (Pyrhoenen 1999). In The Seven-Per-Cent Solution (1976), Sherlock Holmes, under the influence of cocaine-induced paranoia, startles Dr. Watson with an outburst about his archenemy Professor Moriarty. At a small table holding a chess set, Holmes grabs a fi stful of pieces and pounds the board with such feverish intensity that other pieces totter and fall. The chess set, which might seem an incidental prop on a movie set for a nineteenthcentury gentleman’s study, is a potent surrogate for the sophisticated game Holmes and Moriarty are playing in the Conan Doyle stories. The symbolism of chess is also present in subtle and not-so-subtle ways in The Bone Collector (1999). We see its suggestion in the “grid” that crime scene investigator Amelia Donaghy carefully walks to uncover evidence. It is transparent in several scenes: Detective Lincoln Rhyme is repeatedly shown absorbed in a chess game with a computerized opponent, and also when the criminal tells Rhyme that this is a “chess game” and he has won. All of this imagery comes before the climactic ending that pits the criminal, a white man (read as white king), against the detective, a black man (read as black king), in a harrowing checkmate. In this scene at 1:50:00, both men are physically compromised in overt ways; they are opponents symbolically locked in a mental struggle.
Hard Science = Hard Evidence
203
As Irwin argues, the game of chess is “one of the most common tropes” for the criminal/detective mental duel in the analytical detective tradition. Its physical structure comprises horizontal, vertical, and diagonal axial oppositions within a same-sided square. Self-reflective analytical thought can be symbolized by taking the whole game into consideration, a game composed of interwoven oppositions of sameness and difference. For example, an even and odd trajectory begins with the alternation of moves that starts the play sequence. Beginning with an odd number, 1, it then proceeds to an even number with the second turn, and so forth. In this way, the players always fi nd themselves one step away from coincidence. Irwin also notes that when two chess players face one another across a chessboard, the players’ hands directly opposed from one another have different names (left and right), while the players’ hands that are diagonally opposed have the same names (right and right, or left and left). The initial arrangement of chess pieces on the board, however, illustrates a reversal with respect to the two players. Since the king and queen face one another directly, one player has the king side on the right and the queen side on the left, and vice versa. This arrangement leads to the pieces directly opposed having the same names (king and king, or queen and queen), while the pieces diagonally opposed have different names (king and queen). Thus, as Irwin points out, this “internal asymmetry” of the game gives it an added dimension. Chess isn’t just a competition between two players; its structure also symbolizes the competition of one player with his or her own mirror image. These two competitive modes, present simultaneously in the chess game, induce the idea of the “specular double,” which Irwin defi nes as “mentally self-reflective.” The mirror image looks back at the subject, making it into its own mirror image. Thus, the subject reflects on its own act of cognition. The notion of an internal split that cannot be reconciled is the result. Analyzing the act of analysis leads to the subject’s oneness with itself being interrupted by a persistent sense of otherness to itself. How does all this fit with the scheme of a book about chemistry in the movies? Not to be heavy-handed, but it may be a happy coincidence that the archetype movie of this chapter is Kid Glove Killer (1942), since this brings us back to the idea of a glove and thus to the underlying cipher of the book: chirality. Like a chiral molecule, a single glove manifests its own oppositional character when it is turned inside out in three-dimensional space. After performing this maneuver, a glove made for the right hand, for instance, would now fit the left hand. As discussed in the fi rst section of chapter 1, such a form is geometrically equal but not congruent with its partner; three-dimensional pairs of hands and chiral molecule pairs can achieve congruence only through the inversion of either one. These left and right mirror-image pairings are always one step away from being exactly the same or identical to one another. And so, we have paired the dark and bright sides of this book to resonate with this same/different paradox of chirality. Like Irwin’s specular double, each side knowingly
204
ReAction! Chemistry in the Movies
reflects on the other in the layered way that the themes are organized, by opposing one another directly and diagonally in sameness and difference. An example of a theme in direct opposition is chapter 1’s destructive discovery versus chapter 6’s creative discovery, and a diagonally similar theme is the self-experimentation that links chapter 1 to chapter 10. We now know that the chess game clue leads us to chirality. But the solution to ReAction! Chemistry in the Movies may, in one sense, always be out of reach. Perhaps we can grasp it better in the interplay of movie themes that splits and doubles its way self-consciously toward an analysis of the subject of chemistry itself.
THE ARCHETYPE MOVIE: KID GLOVE KILLER (1942) Distribution company: Metro-Goldwyn-Mayer Director: Fred Zinnemann Screenwriter: Allen Rivkin, from John C. Higgins’s story and screenplay for the 1938 short titled They’re Always Caught Short summary: Police chemists Jane Mitchell and Gordon MaKay use a spectrograph to discover traces of vanadium in a bomb’s gunpowder residue Plot description: Jane “Mitchell” Mitchell (Marsha Hunt) has a master’s degree in chemistry from Montana and has just joined the Chatsberg [alias Chicago] police forensic lab. Gordon McKay (Van Hefl in) is her supervisor and only coworker. In the lab, they wisecrack in a battle of the sexes and use Bunsen burners to light the cigarettes they chain smoke. The secretly corrupt special prosecutor Gerry Ladimer (Lee Bowman) plants a bomb under the reform mayor’s car. After the mayor is killed, McKay and Mitchell collect clues such as the metal shards of the makeshift pipe bomb and bring them back to the lab for analysis. In the 2.5-minute scene at 49:30, the police captain joins them in the instrumentation room. Mitchell places some of the bomb’s explosive into the spectrometer’s chamber and lights it while McKay explains to the captain how the instrument works. As the material burns, it emits a spectrum that is captured on a photographic plate. The resulting line spectrum is then compared to known elemental spectra. This sample has the “particularly useful” vanadium triplet. McKay pulls out the “gunpowder bible” and fi nds that the powder came from 38-mm magnum bullets. The police round up suspects so McKay can check the dirt under their fi ngernails for vanadium-containing gunpowder traces. Commentary: In this fi lm, a spectrograph is used to identify the elemental components of gunpowder scraped from a pipe bomb. Vanadium is present not for its explosivity but because it was deliberately added as a tracer. Today, many controlled products contain unique sets of tracers. Every student of general chemistry learns that each element emits
Hard Science = Hard Evidence
205
Figure 7.1. This “spectrum analysis” glass slide was created by the Briggs Company, probably in the 1920s, so that students and researchers could see the line spectra for potassium, rubidium, sodium, and lithium. Image courtesy of the George Eastman House.
a unique line spectrum (figure 7.1). Students also learn that the wavelengths of each line are related to the energy of the photon that is emitted as excited electrons fall from particular higher energy states to unfi lled lower energy states. Since every element has a unique line spectrum, a spectrophotometric analysis can be used to reveal the elemental composition of any substance. This feature fi lm is based on a short fi lm about forensic police work titled They’re Always Caught, directed by Harold S. Bucquet. It was #18 in MGM’s Crime Doesn’t Pay series and was nominated for an Academy Award. Given the public’s current interest in CSI television shows and forensic chemistry, it is a curious fact that these films are not commercially available on video.
CHEMICAL DETECTION MOVIES The League of Extraordinary Gentlemen (2003) Distribution company: Twentieth Century Fox Director: Stephen Norrington Screenwriter: James Robinson, based on the same-titled 1999 graphic novel by Alan Moore and Kevin O’Neill
206
ReAction! Chemistry in the Movies
Short summary: Characters from late Victorian fiction unite to defeat the evil Fantom in 1899; the former nurse Mrs. Harker uses inorganic qualitative analysis to identify the traitor MPAA rating: PG-13 Plot description: In 1899, the Fantom’s men steal Leonardo da Vinci’s map of Venice from the Bank of England while dressed as Germans. Next, his men are dressed as British soldiers while they invade a Berlin dirigible factory, where the Fantom shoots at one of the hydrogen-fi lled dirigibles. As the zeppelins burn, newspapers report and voices speak of British and Germans blaming each other for inciting war. Allan Quatermain (Sean Connery) agrees to lead a team to discover who is behind these crimes because he worries a European War will spread to Africa where he lives. In a room deep beneath London, Quatermain meets M (Richard Roxburgh), who tells him about previous Leagues of Extraordinary Men who have solved other crises. Quatermain reads the resumes of the other team members and learns that Mrs. Mina Harker (Peta Wilson) is a chemist. This prompts the Invisible Man (Tony Curran) to announce his presence. He is Rodney Skinner, gentleman thief, who stole the invisibility formula from a chemist. He has agreed to help so that the government will search for the antidote. Captain Nemo (Naseeruddin Shah) is on the team to transport everyone on his Nautilus submarine. When Harker enters, Quatermain cannot believe they are going to include a woman on the team. M says Quatermain doesn’t know about her powers. The group travels to an East London Dock to fi nd Mr. Dorian Gray (Stuart Townsend). Once there, the Fantom and his men barrage them with machine gunfi re. The league is able to repel them without suffering any losses in part because Tom Sawyer (Shane West) infi ltrated their ranks. When one of the vanquished foes wakes and grabs Harker, she kills him as only a vampire could. She and her husband had been Dr. Van Helsing’s assistants. The group heads to Paris to gather the fi nal member of their league—a monkey man who has terrorized the Rue Morgue in Paris for months. Quatermain and Sawyer hunt down Mr. Hyde (Jason Flemyng) and then restrain him inside the Nautilus. Quatermain offers Hyde amnesty to return to England if he helps. He longs to do so and reverts back to Jekyll when the formula wears off. The Nautilus travels from London to Venice in only 3 days. Nemo discovers some powder in the steering room and gives it to Harker for analysis. In the 3.5-minute scene at 48:00, Harker works amidst her chemical apparatus. Just as her test tube reveals the identity of the powder, Gray emerges from the shadows. She says it is phosphorus, perhaps to create a photographer’s flash. Another character later confi rms that it was used for that purpose and that it created magnesium phosphate as a product. Gray and Harker get intimate while Jekyll watches around the corner.
Hard Science = Hard Evidence
207
Jekyll lusts for Harker as he travels back to his room, where he fi nds that one of his vials is missing. Jekyll tells the others that Skinner must have stolen it. Commentary: This is an enjoyable fi lm because of the many literary and movie references to characters mentioned elsewhere in this book. In addition, Mrs. Harker’s clinical chemistry skills include inorganic qualitative analysis, which she uses to unravel an important clue that reveals there is a traitor in their midst. The scriptwriter greatly expanded Mrs. Harker’s role from the graphic novel and invented the forensic episode. Prior to the use of flash bulbs, photographer’s flash powder consisted of magnesium and phosphorus powders. They combust rapidly and brilliantly when sparked: 6 Mg + P4 + 8 O2 o 2 Mg3(PO4)2 + bright white light. Merci pour le Chocolat (2000) Distribution company: First Run Features, USA Director: Claude Chabrol Screenwriter: Claude Chabrol, based on the 1948 Charlotte Armstrong novel The Chocolate Cobweb Short summary: Jeanne Pollet’s quest for her parentage uncovers murder and the date rape drug Rohypnol Plot description: Mika Muller (Isabelle Huppert) is president of Muller Chocolate in French Switzerland. She is also the second wife of concert pianist André Polonski (Jacques Dutronc) and stepmother to Guillame (Rodolphe Pauly), for whom she prepares a carafe of hot chocolate every night. When Polonski’s fi rst wife gave birth to Guillame about 20 years before, Polonski was shown a newborn girl at the hospital and was mistakenly told it was his. That baby girl is now a young woman named Jeanne Pollet (Anna Mouglais), and she is training to become a concert pianist. While growing up, she heard the story of the hospital mix-up many times and often wondered whether she is indeed Polonski’s daughter. Jeanne’s mother (Brigitte Catillon) is the head of the forensic lab, and Jeanne’s boyfriend Axel (Mathieu Simonet) works for her mother. Without telling her mother, Jeanne decides to visit Polonski’s home and recount the hospital story. When he learns that she wants to enter a prestigious competition, he decides to train her. One night, Jeanne accidently sees Mika spill some of the hot chocolate in Guillame’s room while she looks at a photo of Guillame and his fi rst mother. Jeanne helps her wipe it up, but then Mika insists on washing the handkerchief. Some of it is on Jeanne’s sleeve, so she asks her boyfriend Axel to analyze it. In the 1.75-min scene from 29:15, Axel calls Jeanne to tell her that it contained benzodiazepine, which helps you sleep. He can’t believe it would be fatal, but he hasn’t checked the dose. Then, he asks Jeanne whether she knows what Rohypnol is and says it is the date
208
ReAction! Chemistry in the Movies
rape drug. When he hangs up, she opens one of her mother’s pharmacology books. When Jeanne tells Guillame the story, he fi rst resents her attempts to become part of his household but then tells her the story of his mother’s tragic death when he was 10 years old. Commentary: Flunitrazepam is one of the strongest benzodiazepines (figure 7.2). For years it was marketed in many countries by Hoffman-LaRoche as Rohypnol, a treatment for insomnia and a preanesthetic (Labianca 1998). Benzodiazepines bind to the GABA receptors to induce a hypnotic sedative effect, which is useful for treating insomnia. One of Rohypnol’s side effects is that it causes short-term memory loss. It has never been approved for medical use in the United States, where it is considered to be an illegal drug. The Bone Collector (1999) Distribution company: Columbia Pictures Director: Phillip Noyce Screenwriter: Jeremy Iacone, based on Jeffery Deaver’s same-titled 1997 novel Short summary: Quadriplegic New York City police detective and true crime book author Lincoln Rhyme leads the investigation against a serial killer who leaves clues about the next murder site MPAA rating: R Plot description: While investigating a dead body in a crawl space, an accident left New York City Police Detective Lincoln “Link” Rhyme (Denzel Washington) with the use of only his head, shoulders, and a fi nger. He is a legendary detective on the force and has an encyclopedic knowledge of New York City from the dozen books he has written about true crime stories. In his vast loft apartment, he is confi ned to his bed, where he
H3C
O
N
O2N
N F
Flunitrazepam (Rohypnol)
Figure 7.2. The benzodiezapine flunitrazepam is marketed under the trade name Rohypnol.
Hard Science = Hard Evidence
209
plays computer chess and is cared for by nurse Thelma (Queen Latifah). Rhyme has occasional seizures and worries the next one will turn him into a vegetable. He has arranged for his doctor friend to help him commit suicide on Sunday. Amelia Donaghy (Angelina Jolie) is a straight-thinking street cop who responds to a call about a crime, only to discover a murder scene. She stops a train before it disperses some clues and then photographs the victim and the clues before police detective Captain Cheney shows up. He is angry she stopped the train and threatens to demote her. Interspersed with the above, the audience sees the abduction of a couple returning to New York by plane. Their hired car is not there to meet them so they take a taxi with the number 5Z66. After killing the husband and staging his crime scene off screen, there are gratuitously violent scenes against the wife. Police investigators arrive at Rhyme’s apartment with the information that police chief Murphy wants him to take over the investigation of the couple missing from the airport. From the clues, Rhyme quickly surmises the next killing will occur at 4 P.M. that same day, November 9. He asks for help from the cop who photographed the crime scene, saying that she has natural ability in forensics. The entire forensics team and their instruments are assembled in Rhyme’s apartment as they examine the clues in greater detail. Rhyme couples this with his extensive knowledge of New York to pinpoint the location of the next crime. Unfortunately, the police arrive too late. Donaghy “walks the grid” as she reluctantly collects the deliberately placed clues. In the 4-minute scene starting at 1:04:45, forensic scientist Eddie Ortiz (Luis Guzman) shakes a test tube of white solution in front of the sleeping Rhyme’s nose to wake him up. Ortiz says the bone was a veal shank. It is a cow bone. The hair was probably shaved from a rat. They speculate the perpetrator could be on the police force and certainly understands forensics. Ortiz reports that the dirt on the bone is nitrogen-rich, like manure, but has aged and oxidized to nitrate. These clues lead to them to the next site, but once again, they arrive too late. Commentary: Besides containing some of the most grisly scenes of any movie in this book, it sports a detective who is most like the classical analytical detectives Dupin and Holmes. As a quadriplegic, Lincoln Rhyme represents total rationality. He is able to unravel clues by rapidly associating them with his vast knowledge of past crimes and New York City history. Amelia Donaghy, with whom he communicates electronically, is his remote physicality. Given that a detective is only as great as her or his nemesis, this one must be truly great given the depth and intensity of his foe’s insider knowledge and desire for revenge on both personal and professional levels. Backdraft (1991) Distribution company: Universal Pictures
210
ReAction! Chemistry in the Movies
Director: Ron Howard Screenwriter: Gregory Widen, a former Chicago fi reman Short summary: Arson detectives Donald “Shadow” Rimgale and Brian McCaffrey investigate a series of arson/murders involving magnesium and a liquid called trychtichlorate MPAA rating: R Plot description: Chicago fi refighter Dennis McCaffrey (Kurt Russell in a flashback) died in 1971 while fighting a fi re. His youngest son Brian was with him that day and watched it happen. Twenty years later, his older son Stephen “Bull” (Kurt Russell) is a lieutenant on the force to which he devotes his life. His unsettled son Brian (William Baldwin) has bounced between jobs until fi nally passing fi re training so he can prove himself. On his fi rst day at work, he fights his fi rst real fi re and “saves” a mannequin much to the amusement of his seasoned mates. After the fi re has been quenched, they all smoke cigarettes as they knock down smoldering ceiling tiles. Brian painfully decides to quit fighting fires after he admits he’s afraid. At the urging of a city alderman, he transfers to the arson division to work under Lieutenant Donald “Shadow” Rimgale (Robert DeNiro). Brian is supposed to help speed the investigation of a possible arson/murder. On arriving at Rimgale’s office, Brian learns that Rimgale’s body was burned many years ago while saving the life of an arsonist who set a phosphorus fi re and that it was Brian’s father who saved them both. In rapid succession, there are two more arson/murders, one of which burns a rookie fi reman Brian knew. All the cases involved a backdraft fi re, in which the victim opened the door only to be burned and thrown away from the door as the smoldering flame reignites from the entering oxygen. During the investigation of each fi re, Rimgale smokes a cigarette as he sniffs around. Rimgale shows Brian how to fi nd the source of a fire. He says you have to love fire a little to understand it: “Fire is alive. It breathes. It eats. It waits.” In the 2.0minutes scene from 1:07:15, Rimgale and Brian are in the morgue examining a burnt corpse. The doctor says one common element to the fires has been putty on the doors and the presence of trychtichlorate, which needs magnesium to start. In the 1.5-minute scene from 1:10:30, Rimgale has Brian open a trashcan to teach him how a backdraft works. The final firefighting scene takes place in a chemical factory. Commentary: Clues in the movie lead to a can of solvents with the following ingredients: methylethylketone, dioxoethylene, trychtichlorate, methylene chloride. Of these, trychtichlorate is named as the important compound for setting the arson fi re. It is also a fictional compound, something most viewers wouldn’t know. Its name and properties are reminiscent of trichloroethanol, a paint stripper and industrial solvent. This movie doesn’t actually reveal any real chemistry even though it shows arson realistically. This was probably done so that no viewers could learn how to become an arsonist by watching the movie.
Hard Science = Hard Evidence
211
The special effects (F/X) crew did an excellent job on this movie. The fi re moves as though it is alive and interacting with the actors. The water and explosions hit their marks. F/X is a highly specialized field requiring knowledge of making rain, snow, fog, explosions, fi res, and anything else chemical. F/X workers are also extremely concerned about safety of the crew and the actors when they design their effects. They have a safety adage, “There are old F/X workers, there are bold F/X workers, but there are no old, bold F/X workers” (McCarthy 1992). Batman (1989) Distribution company: Warner Brothers Director: Tim Burton Screenwriters: Sam Hamm and Warren Skaaren, using characters created by artist Bob Kane and writer Bill Finger Short summary: Thug Jack Napier is disfigured in a vat of chemicals, so he uses cosmetics to cover his face; he creates a poison named Smile that forms when certain personal care products are mixed MPAA rating: PG-13 Plot description: Boss (Jack Palance) sees an article in the newspaper about the new district attorney (Billy Dee Williams) cracking down on crime. Boss tells Jack Napier (Jack Nicholson) to steal the records at Axis Chemicals so it will be impossible to prove a connection. Napier is reluctant to do it because of “the fumes in that place.” It turns out that the Boss set Napier up to be killed during the heist because Jack stole his woman. The 4.5-minute scene at the factory takes place at 26:45. At one point Napier takes an axe to a vat, causing a liquid to spew out. During the confrontation with Batman (Michael Keaton), Napier falls into a vat of bubbling green liquid but emerges as the Joker. The Joker in white face with a permanent grin kills the Boss and takes over while wearing cream-colored makeup. He uses his chemical skills off screen to create and mass market an entire line of personal care products. In the 2.5-minute scene at 54:30, a male newscaster reports that two models are dead. A photo shows them in white face smiling like the Joker. A woman newscaster laughs when the deaths are described and then dies with a grin on her face. The broadcast is interrupted by Joker’s advertisement to the world. His “New and Improved Joker Products” contain the new secret ingredient “Smile.” “You’ve bought them already.” In the 4.0-minute scene at 1:14:00, Batman and Vicki Vale (Kim Basinger) are working in the Batcave. The police are searching for one tainted product but Batman discovers that the Joker has arranged for an untraceable toxin to form when two products are mixed, such as hairspray with lipstick. The newspaper headline reads: “Batman Breaks Joker’s Poison Code.” Commentary: In many Batman stories, forensic chemistry provides the critical connection. In this version, however, Batman’s chemical abilities
212
ReAction! Chemistry in the Movies
are upstaged by the Joker’s, with echoes of The Incredible Shrinking Woman (1981), described in chapter 4. The Name of the Rose (1986) Distribution company: Twentieth Century Fox Director: Jean-Jacques Annaud Screenwriter: Andrew Birkin, from the 1983 English translation of the 1980 Italian novel Il Nome Della Rose by Umberto Eco Short summary: Brother William of Baskerville investigates mysterious deaths in a Benedictine abbey while the monks are convinced the apocalypse is coming MPAA rating: R Plot description: Brother William of Baskerville (Sean Connery) and his apprentice Adso of Melk (Christian Slater) journey by foot through the snow to a remote abbey to investigate the mysterious death of one of its monks. William soon deduces that the monk committed suicide but is not yet certain of the reason. Next, another monk secretly reads a book in the scriptorium late at night. After he laughs, he hears a noise, gets nervous, sees a mouse, licks his thumb, and continues reading. The next day his dead body is found in a bucket of blood. In the 1.0-minute scene at 26:00, the apothecary Severinus (Elya Baskin) helps William examine the monk’s corpse. As they work, Severinus describes how various herbs are used while an apparatus bubbles in the foreground. William asks about arsenic, and Severinus replies, “It is most effective for nervous orders if taken in small doses.” William responds, “And what if not so small doses?” Severinus replies, “Hmmm, death.” In a short scene at 41:30, William uses a candle’s flame to reveal Greek writing on a piece of paper in the scriptorium. William explains to Adso that it was written with lemon juice. As the corpses keep piling up, William has to deal with an increasingly nervous group of monks, an angry abbot, and a visit from the Inquisition. William uses logic to comprehend the clues and to determine how and why each monk was killed. Commentary: The movie won a French Cesar for Best Foreign Film, and Sean Connery won a British Academy Award. Any fluid, including water, can be used for invisible writing. To write a lemon juice note (McCarthy 1992), apply the juice from a lemon to paper using a brush or pen, allow it to dry, and then develop the message by holding it to a light bulb or use a match flame. The wetted portion of the paper has frayed more than the rest so it burns fastest to leave a brown-colored message. Moonraker (1979) Distribution company: United Artists
Hard Science = Hard Evidence
213
Director: Lewis Gilbert Screenwriter: Christopher Wood, using character names and title from the 1955 Ian Fleming novel Short summary: Industrialist Hugo Drax alters a South American orchid to produce a hypertoxic molecule MPAA rating: PG Plot description: Hugo Drax (Michael Lonsdale) owns Drax Industries and is fi nancing the U.S. space shuttle program. His company builds shuttles called Moonrakers and trains the astronauts to use them. After one of the space shuttles disappears en route to its proper destination, British MI6 agent James Bond (Roger Moore) is called in to investigate. While snooping around Drax Industries headquarters in California, he meets space scientist Dr. Holly Goodhead (Lois Chiles), who is later revealed to be an undercover CIA agent. Bond learns that Drax has secretly built eight shuttles for his own use, prepared a space station, and is having some special vials prepared by a glazier in Venice. When one of Drax’s own space shuttles was accidentally destroyed, he had to replace it quickly with the one he promised to the United States because he’s on a tight schedule. Bond follows the glass clue to Venice. There, he visits a glass museum only to discover it is a high-tech glazier facility preparing unusually shaped vials (figure 7.3). Bond is surprised to fi nd Dr. Goodhead there before him. After narrowly escaping death, he takes one of the sealed vials for analysis. The results provide a clue that sends everyone to Rio de Janeiro. In Brazil, Bond meets with M (Bernard Lee) and Q (Desmond Llewelyn) for a short debriefi ng session that begins at 1:16:00. Q tells Bond that the nerve gas found in the Venetian vial is hypertoxic to humans but not animals. Drax intends to kill all of Earth’s inhabitants while his chosen crew is in the Space Station. When the toxin has dissipated, they will return to populate Earth with their offspring. Q projects the chemical formula onto a screen (figure 7.4), and before the audience sees half the molecule, Bond identifies it as the “chemical formula for Orchidae anenarum.” The orchid was thought to be extinct but was recently discovered in the region of the River Tipperachi in Brazil, his next destination. Commentary: We know that Bond actually meant to say that it was the “chemical formula for the toxin from Orchidae anenarum.” It is later stated that this pollen toxin caused the sterility that killed off the Incan civilization and that Drax has engineered it to kill people outright. Each Venetian vial contains enough nerve agent to kill one million people. The idea of a deadly orchid toxin appears to have been the creation of the screenwriter, since the 1955 novel of the same title deals with industrialist Hugo Drax who threatens London with a nuclear bomb. The toxin may have been inspired by the French spy fi lm Furia à Bahia
Figure 7.3. James Bond (Roger Moore) examines a suspicious vial from a high-tech glassmaking facility in Venice. MOONRAKER © 1979 Danjaq, LLC & United Artists Corporation. All Rights Reserved. Photo courtesy of the Academy of Motion Picture Arts and Sciences.
ELEMENT BREAKDOWN (SERIES 9)
DS
H3C
C
O CH3 CO
H
C
O C
CH2 (CH2)2
CH
CH2
CH2
CH2
CHCO2
S (C2H3O)2
PO
NO2
Figure 7.4. The chemical formula for the toxin from the Orchidae anenarum. The three cyclohexanes resemble the shape of its glass container. It would be a solid at room temperature given its heavy molecular mass.
Hard Science = Hard Evidence
215
pour OSS 117 (aka 117—Mission for a Killer) from 1965, in which secret agent OSS 117 travels to Rio to stop former Nazis who’ve developed a drug from a rare Amazon flower (Mavis 2001). Within the movie chemistry universe, the orchid toxin structure is most similar to the anticancer agent “peak 37” in Medicine Man (1992), described in chapter 9. Sean Connery, who stars in Medicine Man, had been the fi rst actor to portray James Bond. The molecules from both fi lms are fictitious but chemically possible, with interconnected hexameric rings, an ester, and at least one chiral carbon with indeterminate stereochemistry. The Seven-Per-Cent Solution (1976) Distribution company: Universal Pictures Director: Herbert Ross Screenwriter: Nicholas Meyer, from his 1974 novel The Seven-Per-Cent Solution: Being a Reprint from the Reminiscences of John H. Watson, M.D., which he based on the characters created by Arthur Conan Doyle Short description: Dr. John Watson tricks Sherlock Holmes into meeting with Sigmund Freud in Vienna to confront his cocaine addiction MPAA rating: PG Plot description: In a short scene at 11:30, Dr. John Watson (Robert Duvall) tells his wife that Holmes (Nicol Williamson) must lose his cocaine addiction. He says he intends to seek help from a man in Vienna named Sigmund Freud (Alan Arkin), who wrote an article about addiction in Lancet (a British medical journal). Next, Watson tricks Holmes into traveling to Vienna to meet with Freud. Once there, Freud uses hypnotism to quench his desire for cocaine and learns that Holmes is suppressing a secret about his father, who murdered his mother and then committed suicide. In the scene at 36:00, Watson explains to Freud that Holmes started taking cocaine to relieve the ennui between cases. Freud responds that his own colleague died of cocainism last year and that he, Freud, was partly responsible. It prompted him to write the paper in Lancet. Freud believes that Holmes took cocaine for a reason other than ennui. Holmes is ultimately cured but is depressed until one of Freud’s former cocaine patients ends up in the hospital after she narrowly escapes being murdered. Commentary: The story of Freud and his morphine addict colleague who died of cocaine abuse was summarized in chapter 2. The Andromeda Strain (1971) Distribution company: Universal Pictures Director: Robert Wise Screenwriter: Nelson Gibbing, from the same-titled 1969 novel by Michael Crichton
216
ReAction! Chemistry in the Movies
Short summary: A small team of experts determines the biology and chemistry of deadly alien material; elemental analysis determines the composition of its two parts MPAA rating: G Plot description: A military intelligence satellite crashed near the small town of Piedmont in the New Mexican desert. On February 5, 1971, two soldiers who are looking for the satellite discover that the town’s inhabitants are all dead except for a crazed man in a white suit. The audience watches the officers at the commanding base listen as the two soldiers describe the elderly man’s approach. The soldiers scream and the communication ends. On day 2, Dr. Jeremy Stone (Arthur Hill) and a surgeon don full body suits before they are lowered into the town from a helicopter. They find that the blood of the dead bodies is crystalline, that a baby survived, that several elderly people survived for a while but committed suicide, and that the crazed elderly man in a white suit is just barely alive. They bring the satellite, baby, and elderly man to a secret underground facility located in Nevada. It is disguised as an agricultural experiment station and was designed by Dr. Stone (a Nobel laureate and twice past president of the National Academy of Sciences) many years ago for the express purpose of studying dangerous exobiologicals brought back from the moon or other missions. After the small team of scientists has assembled and cleansed themselves, they set out to examine the satellite on days 3 and 4. At 58:00, Dr. Stone tells them they’ll work in three stages: detection, characterization, and control. In one of the detection steps, Dr. Stone and Dr. Ruth Leavitt (Kate Reid) use a remote camera to examine the satellite’s surface methodically. They fi nd an embedded black pebble with green specks. At 440× resolution, one of the green specks is smooth and occasionally gives a flash of yellow. It is toxic, alive, and growing. In the scene at 1:33:00, Dr. Leavitt subjects the black and the green objects to mass spectral analysis. The elemental percentages of the black object are H 21.07, C 54.90, O 16.00, with Si being just off the edge of the fi lm frame. Provided with this information, one of them says that the “black rock is something like plastic.” Before we see the next readout, Dr. Reid says the “green object is even simpler.” The computer shows elemental percentages of H 27.00, C 45.00, N 5.00, and O 23.00. Commentary: The opening credits list cooperation with the American Society for the Prevention of Cruelty to Animals, which is necessary given the distressing images of the dying monkey and rats shown in the movie. Since the total elemental analysis must equal 100.0%, the black object must have 8.03% Si. With this last bit of information, it is possible to calculate the elemental composition of the black object as C32H146O7Si2. Unfortunately, this composition does not follow the known laws of the chemical bonding since there are too many hydrogen atoms for the number of carbons. For instance, a saturated hydrocarbon with 32 carbons
Hard Science = Hard Evidence
217
would have 2n + 2 hydrogens, or 66. The only way to use the computer’s readout to solve the structure is to assume that the reported percentages are actually the elemental composition, which would give C55H21O16Si8. The relatively high amount of silicon, along with the presence of oxygen, suggests this is a siloxane, a special type of plastic. It is possible to use this information to suggest a structure that differs from the readout by a single hydrogen atom (figure 7.5A). The green object is not simpler in the sense of fewer elements but in the use of whole numbers for the percentages. The elemental percentages yield the following composition: C21H150N2O8. Again, this does not follow the known laws of the chemical bonding because a saturated hydrocarbon with 45 carbons could only have as many as 2n + 2 hydrogens, or 92. When the computer readout is assumed to be composition, we get C45H27N5O23. In this case, the number of hydrogens to carbon is very low and the ratio of oxygen to carbon is quite high, suggesting a high level of unsaturation and a polyaromatic structure. The presence of the amines is consistent with its being an alkaloid (figure 7.5B). Our Man Flint (1966) Distribution company: Twentieth Century Fox Director: Daniel Mann Screenwriters: Hal Fimberg and Ben Starr Short summary: Chemical analysis of a deadly poison dart shows traces of ingredients for bouillabaisse, so Derek Flint searches for the perpetrator in the restaurants of Marseilles
A
B R1 R3
Si
OH
OOH
R1 O
Si 7
R2
OOH
N R4
OOH N
R2
OOH N OOH
R1 =
C H
C
O
R2 =
C
C
C
C
H
R3 =
C
C
C
C
H
R4 =
C H
C H
C H
O
HO HOO
OOH
HOO
N
OOH
N HOO OH
Figure 7.5. Possible structures for the “black object” with the formula C55H20O16Si8, a linear oligosiloxane (A) and the “green object” with the formula C45H27N5O23, an alkaloid (B).
218
ReAction! Chemistry in the Movies
Plot description: Derek Flint (James Coburn) is a retired agent of ZOWIE (Zonal Organization World Intelligence Espionage) who lives with four beautiful young women who cater to his every need. His life of ease and luxury is disrupted by three scientists who are manipulating the world’s weather. They formed a utopian organization called Galaxy in which men are drones who do the work and women are “pleasure units.” The trio believes the world’s governments have failed because they do not cooperate with one another, so they demand that all the world’s governments capitulate to them. It becomes the duty of Lloyd Cramden (Lee J. Cobb), president of ZOWIE, to convince Flint to come out of retirement to deal with Galaxy. While making his request, Cramden is hit by a poison dart blown by Gila (Gila Golan). Flint saves Cramden’s life by cutting open the injection site and sucking out the poison with only three seconds to spare. After Flint accepts the job, Cramden tries to explain various gadgets to him in the scene at 25:15. For instance, one suitcase contains packets labeled C7H6O3, H3H4H6O8, and C3H4O1 and is capable of 53 functions. Flint rejects them all as primitive compared to the 72 functions he’s developed for himself in the form of a cigar lighter. The report fi nally arrives with the news that the dart poison was curare and that the full chemical analysis of the dart also revealed the presence of garlic, saffron, and fennel. Flint asks for their proportions and is told “Two bulbs of garlic, a pinch of saffron, and a dash of fennel.” Flint says those are the ingredients for bouillabaisse as it is prepared in Marseilles. He fl ies his own plane to Marseilles in search of the restaurant. Commentary: With thoughts of impossible chemical compositions fresh in our minds, the proportions of bouillabaisse provide the clue that sends our spy to his next destination. D.O.A. (1949) Distribution company: United Artists Director: Rudolph Maté Screenwriters: Russell Rouse and Clarence Greene Short summary: Accountant Frank Bigelow has been poisoned with a deadly dose of luminous toxin; he has only a short time to fi nd out why because there is no antidote Plot description: As the credits roll, the camera follows a man walking down a corridor at night to the police department. When the credits end, he says, “I want to talk to the captain. I want to report a murder.” Policeman asks, “Where did it take place?” Man says, “San Francisco.” Policeman asks, “Who was murdered?” Man says, “Me.” The rest of the movie is a flashback. Frank Bigelow (Edmond O’Brien) is an accountant from Banning, California, who takes a vacation in San Francisco to cool down the
Hard Science = Hard Evidence
219
relationship with his secretary and girlfriend Paula (Pamela Britton). Once in San Francisco, he checks into the St. Francis Hotel, where he joins a group of conventioneers as they head to The Fisherman jazz club. He leaves them to go to the bar, where he meets a woman named Jeanne. While talking to her, a mysterious stranger swaps his drink. Back in his hotel room, he fi nds flowers and a note from Paula, so he tears up Jeanne’s phone number in guilt. The next morning, Bigelow doesn’t feel well and enters a Medical Building to the sound of haunting theremin music. The doctor says his heart is in good condition and his blood pressure is OK. Then, the lab report is ready and the doctor says he’s been poisoned with “luminous toxic matter.” There is no antidote. He has a day, possibly a week or two at the outside. Bigelow says the doctor is crazy and rushes out the door. In the 3.0-minute scene at 30:00, Frank runs into Southern Hospital University and asks the doctor to examine him for “luminous toxin.” In the next scene, the doctor emerges from an adjacent room to tell him “Yeah, you’ve got it alright” and then turns off the light to show him a test tube that glows. Bigelow spends the next hours discovering who killed him and why. Commentary: We never learn the name of the luminous toxin, but it is chemically exciting when Bigelow discovers he was murdered because he notarized the bill of sale for iridium. Iridium is among the rarest nonradioactive elements. It is part of the platinum group of elements and was also mentioned in Edison, the Man (1940). The movie’s endnote states that “luminous toxin” is a descriptive term for an actual poison according to Edward F. Dunne, M.D., who is listed in the credits as a technical adviser. He may have also helped them set up the scene with the glowing test tube. In 2004, the U.S. Library of Congress selected this fi lm for preservation in the National Film Registry. The Exploits of Elaine (1914) Distribution company: Pathé Films Codirectors: Louis Gasnier and George B. Seitz Screenwriter: Arthur B. Reeve Short summary: Elaine Dodge and Craig Kennedy fight the Clutching Hand, who killed Elaine’s father Plot description: Elaine Dodge (Pearl White) is engaged to Perry Bennett (Sheldon Lewis) and is the daughter of Taylor Dodge (Riley Hatch). In the fi rst episode, Taylor receives a letter that identifies a traitor named the Clutching Hand in his company. To keep it secret, he places blank pages into the real envelope and locks it in the safe. Then, he places the real letter pages into a blank envelope that he hides in a secret compartment in the wall. A masked man, later identified as the Clutching Hand, enters the basement to electrify a floor grating in Taylor’s study. When Taylor steps
220
ReAction! Chemistry in the Movies
on it, he is killed instantly. The Clutching Hand enters Taylor’s study, opens the safe, takes the letter, sprinkles some powder onto a bust on top of the safe, and leaves. Once outside, he realizes he has been duped. Cut to Craig Kennedy (Arnold Daly) in his apartment fi lled with chemical bottles and glassware. When Elaine calls him, he rushes to her house, where they discover her dead father. Kennedy immediately examines the room for clues. He uses some black powder to collect fi ngerprints from the bust, only to discover they are his. The episode ends. In the second episode, the Clutching Hand enters Elaine’s bedroom at night, injects scopolamine into her arm, and then directs her to write a letter to Kennedy. When Kennedy receives the letter from Elaine, he is suspicious, and when he sees her, she denies having written it. She agrees to an experiment in which Kennedy injects her with scopolamine and has her recount her movements of the night before. As she speaks, we see that the Clutching Hand was foiled again, this time by Elaine’s unfamiliarity with her father’s secret hiding place. When she finishes her story, a letter is pushed under the door and the episode ends. Commentary: This was the fourth serial to be produced and was the largest grossing fi lm of its time. The fi rst three serials were Whatever Happened to Mary? (1913), The Adventures of Kathlyn (1913), and The Perils of Pauline (1913) (Lahue 1968; Barbour 1970; Rainey 1999). All three featured women in the leading roles and were successful because young city women would step into a theater on their way home from work to watch the latest 15-minute episode. This serial spawned at least three short spoofs: Pimple’s Clutching Hand (1916) from Britain starring the rubber-faced Fred Evans, Tubby’s Clutching Hand (1916) from Britain starring the pudgy Johnny Butt, and The Mystery of the Leaping Fish (1916) from the United States starring Douglas Fairbanks Sr. Of these, only Leaping Fish survives. Even though all 14 of the roughly 15-minute chapters of Exploits are known to exist in various fi lm archives, they have never been released in their entirety on video. Chapters 1, 2, 3, and 9 are available on Google Video, although chapters 1 and 2 have French intertitles. In 1994, the U.S. Library of Congress selected this fi lm for preservation in the National Film Registry.
8 Chem 101 Learning by Doing
CHEMISTRY’S SYMBOLIC LANGUAGE This chapter stands at the midpoint of the “bright side” of this book. Its oppositional partner on the “dark side” is chapter 3 on chemical arsenals. Thus, education and war making present a core contrast in the uses of chemical knowledge in the movies. We have only to think of the closed world of Dr. Mabuse, symbolized so well by the writings in his notebook, as described in chapter 3. His inward, secret scribblings speak of outward, villainous purposes. In juxtaposition, the writing on the blackboards of this chapter’s movies is exposed; it is to be seen (table 8.1). This writing represents the open, cool, dispassionate transmission of facts in symbolic chemical language. But a little something more has slipped through in this section’s movie example. When we read its blackboard, we see the explosive nature of knowledge itself. “You have to see and read a movie at the same time,” says Harvard professor Tom Conley, whose creative fi lm analyses leave readers with an appreciation of just how rich cinematic art can be (Savisky 2006). In his book Film Hieroglyphs, Conley considers writing as it appears in movies. He shows it is often an incidental and uncontrollable element that inserts itself into our movie-watching experience, leaving us to ponder nonnarrative written material observed on the movie sets. Conley calls the points where story, image, and writing are at odds with one another “ruptures.” Graphically interrupting the flow of the moving image, these points provide “slide areas” for analysis and new insights (Conley 1991). This section examines chemical notations appearing in our movie examples, and how they provide opportunities for an enhanced reading of these movies. The blackboard in The Affairs of Dobie Gillis (1953), a musical campus comedy, ruptures the movie like no other blackboard in movie history (figure 8.1). But fi rst, let us set the scene. Pansy Hammer (Debbie Reynolds) attends Grainbelt University to “work, work, work.” As she and her lab mates begin their fi rst laboratory exercise, she gleefully tells them she really likes chemistry. After a few moments, the professor announces, “Don’t let the bubbles come too fast” and then Hammer’s experiment explodes. This only heightens her interest in chemistry, however, and she 221
222
ReAction! Chemistry in the Movies
Table 8.1. Movies featuring characters as students, teachers, and professors a Title (Year) The Prince and Me (2004) Harry Potter and the Chamber of Secrets (2001) Tortilla Soup (2001) October Sky (1999) The Saint (1997) The Nutty Professor (1996) The Nutty Professor (1963) The Affairs of Dobie Gillis (1953) It Happens Every Spring (1949) Madame Curie (1943) The Chemist (1936)
Educational Institution
Laboratory Scene
Blackboard Scene
University Boarding school
Teaching Greenhouse
Teaching lab None
High school High school University University University University University University University
None Teaching Professor’s Professor’s Professor’s Teaching Professor’s Professor’s Professor’s
Classroom Teaching lab Classroom Classroom Classroom Teaching lab Classroom Professor’s None
a The laboratory scenes take place in a teaching laboratory, a greenhouse, or the professor’s research laboratory. The blackboard scenes take place either in the classroom, teaching laboratory, or in the professor’s research laboratory.
Equivalent Weights H2SO4 + Zn (97) (65.2)
ZnSO4 + H2 (160.2) (2)
KClO3 chlorate KCl chloride
Avogadro’s 6.02 x 1023 atoms 2 H2O
2 H+ + 2 OH2 H2 + O2
Figure 8.1. Blackboard notes in the teaching laboratory of The Affairs of Dobie Gillis (1953).
causes four more explosions throughout the movie. Even though none of the characters ever refers to it, there is a laboratory blackboard that contains information relevant to the narrative. In fact, the blackboard information is so critical that the camera pans across it at least once in every laboratory scene, rupturing the flow of the narrative because this information never changes throughout the semester. To understand the cause of Pansy Hammer’s explosions, we need to read the blackboard. At the top left hand side, we fi nd an equation that
Chem 101
223
demonstrates the principle of “Equivalent Weights,” itself is a demonstration of the Law of Conservation of Mass, which states that matter cannot be created or destroyed during a chemical reaction (see the next section). The reaction equation is for sulfuric acid mixed with zinc metal to generate zinc sulfate salt and hydrogen gas. The upward arrow next to the hydrogen gas indicates it is bubbling out of solution and into the atmosphere. Every student of chemistry knows that hydrogen gas is highly combustible. They also know that some metals react with acids so fast that they generate hydrogen gas and enough heat to catch it on fi re as it evolves. Some students are disappointed to learn during their laboratory exercises that zinc reacts too slowly and without releasing enough heat to cause spontaneous combustion. On the top right-hand side of the board is a demonstration of the naming rules for fully oxidized polyatomic chlorate versus the monoatomic anion chloride. This may refer to sulfate in the “acid plus zinc” reaction, in which monoatomic sulfur ion would be called sulfide. The presence of Avogadro’s number below the reaction equation is used to connect macroscopic and nanoscopic events because it converts mass into numbers of molecules, ions, or atoms. Finally, in the lower right-hand corner are two reactions involving water. In the fi rst, water dissociates to its ions, hydrogen and hydroxide. Today, we always use a set of equilibrium arrows to show that only a small fraction of water dissociates in this way. In the other reaction, water is separated into its elements of hydrogen and oxygen, which are violently explosive when combusted. The blackboard ruptures history to make the viewer conscious of the present because its information differs in a number of ways from current practice. For many decades, students were taught to use an ascending arrow to indicate that gas is evolved during a reaction and to use a descending arrow to indicate that a solid (or precipitate) was formed. Today, students are taught to indicate the states of all reactants and products with letters in parentheses as follows: solids (s), liquids (1), gases (g), or aqueous (aq; i.e., taking place in water). Arrows are now used only to indicate chemical transformation and not movement of a chemical from one physical state to another. To determine when the transition to parenthetical letters took place, we surveyed the many editions of the textbook General Chemistry coauthored by Professor Henry Holtzclaw, Jr. from its 1968 third edition until its 1991 ninth edition. The switch to letters occurred in the 1980 sixth edition (Nebergall et al. 1980), coinciding with the height of the textbook’s popularity. Interestingly, the use of arrows was part of the chemical culture and never codified by the International Union of Pure and Applied Chemistry (IUPAC) until 1988, when it endorsed the use of parenthetical letters (Mills et al. 1993). The result of the IUPAC endorsement is that every general chemistry textbook now uses the letters and not the arrows in this context. Another difference from current practice is the emphasis on “significant figures,” mentioned in chapter 7. Briefly, the number of digits in a
224
ReAction! Chemistry in the Movies
value indicates the degree to which a measurement is reliable. Specifically, the mass for one molecule of sulfuric acid is easily calculated by summing the masses for two atoms of hydrogen, one atom of sulfur, and four atoms of oxygen, each of which is known to high precision and are listed on most periodic tables, to arrive at a mass for sulfuric acid of “97.0 amu” and not just “97.” The use of “97” would suggest we are not confident the next lower digit is a zero and that we don’t realize the uselessness of a value without a unit (in this case “atomic mass unit” or amu). As a demonstration of the principle of “equivalent weights,” the reaction is quite suitable in that each side’s total mass is 162.2 amu. There is a fi nal historical rupture in that the masses given on the blackboard differ from those that would be obtained using the masses found on a current periodic table. Today, the masses would be 98.1 amu for sulfuric acid, 65.4 amu for a zinc atom, 161.5 amu for zinc sulfate, but still 2.0 for diatomic hydrogen. This historical rupture occurs because Dobie Gillis was released in 1953, whereas the defi nition of the “standard mass” changed in 1961 from 1/16 the mass of natural oxygen to 1/12 the mass of carbon-12. The events that led to this change were set in motion by German nuclear chemist J. Mattauch (Mattauch 1958). In a 1958 paper, he discussed the pros and cons for several possible mass standards and concluded that C-12 would be the most suitable. C-12 has six protons and six neutrons to satisfy the physicists by providing a rigorous balance of protons and neutrons. It also gave a value for the atomic mass unit that was similar to one that chemists had been using. The unified atomic mass unit (De Bievre and Peiser 1992) was defi ned as “1/12 the mass of one atom of C-12” by physicists in 1960 when it was adopted by the International Union of Pure and Applied Physics, and by chemists in 1961 when it was adopted by the IUPAC. This choice meant that the atomic mass unit was on fi rm scientific footing, that the two sciences could communicate with a common meaning for atomic mass unit, and that the new atomic mass unit was slightly smaller for both types of scientists. In his atom theory of 1801, John Dalton proposed that chemists should use a value called “atomic mass unit” (amu) equal to the mass of one atom of hydrogen. His expectation was that each elemental atom would have the mass of a whole number since he assumed each was perfect. He didn’t know that atomic nuclei were composed of protons and neutrons, let alone that the mass of a proton differs slightly from the mass of a neutron. He also didn’t know that natural pure elements are mixtures of isotopes, each with different numbers of neutrons that add mass but don’t change the chemical properties. As analytical methods improved throughout the nineteenth century, it became clear that some pure atoms didn’t have masses that were whole numbers. For instance, chlorine was close to 35.5 amu. Francis Aston, chemistry and physics professor at Cavendish Laboratory in Cambridge, invented the mass spectrometer in 1919 (Beynon and Morgan 1978). It rapidly and accurately separated the isotopes of atoms
Chem 101
225
according to their relative atomic masses. He was awarded the 1922 Nobel Prize in Chemistry for identifying the isotopes of neon and chlorine. After a great deal more work, he proposed that the atomic mass unit defi nition be changed to 1/16 the mass of the isotope O-16 because it has an equal number of protons and neutrons (8 and 8). Since neutrons have a slightly larger mass than protons, this meant that all of the earlier chemical literature since Dalton was wrong by the same fraction. Even though physical societies readily adopted his suggestion, none of the chemical societies did. The chemists didn’t want to lose their link with their welldeveloped literature, so they decided to use 1/16 of the average mass of one natural oxygen atom (which is a mixture of the natural oxygen isotopes) so that it wouldn’t change the value as much. It was a nonscientific decision that caused problems for decades when physicists and chemists compared their data.
LAVOISIER’S THE ELEMENTS OF CHEMISTRY (1792) In Tortilla Soup (2001), the fi rst equation we encounter involves zinc metal plus acid to make zinc chloride salt plus hydrogen gas: Zn +_HCl o ZnCl2 + H2. The chemical equation is not balanced, but it is fairly easy to see that we need two molecules of HCl for every one atom of zinc. Only then will there be two atoms of H and two atoms of chlorine in both the reactants and products. After we have balanced the atoms, we can calculate the total mass of the combined reactants to show it is the same as the combined mass for the products. In other words, we can use a theoretical measuring device to show their masses also balance. In Tortilla Soup, the chemistry teacher’s life becomes balanced only when she falls in love with the baseball coach. It takes a week of lectures in the chemistry classroom for an instructor to explain how chemists use the word “balance” as a metaphor, a verb, and a measuring device all at once. The scientific love of searching for mathematical relationships in nature began with René Descartes. In 1637, he published La Géométrie, wherein he created analytical geometry (using a grid with coordinates to map things or movement) to show how mathematical models can be used to represent natural phenomena. La Géométrie was bundled with three other texts that began with his Discours de la Méthode [Discourse on Method]. His Méthode set the guidelines for rational inquiry into all things natural and helped initiate the Age of Reason (Lafleur 1956). He proposed that all people possess enough reason to be able to discover the truth for themselves, that the world is rational and comprehensible, and that one should make empirical observations rather than defer to authorities. His correspondence theory of knowledge proposes that ideas are correct when they are good copies of external reality. Fifty years later, Isaac Newton used the Cartesian method to develop calculus to describe the laws of motion and gravity. Newton’s demonstration was so powerful
226
ReAction! Chemistry in the Movies
that it cemented the idea that if the model did not conform to nature, it was necessary to modify the model until it did. It took another 100 years before Antoine Lavoisier showed how mathematics applies to chemistry. He used a balance to develop the Law of Conservation of Mass and give birth to modern chemistry. Antoine Lavoisier was born into a noble family in Paris in 1743 (Poirier 1996). He received the best possible education and became especially interested in chemistry. In 1768, he was elected to the French Academy of Science primarily because of his family’s influence rather than his accomplishments. The next year, he used the inheritance from his mother to buy one of the 40 seats on the Ferme Générale (literally general farm, but better translated as General Tax-Farm). These 40 fermieres générales (or general farmers), as they were called, purchased the right to collect all manner of taxes on behalf of the crown, a portion of which generated large sums of money for themselves. At age 26, he married 14-year-old Marie Paulze, daughter of another member of the Ferme Générale, who brought an enormous dowry. She became his recording assistant in the laboratory, learned how to paint from Jacques-Louis David, and hosted many dinners that the world’s scientific elite were eager to attend. In November 1772, Lavoisier made the critical observation that sulfur and phosphorus gained mass when they combusted (Perrin 1989). By February 1773, he wrote in his notebook that he planned “a long series of experiments” dealing with all reactions in which the mass appears to change. His team began with the metal combustion reactions and fi nished with acid–base salt-producing reactions. They established the purity of all reactants and yield of all products using the mass balance for solids, the aerometer to measure the specific gravity of liquids, and the pneumatic trough for gas volumes (Holmes 1998). By this careful work, he discovered that a gaseous element from the air was consumed during combustion. He called it “oxygen” (for acid generator) because it generated acids from the non-metals. While this was occurring, in 1782, Dijon chemist Guyton de Morveau proposed using scalar naming rules for compounds. The idea didn’t catch fi re until he copublished it in expanded book form in 1787 with Lavoisier and others titled Méthode de Nomenclature Chymique (Anderson 1984). They proposed to name compounds according to the names of their most basic elements, such as sulfate of lime, and calcareous nitrate. Prior to this, compounds were named according to their properties, their source, their most prominent discoverer, and so forth, and often the same compound had multiple names. This new nomenclature system allowed Lavoisier to fi nish his long study in a way that separated chemistry completely from its alchemical roots. In the preface to his 1789 book Traité Élémentaire de Chimie [Elements of Chemistry], Lavoisier wrote that physical science consists of “three things: the series of facts which are the objects of the science, the ideas which represented these facts, and the words by which these ideas are expressed.” In other words, he combined theory with practice
Chem 101
227
in chemistry for the fi rst time. Part III of the book contains descriptions of the procedures and Madame Lavoisier’s images of the apparatus and experimental design. Five years later, Lavoisier’s neck felt the blade of the guillotine during the Reign of Terror in May 1794 because he was a member of the Ferme Générale. Even before the French Revolution was over, scientists from around the world began courting his now middle-age widow. The next chronological step in creating our chemical equations goes to Swedish chemistry professor Jöns Jacob Berzelius. To keep track of the hundreds of reactions he performed, he began using the one- or twoletter element symbols we use today: Ca, Mg, Cl, O, Se, Th, Ce, and so on. Berzelius earned his medical degree in 1801, earned his living as an assistant pharmacist while he experimented with electricity, discovered cerium, and then secured a permanent position at the Medical School of Stockholm in 1808 (renamed the Karolinska Institute in 1810) (Melhado 1981; Rocke 2008). During his fi rst year as a professor, he combed the chemical literature for theories and experiments to present in his 1809 chemistry textbook for medical students. He realized he could use experimental stoichiometry to test the theory of mineral salts, so he carried out every known stoichiometric reaction and analyzed them using the best quantitative methods. He summarized his work in a book-length journal article that was soon translated into German. In 1813, the indefatigable Berzelius turned his attention to compounds from animals and vegetables but was frustrated they did not react stoichiometrically with each other or with inorganic compounds (Klein 2003). The only exception was the organic acids. He felt this nonquantitative behavior would forever prevent them from being characterized and caused him to make the sharp distinction between inorganic and organic compounds that we still use today. Organic compounds actually yielded to study but only after other chemists isolated pure organic compounds and developed noncombusting transformative reaction methods. Nevertheless, it was the high compositional similarity of different organic compounds that caused him to adopt Latin letters as a shorthand for elements and superscripts for stoichiometries (today we use subscripts). In 1826, he published his fi nal revision of atomic weights and molecular formulas for all known compounds. The power of this information became immediately obvious to organic chemists, and they soon adopted his naming system (Klein 2003). Berzelian formulas serve as a useful surrogate name, provide the empirical stoichiometric relationships of the elements, and provide a method to calculate the molecular weight from the atomic weights. A practical reason for the rapid adoption of his system was that fewer than 100 organic compounds were known in 1800, but several thousand by the 1850s. Since most were produced by artificial means, organic chemistry became synonymous with carbon chemistry. The last component of the Berzelian names to be established was the order in which the elements are listed. In
228
ReAction! Chemistry in the Movies
1900, Edwin Hill of the U.S. Patent Office developed the indexing system we still use today (Hill 1900). For organic compounds, the number of carbon atoms is given fi rst, the hydrogen atoms second, and then all other atoms in alphabetical order. The elemental composition of inorganic compounds is listed alphabetically. The barbed arrow was the last component of the chemical equation to be adopted (Jensen 2005). The fi rst linear chemical equations appeared in the seventh edition (1864) of George Fownes’s textbook Manual of Elementary Chemistry, Theoretical and Practical. He used an equal sign to separate the reactants from the products. The fi rst use of the double barbed arrows to show reversible equilibrium is found in the 1898 published lectures of Jacobus van’t Hoff, indicating a pedagogical origin for this most common convention.
THE SENSE OF WONDER AND THE MOST BEAUTIFUL CHEMICAL EXPERIMENTS “Sense of wonder” is a phrase that describes the feeling evoked by scientific discovery (Gruber 1981), by reading science fiction (Clute and Nicholls 1993), and by children encountering nature for the first time (Carson 1956, 1965). In each case, the sudden fresh thought causes an individual to reevaluate his or her ideas. Sense of wonder is related to the scientific Eureka moment, the psychological “Aha,” and the religious epiphany. Publisher Hugo Gernsback coined the phrase “sense of wonder” to defi ne the feeling he hoped writers would invoke in readers of his pulp magazine Amazing Stories, founded in 1926 (Clute and Nicholls 1993). He also coined the name “scientifiction” to defi ne this genre but then relaxed it to “science fiction” in 1930. The editors of the 1993 edition of The Encyclopedia of Science Fiction suggest that “sense of wonder” is unique to their genre. The essence is that something had not been properly understood until the transformative moment in the text. The Sense of Wonder, with glossy nature photos by Charles Pratt, was a book published in 1965 from an essay that Rachel Carson wrote for Woman’s Home Companion in 1956 (Carson 1956, 1965). In it, she discussed the nature walks she had taken with her nephew when he was a young boy. She was surprised by the things he learned from their walks. She had not consciously taught him anything but only pointed out the things that interested her the most. She believed that a child’s sense of wonder is fueled by nature walks with an adult, simultaneously experiencing the sights, smells, feel, and sounds. In November 2002, the editor of Chemical and Engineering News, the news magazine published by the American Chemical Society (ACS), asked readers to nominate entries for the “ten most beautiful chemical experiments” (Freemantle 2003). Beautiful was defi ned as “elegantly
Chem 101
229
simple but significant.” Note that scientists use the word “elegant” to describe experiments that are ingenious and simple. Seventy readers responded with 55 discoveries. Twenty-five entries were nominated more than once and were ranked by a panel of chemists and science historians. The top entry was also the most nominated entry: Pasteur’s 1848 separation of asymmetric tartrate crystals (briefly described in chapter 6). The second entry was Lavoisier’s metal oxidation experiments that generated theories of combustion and oxidation (described above). The tenth entry was the Curies’ isolation of radioactive elements (see description of Madame Curie, below). Stimulated by the ACS poll, the Royal Chemical Society commissioned science writer and editor Philip Ball to develop a list of “ten beautiful experiments in chemistry” based on his personal study of chemical history (Ball 2005). He chose experiments that were deemed significant at the time they were made, that did not involve a great deal of serendipity, and that contained different aspects of beauty. He ordered his list chronologically and included the conceptual simplicity of Pasteur’s crystal separation and the Curies’ patience while conducting their experiment. Thus, it appears the two ways of identifying the most beautiful and wondrous experiments and discoveries in chemistry have produced two stories in common: Louis Pasteur’s chiral crystals and the Curies’ radioactive elements. The isolation of radioactivity by the Curies is told in Madame Curie (1943), but Pasteur’s chemical discovery is omitted from his biographical picture The Story of Louis Pasteur (1936). Nevertheless, from this lesson we’ve learned that these two discoveries were beautiful at the time they were performed, have endured beyond their historical context, and are instructive.
MOVE OVER “SAFETY FIRST” Few movies show any concern for laboratory safety. A recent exception is Harry Potter and the Chamber of Secrets (2002) in which the students are told to make sure their ear muffs are fi rmly in place before they repot their young mandrakes. In reality, chemistry instructors pay a great deal of attention to the safety of their charges. Students are required to wear safety goggles, remove dangling jewelry, and sometimes even to wear gloves to reduce the probability an accident will happen. Goggles don’t obstruct their vision but put them in the mood for observing and thinking about their observations. Even though blackboards are for teaching and laboratories are for learning, it is important to realize that teaching is more than talking and learning is more than listening or doing. Work is required from both teachers and students in the training process—future scientists need training in more than just safety. They have to learn to adapt to the continually changing landscape of our mobile, global, interconnected culture. This
230
ReAction! Chemistry in the Movies
is an enormous task with fuzzy boundaries, so it is helpful to start with an ideal goal. Danish science education professor Kathrine Eriksen proposes that college chemistry majors should be able to act as social agents outside their narrow academic context (Eriksen 2002). She suggests that chemistry students need to reflect on how to do the right thing and why, to understand the nature of chemical knowledge, and be cognizant of the role chemical products play in society. She wrote, “The tools for reflection must be found outside the ‘world’ of natural science and can be for example historical, psychological, or sociological analyses of the scientific enterprise.” To help students integrate learning with their future lives, the University of Nebraska–Lincoln (UNL) approved a systemwide general education program in 2008 called Achievement-Centered Education (ACE; see www.unl.edu/ous/). According to its statement, “ACE enhances the undergraduate experience by providing exposure to multiple disciplines, complementing the major, and helping students develop important reasoning, inquiry, and civic capabilities.” At UNL, students are required to take courses meeting 10 specific learning outcomes falling under four major objectives: skills, knowledge, responsibilities, and integration. The integration outcome is a creative or scholarly product “that requires broad knowledge, appropriate technical proficiency, information collection, synthesis, interpretation, presentation, and reflection.” For chemistry majors, this will be achieved by modifying an advanced lecture-andlaboratory analytical or inorganic course to ensure that all these objectives are components of the laboratory project reports. The reports will be written using the guidelines provided by the relevant division of the American Chemical Society to ensure technical proficiency and an appreciation for presentation. The background section will require collecting relevant information from the literature whereas the results section will require collecting data at the lab bench. While writing their reports, the students will have to synthesize and interpret their data in light of the published information they collected. The discussion section will provide space for reflection on the entire enterprise. Harvard psychologist Howard Gardner, renowned for his 1980s groundbreaking theory of multiple intelligences, has identified “five minds for the future” that can be cultivated “at school, in professions or in the workplace” (Gardner 2006). In his estimation, these minds— disciplined, synthesizing, creating, respectful, and ethical—are at a “premium” in today’s world and will be needed even more in the future. Individuals will need to consider their roles both as workers and as citizens to develop ethical minds. Gardner suggests four “signposts” that would help realize this. The fi rst is “mission,” or having a specific knowledge of the goals to be achieved from an activity. The second is “models,” or exposure to individuals who do good work. The third is the “mirror test—individual version,” to fi nd out if, on reflection, the individual is going forward in a way he or she would personally approve; the fourth is
Chem 101
231
“mirror test—professional responsibility,” which involves an obligation to monitor peers and hold them accountable. It seems Dr. Jekyll’s idea of bringing the mirror into the lab (see chapter 1) enjoys a new currency when seen from this perspective. We are asking him now, however, not only to verify that his experiment worked, but also to look again into that mirror to see what effect it will have on himself and society.
THE ARCHETYPE MOVIE: THE NUTTY PROFESSOR (1963) Distribution company: Paramount Pictures Director: Jerry Lewis Screenwriters: Jerry Lewis and Bill Richmond Short summary: Nerdy Arizona State Chemistry Professor Julius Kelp creates an elixir that transforms him into chauvinist swinger Buddy Love Plot description: We are prepared for the movie’s chemical theme by a montage of colorful chemical demonstrations presented to an attentive class of students during the opening credits. Solutions boil out of graduated beakers onto tables or the demonstrator’s unquavering hand. The demonstrator’s face is never visible to the camera. As the credits end, the last demonstration explodes, a meeting is disrupted, and the university president calls for Kelp. In the meeting between the disheveled über-nerd Dr. Julius Kelp (Jerry Lewis) and President Mortimer S. Warfield (Del Moore), we learn that Kelp has been with the university for two years and 22 minutes. It is the fi rst day of class at Arizona State University, and he has destroyed his classroom for the second time. The president reminds Kelp of his promise two years ago to desist from engaging in further research. Six weeks after Kelp’s arrival, he was working on a gasoline additive composed of three parts carbon, five parts hydrogen, one part nitrogen, and three parts oxygen. Kelp says, “Yes, nitroglycerin.” Back in the classroom, Dr. Kelp is about to begin his lecture when a football player wishes to be excused for practice. Kelp refuses, so the student forcibly lifts him into a glassware cabinet and class is dismissed. Stella Purdy (Stella Stevens) helps him climb down but drops her copy of Allure magazine. Kelp picks it up for her, but she is gone. In it, he reads an advertisement for a muscle-building gym. After a quick series of body building gags, he visits Dr. Levee (Milton Frome) in the university’s medical center to say he has actually lost weight after six months of training. Levee asks him about his next step, and Kelp says, “I’m a chemist so I’d like to approach it chemically.” On the doctor’s bookshelf he fi nds a book titled Man, Muscles, What Are They, and How to Get Them. It says that man’s body can grow by the addition of elements in the same
232
ReAction! Chemistry in the Movies
way that man’s knowledge grows by study. Kelp is intrigued and brings the book back to his classroom to read. On the blackboard behind him are the oxidation half reactions for six oxidation states of nitrogen, from nitrate to hydrazine. He is inspired and reads many more books. In his laboratory, the blackboard reaction shows manganate reduction to manganese ion. He writes in his notebook that he has made 13 experiments in seven weeks without the desired result. This causes him to read even more books, except that this time he also drinks milk from a graduated beaker. After Purdy invites him to a midterm celebration at the Purple Pit, he writes at 28:45 that his research is complete and he will try some of the formula tonight. He sneaks into his lab, mixes up the formula, and drinks it. He convulses, knocks over his glassware, spills colored solutions on the floor, and fi nally gains a hairy arm plus even worse teeth. In an abrupt cut, a man tells the camera the suits will be ready no later than next week. Everyone stares into the camera as it crosses the street and enters the Purple Pit. Inside, the musicians and dancers stop when they see the camera. He is not the monster we observed in the laboratory—he is Buddy Love, dressed in a light blue suit with black accents, thin black tie, and a pink shirt. He has greased black hair, a cool but hyperchauvinist attitude, chain smokes, and can drink alcohol strong enough to knock out a bartender. Commentary: The full color demonstrations during the opening credits end with an explosion, which we are told was due to Kelp’s C3H5NO3 nitroglycerin. After a moment’s reflection, chemists will realize Kelp’s nitroglycerin differs from the more common variety with elemental composition C3H5N3O9 (figure 8.2). This is the fi rst Jekyll and Hyde adaptation to infer that the transformative compound was a real entity. In this case, it was a muscle-building formula with many attributes of androgenic-anabolic steroids (figure 8.3). This movie is also a satiric inversion of the 1931 Jekyll and Hyde in which the “bright” character is an ugly, weak nerd with the simian teeth of Fredric March’s “dark” neanderthalic character. The point-of-view camera shot
O2N O H
NO2
O2N C
C
C
H
H
H
Kelp's Nitroglycerin
H
H
O
O
O
C
C
C
H
H
H
NO2 H
Nitroglycerin
Figure 8.2. Kelp’s experimental nitroglycerin (C 3H5NO3) is missing two nitrate groups compared to common nitroglycerin (C 3H5N3O9).
Chem 101
233
of good-looking but bad Buddy Love buying suits and entering the Purple Pit is an inversion of the opening sequence of March’s philanthropic but manic Dr. Henry Jekyll that begins from Jekyll’s perspective. In 2004, the U.S. Library of Congress selected this fi lm for preservation in the National Film Registry. In 1995, Jerry Lewis wrote about this movie saying, “For ten years I wrote and re-wrote the story” (Neibaur and Okuda 1995). In other words, he began writing in about 1953 when he was fi lming Living It Up, a remake of Nothing Sacred (1937), one of the earliest screwball comedies that starred Fredric March. Did Lewis watch Nothing Sacred (1937) and then March’s 1931 Dr. Jekyll and Hyde at this time? Or, was he inspired by the Russian weightlifting team’s performance at the 1952 Olympics? (They used steroids to win.) Androgenic-anabolic steroids are synthetic compounds with structures similar to the natural male sex hormone testosterone (figure 8.3). Its anabolic effect is that it increases the growth of muscular and skeletal tissues. At its normally low biological levels, its androgenic effect is to increase the development of male sexual characteristics. The four-ring structure of testosterone is found in all steroids and shared by many other hormones. Hormones move from their site of biological synthesis to other cells by way of the bloodstream. Certain cells have proteins called steroid hormone receptors on their outer membranes. There are a few dozen combinations of these receptors, and some of them bind certain steroids better than others. When a steroid binds to its receptor, it stimulates that cell to produce certain proteins so that the cells grow. Several hundred different synthetic steroid hormones have been synthesized, and they each have different effects at different doses. Many hormones don’t effect sexual development but do suppress or stimulate other cells to grow. For instance, the most medicinally important steroid is cortisone, a non-sex-related hormone used to treat a variety of medical problems. Chronic high doses of steroids cause a variety of problems but especially liver toxicity. The liver synthesizes enzymes that degrade many small compounds such as hormones, drugs, and toxins, as described in
CH3
OH
CH3
O
Testosterone
Figure 8.3. Testosterone is a natural androgenic-anabolic steroid.
234
ReAction! Chemistry in the Movies
chapter 5. The enzymes that degrade steroids are particularly active, keeping the hormone levels low. When the steroid degradation products build up too quickly, the result can be jaundice, liver failure, or even liver tumors. Controlled studies as early as the 1930s revealed that chronic low use of testosterone caused many unwanted side effects, such as atrophy and shrinking of the testicles, menstrual cycle changes or cessation, male impotence, male hair loss, and facial hair growth for women. In both sexes, anabolic steroids cause a tendency toward acne, trembling, bad breath, and high blood pressure. Longtime users suffer uncontrollable rage as a side effect of high blood pressure and stress. Testosterone was fi rst synthesized in 1935 and was introduced into the sporting arena in the late 1940s. Their earliest high-profi le use in the sports world was when the Russian weightlifting team won the top honors in the 1952 Olympics. At the 1954 World Weight Champion competition, the Russian team physician freely admitted to the U.S. team physician Dr. John Ziegler that their team was using synthetic testosterone (Yesalis 2000). This spurred Dr. Ziegler to escalate the steroid wars by developing the synthetic androgenic-anabolic steroid sold as Dianabol (Yesalis 2000). Still among the most potent, its use immediately skyrocketed among collegiate weightlifters, football players, and wrestlers. Even though the International Olympic Committee (IOC) voted to disallow doping in the late 1960s, the use of steroids to build muscles in the athletic world had spread to most other sports and was fi rmly entrenched. Since athletes continued to use steroids covertly, the IOC instituted steroid testing for the fi rst time during the 1976 Summer Olympics in Montreal. The problem has not abated, and a 1993 survey by the American Medical Association found that more than one million U.S. citizens use or have used androgenic-anabolic steroids (Yesalis et al. 1993). They found that the average age to start using them was 18 years. The cat-and-mouse game of developing new stimulating substances and new diagnostic tests led to the creation of the World Anti-Doping Agency in 1999, an organization with members from the IOC and other amateur athletic organizations (World AntiDoping Agency 2002).
CLASSROOM CHEMISTRY MOVIES The Prince and Me (2004) Distribution company: Paramount Pictures Director: Martha Coolidge Screenwriters: Mark Amin, Jack Amiel, Katherine Fugate, and Michael Begler Short summary: University of Wisconsin premedical student Paige Morgan meets undercover Danish Prince Edvard and falls in love MPAA rating: PG
Chem 101
235
Plot description: Paige Morgan (Julia Stiles) attends the University of Wisconsin, where she studies hard so she will qualify for admission into a medical school program. She would like to join “Doctors Without Borders,” an international benevolent association in which physicians provide free care to people without ready access to medical services. Her parents run Morgan Dairy in Manitowoc, Wisconsin. Her life is contrasted with Edvard “Eddy,” the playboy Prince of Denmark (Luke Mably). He drives hot cars too fast through the narrow streets of Copenhagen while being pursued by paparazzi. They get what they want when he kisses two beautiful women. He decides he needs to get out of Denmark and chooses Wisconsin after seeing an advertisement on television for a Wild College Girls video, supposedly made in Wisconsin. There are several short scenes beginning at 22:00 of Morgan’s organic chemistry laboratory. In the fi rst short scene, her organic chemistry professor announces, “Science is an open quest.” On the blackboard, he writes that their grade will be based on 40% procedure, 50% lab work, and 10% attendance. Eddy arrives late, but the professor allows him to enter anyway, and he becomes Morgan’s lab partner. She lets Eddy know this course is important to her and he had better carry his weight. When Eddy doesn’t show up for the next lab period, she works alone to add hydrogen chloride slowly to a solution that foams up and squirts out of its apparatus. Commentary: Morgan and Eddy are polar opposites and yet strangely attracted to one another. When Morgan does poorly on her English composition assignment, Eddy reveals his deep understanding of Shakespeare’s love sonnets. In return for his help, she teaches him how to use a washing machine and so on, but he still earns a D-minus in chemistry. Earlier, Morgan’s mother told her, “There’s more to chemistry than what you learn in the classroom.” Harry Potter and the Chamber of Secrets (2002) Distribution company: Warner Brothers Director: Chris Columbus Screenwriter: Steven Kloves, based on the 1998 novel by J. K. Rowling Short summary: Someone has petrified muggle-born students and the cat Mrs. Norris; mandrake root juice restores them MPAA rating: PG Plot description: The second year for Harry Potter (Daniel Radcliffe) and his classmates at Hogwarts School for Wizards begins smoothly enough. The students are in the greenhouse, where they are required to wear protective earmuffs before repotting their mandrakes. In the 1.75-minute clip at 31:30, studious Hermione Granger (Emma Watson) explains that mandrake root, or mandragora, is an antidote for petrification. Then, Professor Sprout demonstrates how to lift the mandrake out of its pot and into a new one. Its cries are deafening, but the students follow her directions. The message is
236
ReAction! Chemistry in the Movies
that if you are going to use herbal remedies, you need to learn how to maintain a garden. The scene gains weight when it is next revealed that someone or something at the school is petrifying muggle-born students. Commentary: Mandrake tuber actually does exist (botanical name Mandragora officianarum), and it does have a humanoid shape. Flavius Josephus even created a myth in 93 C.E. that its screams will kill anyone who uproots it. He recommended, therefore, that you starve your dog for a few days, tie it to the mandrake’s greenery, and then use some meat to tease it enough to pull up the mandrake for you. In contradiction to Hermione’s description, though, mandrake root has been used since Greek and Roman times as a “sleeping draft” (called an anesthetic today) at low doses and as a deadly poison at high doses. In the late 1800s, Scottish surgeon Benjamin Richardson was searching for general anesthetics (Robinson 1946). Even though ether and chloroform anesthesia had revolutionized dentistry and surgery in the 1840s (see the description of The Great Moment in chapter 9), they had a number of undesirable side effects, especially nausea. In 1874, Richardson published his experiments describing how he prepared an alcoholic extract of the mandrake root, tested its effects on lab animals, and then selfexperimented. It made him nauseous, so he decided it was unsuitable. For many herbal remedies, chemists were able to fi nd a single active ingredient that was responsible for the physiological activity. King’s American Dispensary indicates that French chemist Crouzel isolated the active ingredient from mandrake root and named it mandragorine (mandragora + amine, for the nitrogen) but that later chemists showed it was identical to the molecule named atropine isolated from the deadly belladonna plant (Atropa belladonna) (Felter and Lloyd 1898). The structure of atropine and cocaine were described in chapter 2, and atropine’s use as an antidote for certain types of nerve gases (see, e.g., The Rock, 1998) is noted in chapter 3. Tortilla Soup (2001) Distribution company: Samuel Goldwyn Company Director: María Ripoll Screenwriters: Tom Musca, Ramón Menéndez, and Vera Blasi, based on Ang Lee’s 1994 script for Eat Drink Man Woman, a Taiwanese fi lm directed by Ang Lee Short summary: Martin Naranjo is retired master chef who prepares magnificent weekend meals for his three grown daughters, one of whom is chemistry teacher Leticia MPAA rating: PG-13 Plot description: In the long opening sequence, a man purchases delectable groceries at the Big Saver, gathers produce from his lovely garden, and then prepares elaborate foods such as squash blossom soup. Martin Naranjo (Hector Elizondo) is a retired master chef who prepares
Chem 101
237
meals for his three grown daughters every weekend at his home in Los Angeles. Interspersed with the preparation scenes, his oldest daughter Leticia (Elizabeth Peña) explains to her friend that she’s Christian and not Catholic. Her father doesn’t understand. When the three daughters converge to eat the meal, they fi nd it tasteless, and the youngest one tells him so. This sets off a family argument during which the middle sister announces she’s going to take a job in Barcelona. In the scene at 28:00, Leticia is about to begin discussing the fi rst equation on her blackboard (figure 8.4) when a baseball crashes through the classroom window. She tosses it back to the new coach, Orlando Castillo (Paul Rodriguez), as a student writes, “Love is a chemical reaction” on the blackboard. The next day, in the scene at 35:00, Leticia fi nds a handmade card on her desk and assumes it is from Castillo. An image of a chain reaction (one sphere hitting another sphere and four spheres scattered from it) is on the blackboard along with the words “Fluorescent,” “gamma,” “infrared,” “laser,” and “photon.” On day 3, another love letter arrives and the blackboard says “Acids and Bases.” On day 4, the card is cradled in a molecular model on her desk. It says, “I love you and hope you feel the same.” At last, her resistance falls and she walks across campus to kiss him while he is coaching a game. He is surprised and pleased but didn’t write the letters. When she fi nds out that her students wrote the letters, she is embarrassed until he explains he would have sent them if he could write like that. They kiss privately in her classroom. Commentary: Love is the equation that balances the chemistry teacher’s life. On the fi rst day, she cannot balance the equation because the baseball interrupts her lecture. On the next day, her love explodes in a chain reaction. On the third day, the opposites of acids and bases neutralize and balance each other. On the fourth day, a love letter is cleverly cradled in a molecular model on her desk, both of which were placed there as offerings to her heart.
1. Zn + __ HCl
ZnCl2 + H2
2. N2 + __ H2 3. __ H2O
2 NH3 2 H2 + O 2
4. __ NaCl + H2SO4
Na2SO4 + 2 HCl
Figure 8.4. Leticia Naranjo was about to use these four equations on her blackboard to teach her students how to balance chemical equations when the baseball flew into her classroom.
238
ReAction! Chemistry in the Movies
October Sky (1999) Distribution company: Universal Pictures Director: Joe Johnston Screenwriter: Lewis Colick, from the 1998 autobiography Rocket Boys: A Memoir by Homer Hickman Short summary: Homer Hickman develops rocket fuels after learning about Sputnik; based on a true story MPAA rating: PG Plot description: On October 4, 1957, a radio broadcaster announces the Soviets have launched the satellite Sputnik, and everyone rushes outside to watch it streak across the sky. The Hickman family lives near Olga Coal Company in Coalwood, West Virginia. High school student Homer Hickman (Jake Gyllenhaal) is inspired to collect explosives and use them to power a small rocket. When it explodes rather than lifts off, Homer asks the class nerd Quentin Wilson (Chris Owen) for advice. Four friends build a rocket in Hickman’s basement using a mixture that includes sulfur for fuel. It works but hurtles into the mine buildings. John Hickman (Chris Cooper) is a taciturn but respected coal mine supervisor who doesn’t believe there is much need for education when you work the coal mines. He wants his son Homer to follow in his footsteps. Homer’s high school teacher Miss Frieda Riley (Laura Dern) encourages Homer to follow his dreams, so the boys build a launch site eight miles away. The brief scene in the high school chemistry laboratory begins at 28:30 with a shot of the blackboard (figure 8.5), even though we never see the students engaged in their organic qualitative analysis exercise. Instead,
QUALITATIVE ANALYSIS 1) Density 2) Boiling (g/cm3) Point (°C)
acetanilide ethyl acetate n-hexane ? i-propanol cyclohexane
1.22 0.90 0.66 0.79 0.80 0.79
304 77 69 65 97 81
SOLUBILITY WATER
S S S ? ? ?
ETHANOL 1.0 M HCl
S S S ? ? ?
S S S ? ? S
Figure 8.5. The laboratory blackboard lists the physical properties for a collection of organic liquids. The information can be used to design a protocol capable of distinguishing the six liquids. Information hidden from the camera is indicated with a question mark.
Chem 101
239
Wilson describes to Homer how potassium chlorate plus sugar would be a great fuel for their rocket. He mixes it together but pours it down the drain when the teacher walks past. When another student tosses a burnt match down a different drain, all the drains simultaneously ignite. After many launch failures and with the help of many people, they fi nally succeed with the aid of an engineer who helps them control the gas emissions. Miss Riley encourages the boys to enter their project in a science fair. Upon winning, their fathers become angry at them and Miss Riley. When it is determined that only one boy can go to Indianapolis for the National Science Fair, Homer is chosen. On the night of his arrival, he pays to see The Incredible Shrinking Man while someone sabotages his display. With help from everyone back home, including his father, the display is rebuilt and they win fi rst prize. At the end of the movie, we learn that all four of the rocket boys attended college, Wilson became a chemical engineer, and Homer became a NASA engineer. Commentary: The movie provides much less information about the various rocket fuels than the book. In the book, they started with ordinary black gunpowder (charcoal, potassium nitrate, and sulfur), moved to the “rocket candy” (sugar and potassium nitrate), and tried other things including “zincoshine” (“moonshine” ethanol, sulfur, and zinc). Even though it is a great special effect, when dry potassium chlorate and sugar are mixed and poured down a drain, they will not combust explosively as shown in the movie. Instead, the chlorate has to be heated until it is melted, at which point it decomposes to form molten potassium chloride and molecular oxygen. According to Homer Hickman’s website, October Sky is an anagram of Rocket Boys. The Saint (1997) Distribution company: Paramount Pictures Director: Philip Noyce Screenwriters: Jonathan Hensleigh and Wesley Strick, based on the fictional master criminal created by Leslie Charteris in 1928 Short summary: Simon Templar is hired to steal a cold fusion formula from Oxford University electrochemist Dr. Emma Russell and then falls in love with her MPAA rating: PG-13 Plot description: Simon Templar is an orphan with a false name who has grown up to be a master criminal. This adventure begins when he accepts a contract from Ivan Tretiak (Rade Serbedzija) for $3 million to steal the cold fusion formula from Dr. Emma Russell (Elisabeth Shue). Tretiak is a ruthless Russian billionaire who wants to use the limitless cold fusion energy to overcome the harsh Russian winter. He believes this will make him a hero in the eyes of the Russian people, who will then elect him as the next leader of the Russian Federation.
240
ReAction! Chemistry in the Movies
In the 3.75-minute clip beginning at 29:00, electrochemist Dr. Emma Russell holds a press conference at Oxford University to announce she knows how to achieve cold fusion. She is about to show a chalkboard full of theory when the scene changes. (The chalkboard contains an iodateiodine reduction-oxidation reaction, which has no relationship to cold fusion.) Templar breaks into her apartment and downloads the notes she has stored on her computer. At 33:00, we are shown a series of three pages from Russell’s Macintosh Note Pad: “Page 1” is a schematic diagram of the fusion reaction (deuterons enter the palladium matrix and then fuse); “Page 2” are the electrochemical reactions shown in the previous schematic; and “Page 3” is a summary of the three possible cold fusion reactions (to make He-3 and a neutron, to make H-3 and a proton, or to make He-4 and lots of energy). From objects in her apartment, Templar deduces she has a poetic heart, so he dresses as a poet himself and seduces her to fi nd out she has other notes hidden in her black lace brassiere. His mission accomplished, he sends the notes to Tretiak but then realizes he’s in love. Before the movie ends, Templar and Russell must prevent Tretiak’s electrochemist from using the notes and, consequently, prevent the downfall of the current Russian president. While running for their lives from Tretiak’s hit men, Russell decides to give the fi nal missing part of her formula (figure 8.6) to the Russian scientist for free so the world
Γdd =
σT =
KOd ON σT 4 πα h2
∫ λdE = ∫ ( 2π ∞
MH
dN ) dE d
M H = ∑ ∫∫∫ ψ1n θ ERψ n d 3 r 2
o
n=
θ ER = 4 πζδ (r − r') + ϕ nd ϕ dd ∴Γdd = 1019 s−1 at STP
⇒ 1.39 Megawatt!! Figure 8.6. Dr. Russell’s final note supposedly “explains” how cold fusion works. The circles and arrows have been added to show how each line relates to the next. The first equation describes the rate of deuteron–deuteron fusion *dd, which is only known from the energy absorption cross section VT, which is only known from the strength of coupling between the initial and final states |MH|2, which is only known from the density of deuterons in the lattice as measured by the Hamiltonian operator TER. When you know the density of deuterons in the lattice, you can solve the Hamiltonian and then all of the other equations.
Chem 101
241
can have cheap energy. Her note indicates that 1.39 megawatts of energy can be generated from a deuteron–deuteron interaction rate of 1019 per second at standard temperature and pressure (STP). Commentary: Cold fusion is nuclear fusion at ambient temperature. On March 23, 1989, the University of Utah electrochemists Stanley Pons and Martin Fleischmann held a press conference that shook the energy world. They announced reproducible cold fusion with 10% more energy released than they supplied. The conditions were amazingly simple: pass an electric current through palladium or platinum wires in a container of heavy water and lithium sulfate. In heavy water, the two hydrogen atoms are replaced with deuterium, called heavy hydrogen because it has one proton and one neutron. During “heavy water” cold fusion, two deuterium atoms are brought together to fuse and form a helium atom according to the reaction 2 D2O o 2 He + O2 + energy. The barrier to nuclear fusion is coulombic repulsion. The positively charged protons in each nucleus repulse as the atoms approach each other. In hot fusion, the repulsion is overcome by using either 50 million degree temperatures or particle accelerators. In cold fusion, the negative charge of the metal catalyst is supposed to overcome it. There are three possible deuterium fusion outcomes: (1) D + D o T + H; (2) D + D o 3 He + neutron; or (3) D + D o 4He + J (24 MeV). The last outcome is the most desirable. Many labs have extensively tested the claims made by Pons and Fleischmann. They have found that no tritium, neutrons, or gamma rays are generated and that the He concentration does not change but that there is a source of unexplained minor excess energy production. Whatever the source of that energy, it is enough to keep researchers working in this area despite the heavy negative press. Dr. Eugene Mallove, a strong proponent of cold fusion, was the “science consultant” for this movie. He undoubtedly supplied the formulas on the computer, Post-it Notes, and the fi nal missing part of the solution (figure 8.6). The Nutty Professor (1996) Distribution company: Universal Pictures Director: Tom Shadyac Screenwriters: David Sheffield, Barry W. Blaustein, Tom Shadyac, and Steve Oekekerk, based on the 1963 Jerry Lewis movie Short summary: Overweight genetics professor Sherman Klump creates a quinone compound that transforms him into thin, aggressive Buddy Love; Klump takes it because he’s in love with chemistry professor Carla Purty MPAA rating: PG-13 Plot description: Overweight genetics professor Sherman Klump (Eddie Murphy in one of seven roles) studies obesity and has successfully
242
ReAction! Chemistry in the Movies
developed a genetically based weight loss formula that works on hamsters. His colleagues all know about his work, and the college dean (Larry Miller) even helps him get funding by arranging for him to meet with the school’s wealthy alum Harlan Hartley (James Coburn). Then he falls in love with the new chemistry professor, Carla Purty (Jada Pinkett Smith), who says she’s read all his papers. After she agrees to a date, he attempts to lose weight, but he’s just as portly as ever when they take their seats at “The Scream” comedy club. He is mortified when the comic picks him out for insults about his weight. He returns home depressed, eats an entire carton of ice cream, and falls asleep only to dream that he is the size of Godzilla. When Klump wakes, he resolves to test his formula on himself. In his laboratory, an overweight hamster watches as Klump connects himself to the monitors and then has the computer calculate the dose based on his weight. The numbers rise rapidly as the camera pulls away to his surprised face. At 33:00, a molecule rotates on the computer screen that must be the quinone compound Klump’s been working on (figure 8.7). He adds a drop of something to a test tube solution and it glows blue. He drinks it and almost instantaneously becomes the slim, aggressive Buddy Love. Every time Love appears, his testosterone is higher until it reaches dangerous levels, allowing Klump to fight back. Commentary: Rick Baker and David Roy Anderson won the Academy Award for Best Makeup for their efforts on this fi lm. Eddie Murphy gained the appearance of 250 pounds through the use of body and facial prostheses composed of molded foam rubber. The proposed compound that reduces weight genetically resembles dATP, one of building blocks for DNA (figure 8.7). The chemical love story gets even better when considering that the proposed compound just might work in the real reaction that caused the solution to glow blue. One of the more sensitive methods for determining ATP concentration
NH2
N
O N O
O
N
O N
-O
P O-
O
P O-
O
P
O
OO
Adenine Quinone Triphosphate (AQTP) Figure 8.7. The molecule on the computer monitor at 33:00 has a structure in which the deoxyribose of deoxyadenosine triphosphate (dATP) is replaced by a quinone. It would reasonably be called adenine quinone triphosphate (AQTP).
Chem 101
243
involves the enzyme from fi refl ies that allows them to glow when they are trying to attract a mate (figure 8.8). For every molecule of ATP that is consumed by the enzyme luciferase, one molecule of luciferin is converted to oxyluciferin and one photon of light is emitted. The Affairs of Dobie Gillis (1953) Distribution company: Metro-Goldwyn-Mayer Director: Don Weis Screenwriter: Max Shulman Short summary: Pansy Hammer and Dobie Gillis attend Grainbelt University; she loves to mix chemicals together until they cause an explosion Plot description: Pansy Hammer (Debbie Reynolds) attends Grainbelt University to “work, work, work,” so she signs up for English and chemistry. Dobie Gillis (Bobby Van) is only there to have fun. He sees Hammer from a distance, and tells the registrar that he’ll take whatever courses she’s taking. On the fi rst day of class, the decidedly uncheery chemistry HO
N
S
S
N
O
Luciferin
O-
ATP
Firefly Luciferase PPi HO
N
S
S
O2
HO
N
O
AMP-Luciferin
N
O
AMP
S
S
N
O O
oxyluciferin
O
O
AMP
H HO
N S
Oxyluciferin
S N
+CO2 +AMP O
+light
Figure 8.8. The Firefly Luciferase reaction that was undoubtedly used in the movie to create the glowing blue solution that Professor Klump drank. It is possible that AQTP from figure 8.7 would substitute for ATP in the reaction.
244
ReAction! Chemistry in the Movies
professor tells the students, “This is the hardest class you’ll ever take.” Hammer is happy anyway and tells her lab mates how much she likes chemistry just before her experiment explodes. After a montage of school days passing and two more explosions, Hammer succumbs to Gillis’s charms, and they skip class for a couple of weeks. When the chemistry lab is “due tomorrow morning,” they decide to sneak into the laboratory to do it that night. At the beginning of the 4.0-minute scene from 41:30, we get a good look at the blackboard (figure 8.1) as the professor turns out the lights and leaves the room. Hammer and Gillis emerge from their hiding place in the closet and begin the inorganic qualitative analysis exercise. Hours later, they have determined that her unknown is silver chloride. They’re both tired so he decides to nap while Pansy cleans up. She pours all the reagents into a large round flask until it bubbles over and explodes. Commentary: The blackboard from this movie was described in this chapter’s fi rst section. It Happens Every Spring (1949) Distribution company: Twentieth Century Fox Director: Lloyd Bacon Screenwriter: Valentine Davie Short summary: Chemistry Professor Vernon Simpson discovers a nitrocyclohexane compound that causes baseballs to avoid wooden surfaces, including baseball bats Plot description: Vernon Simpson (Ray Milland) is the new chemistry professor at the university. His girlfriend, Deborah (Jean Peters), is the president’s daughter and is taking his class. In the 2.5-minute scene from 10:30, he is lecturing about carboxylic acids, their acidic strengths, and their dissociation constants. Behind him, the blackboard shows the acid dissociation constants for a variety of compounds (figure 8.9), plus the
Acetic Acid Chloroacetic Dichloroacetic Trichloroacetic Benzoic Oxalic
KA 1.8 x 10-5 1.54 x 10-3 5 x 10-2 2 x 10-1 6.7 x 10-5 3.8 x 10-2
CH3—CO2H CH2Cl—CO2H CHCl2—CO2H CCl3—CO2H —CO2H HO2C—CO2H
Figure 8.9. The acid dissociation constants for a variety of carboxylic acids on the blackboard of Simpson’s chemistry classroom.
Chem 101
245
Lewis dot structure for acetic acid (not shown here). During his lecture, he reaches down to get a bottle from a low shelf behind the counter and takes the opportunity to turn on the radio. He continues his lecture while listening with one ear to the baseball game. For his fi rst demonstration, he places some acetic acid in a test tube and adds a few drops of methyl orange to it. The bright pink color proves it is an acid. Class ends before he is able to discuss the information on the blackboard. Deborah meets with Professor Simpson after class. They can’t get married because he doesn’t have enough money, but he tells her he is almost certain his biophobic is going to be a success. A biophobic repels insects from wood, and his work is so promising that Norworth Laboratories is interested in his nitrocyclohexane compound. She is eager to see the experiment, so they step into his laboratory next door. His apparatus is bubbling away near the window, and they are overjoyed to see the white precipitate come out at the end of the distillation apparatus. As they step aside so he can write a note in his book, a baseball crashes through the window, destroying the apparatus. In the 3.5-minute scene from 13:00, he says it has taken him months of preparation to get to this point. After his fiancé leaves, he picks the baseball out of the milky solution and tosses it aside, only to see it curve away from the wooden shelf. As a test, he wipes off the ball’s surface and now it rolls into the wood. Next, he rolls the ball in the solution, and it avoids wood once again. Excitedly, he writes his fi ndings in his notebook, and his voice-over narration explains the biophobic’s properties. He suspends the baseball with some string and then swings at it with a piece of wood. The ball jumps out of the way. He pours the solution into a container and makes his way to St. Louis to try out for the baseball team. Against all odds, he becomes the pitcher and takes the team all the way to the World Series after a season of strikeouts. Commentary: Simpson’s deferred lecture concerned polar effects in molecules. Gilbert Newton Lewis was the fi rst to offer a lucid explanation for the ability of one functional group to alter the properties of other functional groups (Lewis 1998). He was head of the chemistry department at the University of California–Berkeley and had fi rst described his fruitful thoughts in a 1916 journal article. His role as chemical warfare adviser to General Pershing during WWI interrupted his work in this area, and it wasn’t until 1923 that he was able to follow up. In that year, he published Valence and the Structure of Atoms and Molecules, in which he described how the electrons in the bonds between atoms in a molecule influence the overall properties of the molecule (Lewis 1923). For his example of “polar effect through chain,” Lewis compared acetic acid with chloroacetic acid (figure 8.9). The methods to determine the absolute strengths of acids and bases were developed in the 1890s and 1900s. By 1916, the strengths of only two dozen or so organic acids were known, and each was different. Lewis explained how their structures influenced their acidity, and his explanation is still used in chemistry
246
ReAction! Chemistry in the Movies
classrooms today. He writes that the electrons “drawn toward the chlorine, permit the pair of electrons joining the methyl and carboxyl groups to approach nearer the methyl carbon . . . producing a greater separation of the electrons from the hydrogen on the hydroxyl, and thus a stronger acid.” On page 139 of his book, he restates this principle and ends it with “the substitution of a second and third chlorine heightens the effect,” but he doesn’t elaborate further. Bond polarity and inductive effects in general remain one of the most reliable aspects of the structure–activity relationships used by pharmaceutical chemists to predict molecular properties. Madame Curie (1943) Distribution company: Metro-Goldwyn-Mayer Director: Mervyn LeRoy Screenwriters: Paul Osborn and Paul Hans Rameau, from Eve Curie’s 1937 same-titled biography of her mother Short summary: Biography of Marie Sklodowska Curie, who discovered radium with the help of her husband Pierre; based on a true story Plot description: Marie Sklodowska (Greer Garson) is working toward two degrees, in math and in physics, at the Sorbonne in Paris. She doesn’t eat to save money and faints in Dr. Perrot’s (Albert Bassermann) class one day. Since she has no friends or family living in Paris, Dr. Perrot arranges for Dr. Pierre Curie (Walter Pidgeon) to accept an assistant. When Curie learns that the student is a woman, he accepts her anyway, but says women are a distraction. Six months later, Marie’s studies are nearly complete and she prepares to return to Warsaw. Curie tells her she can’t return because, “You have a gift for research. Anyone can teach.” After she earns top honors in physics, Curie arranges for her to spend the weekend with him and his parents. At the cottage that night, he barges into her room to talk of science and collaboration. He never mentions marriage but she knows exactly what he’s asking. She agrees to stay in Paris, and Curie leaves to tell his parents “we are engaged.” They marry and have a short honeymoon. Sklodowska Curie decides to work on Bequerel’s fi nding for her Ph.D. She is soon discouraged because things don’t add up. In the 8.5-minute scene from 57:15, they are in the lab and the electrometer is functioning perfectly. She recounts her fi ndings: Uranium and thorium are the only known radioactive elements, whereas crude pitchblende is more radioactive than pure uranium. She demonstrates that the pitchblende electrometer reading always gives eight units, the uranium contributes two units, and the thorium contributes two more units, leaving four missing units. He asks whether she has checked the chemical analysis of pitchblende, and she shows him a blackboard with the pitchblende composition (figure 8.10). After talking through the list, she looks again at the
Chem 101
Uranium Oxide (UO) Thorium Oxide (ThO) Lead Sulphide (PbS) Silicon Dioxide (SiO2) Calcium Oxide (CaO) Barium Oxide (BaO) Iron Oxide (FeO) Magnesium Oxide (MgO) Other extraneous matter
247
75 % 13 % 3% 2% 3% 2% 1% 0.99 % 0.001 %
Figure 8.10. The laboratory blackboard lists the mass percent of each pitchblende component.
last line that reads “Other extraneous matter 0.001%.” She thinks out loud, “Our universe has a fi xed composition, hasn’t it?” He says, “Go on.” “Men originally thought everything was made of four elements. Now, we know of 78. What if there’s a new kind of matter that isn’t fi xed? We’d have to change our conception of matter. Pierre, we have discovered an active element.” They are given permission from a group of Sorbonne administrators to use an unheated shed in the courtyard that leaks in the rain. In the 3.75-minute montage from 1:10:15, the narrator describes the procedure for isolating radium: boil pitchblende, add acids to dissolve the salts, melt down into separate cauldrons, release the gas fumes, fi lter and refi lter. As the months run into years, they fi nally have only barium and another element with similar properties that they believe is radium. The problem is how to fi nish the separation. Sklodowska Curie’s fi ngers hurt, so she visits a doctor, who advises her to abandon her experiments. He thinks the burns on her fi ngers could become cancerous. She decides to continue and even speculates whether the unknown element might have the power to destroy healthy and unhealthy tissue. It might be used to cure cancer. In the second montage, the narrator describes the action again, and this time the key is sequential crystallization. It took two years to carry out 5,677 crystallizations (figure 8.11). They started with eight tons of pitchblende, and the entire process took four years. When they peer into the last bowl, they see only a stain and are upset. They return home, deal with their children, attend a New Year’s party, return home, and go to bed. Marie can’t sleep, and says to Pierre, “We didn’t test the stain, did we? What if it’s very active and you don’t need much?” They rush back to the lab through the snow. In the dark, they see that it glows. News reporters are at their home, and they announce they plan to give their discovery away for free. Commentary: The discovery of radium is one of the most frequently told tales in chemistry textbooks because it brings together so many aspects: the composition of the atom, the nature of isotopes, the nature of
248
ReAction! Chemistry in the Movies
Figure 8.11. Marie Curie (Greer Garson) works on the final separation of radium from the chemically similar barium while husband Pierre (Walter Pidgeon) watches. This process required 5,677 sequential crystallizations. MADAME CURIE © Turner Entertainment Co. A Warner Bros. Entertainment Company. All Rights Reserved. Photo courtesy of the Academy of Motion Picture Arts and Sciences.
radioactivity, the birth of the atomic age, Sklodowska Curie’s two Nobel prizes, and so on. Even so, chemistry students never learn the personal drama behind the discovery unless they watch a movie such as this one. They would never know the Curies spent four years stirring boiling cauldrons fi lled with acid to the detriment of their own health (Curie 1937; Reid 1974). The Curies also refused to patent their procedure, to their economic detriment, and they never had enough funds to carry out their own research. On the other hand, so much of the Curies’ life story was omitted that it makes them appear to be altruistic scientific saints rather than real-life vulnerable human beings (Reed 1989; Elena 1997; Flicker 2003). The movie took five years to make, from the fi rst script treatment in 1938 to its debut in December 1943, longer than the four and a half years it took the Curies to carry out their experiments (Latham 1971). In addition, the $1.4 million it cost to produce the movie “would have bought and sold many times over the leaky shack which the Curies used as their radium lab.” The fi rst two writers assigned to the project were the legendary F. Scott Fitzgerald (Latham 1971) and Aldous Huxley (Dunaway 1989). The problem in bringing Eve Curie’s 1937 biography of her mother
Chem 101
249
to the screen was balancing the love story with the scientific story, and no one seemed to do a satisfactory job. Despite the absence of a satisfactory script, California Institute of Technology physicist Dr. Rudolph Langer was hired in early 1942 as a consultant by the fi lm’s producer. The two of them reenacted the major experiments and discussed the visual melodrama. When Langer objected to spicing it up with sizzling sparks or emotive conversations, he became known as “I-Don’t-Like-ItDr.-Langer” to the writers at MGM. He insisted that “no matter how tense the moment, a scientist would always remain calm.” In December 1942, the fi nal script was ready. Director Mervyn LeRoy completed this unique scientific romance in time for a Christmas release and later wrote that it was one his own favorites among his fi lms (LeRoy and Kleiner 1974). He made sure he understood every scene because, “If I didn’t understand it, I knew that the audience wouldn’t understand it either.” The movie was a critical and popular success. The Chemist (1936) Distribution company: Educational Pictures Director: Al Christie Screenwriter: David Freedman Short summary: Duncemore College chemistry student Elmer Triple invents four things, including a noiseless explosive Plot description: In the Duncemore College “Laboratory,” chemistry student Elmer Triple (Buster Keaton) appears to be hard at work until we realize he has retrieved a potato and egg from the boiling water in a round bottom flask. Next, he uses tongs to toast some bread using a Bunsen burner flame and pours coffee from a glassware setup. When he squirts some milk into it from the fi ngers of a suspended rubber glove, the professor (Earle Gilbert) approaches him to say, “Elmer, I always thought you would be my prize student, but look at this.” Elmer defends himself by saying he’s invented a Businessman’s Breakfast Powder. He mixes the solid food in a graduated beaker, adds the coffee, and pours in the powder until it foams. He drinks it in one gulp saying, “The eggshells are hard to digest.” His next three inventions are Elmer Triple’s Tripling Powder, a love potion, and a noiseless explosive. When safecrackers read about Triple’s success on the front page of the newspaper, they don caps and gowns to sneak onto campus to kidnap Triple. Commentary: The narrative has clear links to inventors with its oxymoronic noiseless explosive. Even so, the action in this movie takes place at a college and is carried out by a chemistry graduate assistant in the laboratory, which places it fi rmly in the category of chemical education. The only thing missing is a blackboard in this otherwise perfectly amusing 20-minute short by the incomparable Buster Keaton.
9 Good News Research and Medicinal Chemists Making a Difference
SOLVING PROBLEMS FACTUALLY AND FICTIONALLY Stories of people doing their jobs well, treating each other with respect, and trying to make the world a better place are all examples of “good news.” Such stories don’t generate many website hits, nor do they bring people into the theaters. Instead, it seems readers and movie viewers would rather have the double pleasure of learning about bad behavior and its comeuppance. Five movies in this chapter overcome this problem; they are based on true stories (table 9.1). The advantage of such stories is the sympathy viewers feel as they appreciate the adversities the chemist has overcome to make their celebrated fi ndings. For instance, in the documentary Me & Isaac Newton, which explores the motivations of seven scientists, pharmaceutical chemist Gertrude Elion is warm and charming as she describes why she decided to become a chemist. When she later describes her struggles to enter graduate school and then get a job as a chemist, the viewer is struck by her matter-of-fact, water-underthe-bridge tone. This all happened before she understood there was a climate of active discrimination against women that had nothing to do with their drive or abilities. Still later, she says the ultimate reward for all her work comes when someone thanks her for having developed the drug that cured a loved one. The disadvantage of using true stories is the need to create dramatic tension. The important moments in people’s lives rarely coincide with obvious indications that “this is the moment when everything fell into place,” whereas a movie’s linear narrative has to make that point clear to the audience. Another problem for moviemakers is that most people just aren’t very curious about the origins of everyday things. This is a challenge because very few chemicals cause the imagination to soar (unless you are a chemist), which may explain why all five movies based on true stories are about medicinal chemistry, which can be seen as the external evidence of the chemist’s desire to do good things for other people. Fictional movie chemists are less likely to develop medicines (table 9.1). Like the chemistry professors in chapter 8, they tend to develop chemical products for more selfish reasons. For instance, Elvis in Clambake 250
Good News
251
Table 9.1. Movies Featuring Chemists Solving Problems Title (Year)
Based on a True Story?
Medicinal or Other
Me & Isaac Newton (1999) Medicine Man (1992) Lorenzo’s Oil (1992) The Serpent and the Rainbow (1988) Clambake (1967) Forever, Darling (1956) Monkey Business (1952) Strange Impersonation (1946) The Great Moment (1944) Dr. Ehrlich’s Magic Bullet (1940) The Love Test (1935) The Schemers (1922)
Yes No Yes Yes No No No No Yes Yes No No
Medicinal Medicinal Medicinal Medicinal Varnish Insecticide Medicinal Medicinal Medicinal Medicinal Fireproof fi lm Gasoline substitute
identifies himself as a (chemical) engineer. He uses his chemical knowledge to gain the love of a girl and his father’s respect by winning a boat race. He is also willing to self-experiment by racing the speedboat before his prototype superhard, super-fast-drying varnish has fi nished drying. In his book about fi lm genres, Wesleyan fi lm studies professor Joseph Reed wrote some things about the “inventor movie” that are relevant to our good-news chemists. He notes that heroes are cold characters, so they need adversaries, even if those adversaries are diseases. They can also be softened with the handicaps of being obstinate, forgetful, or eccentric or by having partners who provide them with the critical insight at just the right moment. Finally, Reed believes Americans tend to be anti-egghead, so the good eggheads are often pitted against the budget-cutting bad eggheads.
ALFRED NOBEL AND THE RESPONSIBILITY OF DISCOVERY The international Nobel Prizes are beacons to the world for what’s good about physics, peace, medicine or physiology, literature, and chemistry. They have been awarded nearly every year since 1901 and were created by Swedish industrial chemist Alfred Nobel, who invented blasting caps, dynamite, and smokeless gunpowder. He also founded one of the fi rst multinational corporations and became the world’s loneliest millionaire, according to one popular biography (Evlanoff and Fluor 1969). He created the three science prizes because he believed the sciences were the driving force of the modern world economy. He created the literature prize because he was an avid reader of novels and even wrote a novel and some poetry himself. And, he created the peace prize because of his lengthy correspondence with Bertha von Suttner. She kept him apprised of the growing European peace movements and was awarded the 1905
252
ReAction! Chemistry in the Movies
Nobel Prize in Peace. In 1969, the Bank of Sweden created the “Prize in Economic Sciences in Memory of Alfred Nobel” that they maintain at the same funding level as the original five Nobel prizes. It is often called the Nobel prize in economics. Alfred Nobel’s father, Immanuel, was an inventor whose fortunes rose and fell many times (Bergengren 1962). In 1833, the year Alfred was born in Stockholm, Immanuel went bankrupt and left the country, fi rst for Finland and then for Russia. During the next eight years, Alfred and his two brothers were raised by their mother, who earned money running a small dairy and vegetable shop. When their father fi nally secured funding from the Russian government to build a sea mine factory in 1842, they joined him in St. Petersburg. Russia didn’t have a strong navy and wanted to secure its coasts by planting the mines in strategic locations. Alfred’s private tutor gave him an excellent education, but he never earned a degree of any kind. At age 16 in 1850, Alfred was sent by Immanuel on a trip to Italy, Germany, and the United States. Then, he spent 1851 in Paris working as an assistant to chemist Dr. Théophile-Jules Pelouze, who had retired from academic life a few years earlier. Alfred sought him out because he had worked with the new “safe explosive” guncotton since 1838. Guncotton research caught fi re in 1846 when Swiss chemist Christian Schönbein described a reliable and safe method of its preparation and storage. It was safe when wet and could be stored indefi nitely that way. In the years since, guncotton has replaced black gunpowder (75% g/g potassium nitrate, 15% sulfur, and 10% charcoal) as the explosive used in ammunition. While Alfred Nobel was working with Pelouze, he met Pelouze’s former graduate student, Ascanio Sobrero, who had become a chemistry professor at the University of Turin. Sobrero synthesized nitroglycerin in 1847 (figure 9.1) (Noyes 1996; Dolan 1997) by substituting glycerol for cotton in the nitration reaction to produce a straw-colored oil. He reported that it was unpredictably explosive, had a sweet but burning taste, and gave him a violent headache afterward. Physicians later showed that nitroglycerin dilated blood vessels in the same way as amyl nitrate
NO2
O2N
4
H
O
O
O
C
C
C
H
H
H
NO2 H
12CO2 + 10H2O + 6N2 + O2
Nitroglycerin Figure 9.1. Nitroglycerin combusts in the absence of atmospheric oxygen to release four gases.
Good News
253
(Murrell 1879), which had been synthesized by a similar method and was reported to yield a similar taste and headache. These effects occur because these compounds expand the blood vessels to increase blood flow and reduce blood pressure. Toward the end of his life, Alfred took nitroglycerin to treat his angina, and it is still used for this purpose today. Sobrero never exploited nitroglycerin and eventually stopped working with it altogether because it became more explosive the longer it was stored. It slowly degrades to more explosive compounds (see figure 8.2 for The Nutty Professor’s nitroglycerin for one possibility). On the other hand, when nitroglycerin is fresh, it is so stable that it is hard to detonate. Alfred Nobel returned to St. Petersburg, where his father’s business was soon booming because of the Crimean War. Three years after the Crimean War ended, Immanuel went bankrupt again and returned to Sweden with his wife and youngest son, Emil. The three older sons remained in St. Petersburg, but Alfred soon joined his parents in Sweden. In Sweden, Immanuel Nobel built a factory to begin producing “Swedish Blasting Oil,” a mixture of nitroglycerin and black gunpowder. The blasting oil was used for mining, construction, and demolition. In 1864, young Emil Nobel was killed by one of a string of deadly explosions at the factory. Alfred took control of the company after this and soon invented “Nobel’s Patent Detonator.” Also known as the blasting cap, it allowed the user to set off the nitroglycerin from a distance. It consisted of mercury fulminate, which explodes in response to electrical shock. In 1866, Alfred mixed nitroglycerin with a wide variety of materials to reduce its unpredictability. He settled on a mixture of one part nitroglycerin with four parts kieselguhr, a porous, inert diatomaceous earth. It formed a moldable, clay-like substance that was much safer to handle. The orders for dynamite, or “Nobel Blasting Powder,” as he called it, soared, and he built his fi rst factory outside of Sweden. The German factory exported dynamite to the United States, Australia, and other parts of Europe, each of which was exploiting its buried natural resources and building railways. After more than two decades traveling the world to open manufacturing plants, Alfred Nobel settled in Paris at age 43 (Gray 1976; Abrams 1988). In fact, Victor Hugo called him “Europe’s richest vagabond.” He soon placed an ad in the newspaper that read, in part, “wealthy, highlyeducated elderly gentleman seeks a lady of mature age, versed in language, as secretary and supervisor of the household.” The most qualified applicant was Austrian Baroness Bertha Kinsky, but she worked for him only a week before she eloped with Count Arthur von Suttner. Nevertheless, the relationship was forged, and they maintained a lifelong correspondence. She and her younger husband read widely, and Bertha became convinced that an idealistic European society could evolve along Darwinian principles. In 1884, she published her influential novel Die Waffen nieder! [Lay Down Your Arms!] that was translated into many languages. It was read by nearly as many people as Uncle Tom’s Cabin. In her novel, a woman
254
ReAction! Chemistry in the Movies
marries an officer who then dies in a war. She remarries to a second officer and, even though he shares her hatred of war, he dutifully goes to war twice. After he resigns his commission, they live in Paris until the day he is shot as a suspected spy. The novel propelled von Suttner to leadership in the peace movement. Most significantly, she founded the Austrian Peace Organization in 1891 and began publishing a newsletter with the same title as her world-famous novel. She believed people would evolve in a better direction if they channeled their energies and thoughts into causes much loftier than wars. Alfred maintained an active interest and even donated funds to her various movements. Upon his death in 1896, Nobel bequeathed $9 million to create a fund; the interest was to be used to award annual prizes to individuals but soon expanded to include organizations. The fi rst Nobel Prizes were awarded in 1901, and Bertha von Suttner earned the Nobel Peace Prize in 1905.
C 4 H 8 N 8 O 8: BOOM! Thermal explosives reveal themselves by their high numbers of oxygens and nitrogens relative to their carbons and hydrogens. The oxygen reacts with the carbons and hydrogens in the compound to yield gaseous carbon monoxide, carbon dioxide, and water plus lots of energy. Nitrogen is the key to many of these explosives because it stabilizes a high-energy form of the oxygen but reacts to form stable diatomic nitrogen and energy during the reaction. There are two mechanisms by which chemical reactions lead to an explosion (Atkins 1978). In a thermal explosion, the volume is held constant while the reaction temperature increases exponentially. For a reaction in which the rate increases with temperature, the exponentially rising temperature leads to such high pressures in the confi ned volume that the vessel bursts catastrophically. The other mechanism to generate an explosion is called a chain reaction explosion. The rate of the reaction increases to supersonic speeds when the reaction produces radical intermediates exponentially. Some organic polymerization reactions do this when too much of one of the reagents is added. A thermal explosive reacts to generate gases rapidly and exothermically (Atkins 1978; Patnaik 1992). When such reactions occur at rates faster than the speed of sound, it produces a shock wave capable of shattering its container. For instance, when nitroglycerin liquid is shock-activated, it generates a mixture of four gases within microseconds (figure 9.1). Then, the heat raises the temperature to greater than 2,000°C. As the Ideal Gas Law tells us, this will cause the container’s pressure to rise if the volume of the container is fi xed. The only release for the high pressure is to burst the container. The chemistry of explosivity wasn’t understood until relatively recently (figure 9.2, table 9.2), which means that the earliest explosives
Good News
255 H 3C
CH3 O2N
NO2
N
NO2
O 2N
NO2
NO2 O 2N
N
N
NO2 O2N
N N N
N NO2
TNT
NO2
Tetryl
NO2
N
RDX
O 2N
HMX
Figure 9.2. Four powerful explosives.
Table 9.2. The world’s most commonly produced chemical explosives Compound
First Synthesized
Nitroglycerin TNT Tetryl RDX HMX
1847 1863 yellow dye 1910s 1890s heart stimulant 1930
a
Explosivity Discovered
Explosivitya
1847 1890s 1910s 1920 1930
1.5 (1.0) 1.3 1.6 1.8
Explosivity is always reported as relative to trinitrotoluene (TNT).
were discovered accidentally. For instance, when trinitrotoluene (TNT) was originally prepared, its bright yellow color led to its use as a dye. In fact, many dyes have nitro groups and are not explosive. Thirty years passed before TNT was discovered to have explosive properties. When RDX was synthesized, it was originally used as a heart stimulant. Its explosivity was discovered 30 years later.
OPTIONS AND FUTURES For the past four or five decades, the chemicals manufactured in greatest abundance have been dominated by the raw materials to make fertilizers and plastics (Anonymous 2007). Plastics, antibiotics, and ammonia fertilizer were developed in the early twentieth century, and their use has continued to grow ever since. In fact, their production will continue to grow. As more nations develop, their citizens will also want longer life spans, fewer famines, farming efficiency, population expansion, and consumer choice. For these chemicals, the near future seems certain. Almost everything else about chemistry is changing fast. During the past five decades, chemical research has diffused away from its classical focus on the chemical structures resulting from strong bonds, toward interdisciplinary projects in which the weaker bonds within and between molecules is of greatest interest. Journalist Stephen Ritter described these changes as follows (Ritter 2004): “Chemistry has evolved from largely being a study of the elements, to being the study of
256
ReAction! Chemistry in the Movies
molecules, to currently being the study of molecular interactions.” This trend is likely to continue for all scientific research since funding agencies favor larger, more complex projects with high social impact. For instance, the National Science Foundation’s Five-Year Strategic Plan released in 2006 (National Science Foundation 2006) indicates that their goal is to sustain the nation’s competitive edge in response to a continuously changing landscape. In the summary of their introduction, they wrote, “Today’s research requires globally-engaged investigators working collaboratively across agencies and international organizations to apply the results of basic research to long-standing global challenges such as epidemics, natural disasters, and the search for alternative energy sources.” This type of research will solve problems and sometimes even change society at the same time. Harvard chemistry professor George Whitesides recently devised one clever way to identify research areas with the greatest societal impact (Whitesides 2004). He generated a personal list of nine assumptions that society takes for granted but that he believes are vulnerable to disproof by scientific inquiry. He didn’t limit the assumptions to those that could be overturned by the chemical sciences, but he did note that chemistry is the science of the real world and will always be involved in some way. We summarize his fi rst assumption here and leave it to our readers to learn about the other eight. Whitesides’s fi rst assumption was: “We are mortal.” He noted that infections caused the largest numbers of deaths a century ago but today cause only a small fraction. Today, we die of old age in the form of cardiovascular disease, cancer, Alzheimer’s disease, diabetes, or other degenerative diseases. Of these, the nature and treatment for cardiovascular disease is reasonably well understood. Its rate of occurrence is already decreasing due to the ready availability drugs that control blood pressure, cholesterol production, and blood clotting. Changes in diet and lifestyle can reduce the incidence even further. This leaves cancers, which are problematic because the cancerous cells have been irretrievably transformed and damaged by the time they are detected. Nevertheless, common types of damage have been identified, and new treatments are being devised. On the other hand, cancer prevention will retain its dominance over treatment since the most common environmental triggers are cigarette smoking and exposure to the sun’s ultraviolet rays. Aging and the other diseases are still poorly understood but could result in very significant societal changes when they are overcome. Even though immortality would obviously transform our world (and has been the subject of much fiction), so would a life span greater than 200 years, especially if it were coupled with 100 years of fertility. This theoretical future seems to be within reach but raises its own issues. What if only the very, very rich could afford the costs? What if the life spans in developed nations were 10 times longer than those in less developed nations?
Good News
257
The movies based on true stories in this chapter demonstrate a few ways in which science has transformed society. Dr. Ehrlich’s Magic Bullet (1940) presents the birth of modern drug discovery, which has given us safe and effective treatments for many diseases. The Great Moment (1944) presents the fi rst public demonstration of general anesthesia, which has given surgeons more time to do their work. These discoveries have been so transformative that a world without them seems almost unimaginable.
THE ARCHETYPE MOVIE: DR. EHRLICH’S MAGIC BULLET (1940) Production company: Warner Brothers Director: William Dieterle Screenwriters: John Huston, Heinz Herald, and Norman Burnstine, from a story idea by Burnstine Short summary: Biography of Paul Ehrlich, who discovered a chemical cure for syphilis and gave birth to modern drug discovery Plot description: Dr. Paul Ehrlich (Edward G. Robinson) begins his career as a dermatologist in the “Hospital,” where it disturbs him to prescribe a syphilis treatment he knows doesn’t work. His superior tells him that he must prescribe it because it is the official treatment. To console himself, he works late to fi nd which “new aniline” dyes stain which cells and tissues. While attending a seminar on tuberculosis given by Dr. Robert Koch (Albert Bassermann), he learns that its diagnosis is hindered by the difficulty in visualizing the small bacterial cells on slides. Ehrlich suggests it should be possible to fi nd a stain for them. Koch is skeptical of the young man but urges him to go ahead. Ehrlich loses his job at the hospital and spends months testing dyes until one day he fi nds one that works— his wife Hedwig (Ruth Gordon) fi red up the stove in his laboratory and it heated the slide; the heat allows the stain to enter the cell membrane. He presents his fi nding to Koch, who is under fi re by the budget committee to produce results. Koch is so happy with Ehrlich that he offers him a position in his Institute for Infectious Diseases. Ehrlich gratefully accepts but, unfortunately, must take a leave of absence because he contracted tuberculosis. During his six-month recuperation in sunny and dry Egypt, Ehrlich is called to treat a father and son who have been bitten by an adder. The father shows no symptoms, but the son is gravely ill. He fi nds out the father was bitten four times in his life, and each time his symptoms were weaker. When Ehrlich returns to Germany, Koch shows him a lab full of every dye imaginable, but Ehrlich wants to study how animals respond when injected with snake venom. (He later won the 1908 Nobel Prize in Medicine for his work in discovering the immune response.) One day, his
258
ReAction! Chemistry in the Movies
colleague Emil Behring (Otto Kruger) enters to say that his diphtheria serum didn’t work and many children have died—it had worked on all the animals but not on the children. Ehrlich tells him that every time his test horse is injected with snake venom, it produces more antibodies in the blood to protect itself. They become heroes when their 100-times more effective diphtheria serum is able to squelch a diphtheria epidemic. As an intertitle scrolls up the screen, 15 years pass. Ehrlich has won the Nobel Prize, is now director of an institute in Frankfurt, and is searching for chemical “magic bullets.” He tosses four members of the budget committee out of his office when they fail to understand why his team infects mice with trypanosomes and then tries to cure them. The committee is then disappointed to learn the “arsenic preparation” cured the mice but caused them to go mad such that they jump wildly in their cages. Ehrlich argues that chemicals can be altered to have any structure; they just need to find the one that works. When he learns that the bacterial cause of syphilis has been discovered and that it has a corkscrew shape like trypanosomes, they begin to work on syphilis, too. In the 3-minute scene from 1:07:30, Behring tours Ehrlich’s institute in what seems to be just a friendly visit. When Ehrlich says chemistry will loom large in the future of medicine, his old friend Behring responds that he is on a committee to review Ehrlich’s funding. He asks him to give up on this crazy idea of a chemical cure: “The idea of injecting chemicals into people’s veins fills me with horror.” Ehrlich gets angry and Behring leaves. In the 4-minute scene from 1:10:45, Ehrlich says, “I must work doubly fast before the committee can cut off my funds.” There is an overlapping montage of charts, chemicals, rats in cages, and researchers. The committee votes to cut Ehrlich’s budget by 50%, which prompts Hedwig Ehrlich to visit the rich widow Francisca Speyer (Maria Ouspenskaya) in hopes of getting more funds for her husband. In the 4-minute scene from 1:14:45, there are many beautifully dressed people dining at a long table engaged in jovial conversation. Someone asks Ehrlich about his research area, and he responds “syphilis.” A quick series of shocked faces follows. A woman breaks the silence to smooth things over, but it isn’t necessary because Speyer wants to know more. He talks of his “side chain theory,” where the disease is a keyhole and the molecule is the key. He says he is trying to fi nd a key that will only fit the disease and not the brain. She is fascinated and decides to give him funding. This is followed by another research montage consisting of numbered bottles, bubbling apparatus, a notebook page with the word “failure,” then ending with “Test 606: Complete Recovery.” Ehrlich calls together his staff (see figure 10.1 in chapter 10) to say he thought 100 tests would be more than enough and that they would fi nd the cure in a few months but it took years. Now, he says they need to test it in humans, and “I want you to promise that, if it fails, no one will ever know.” Ehrlich and most of the others leave, but Morgenrath (Edward Norris) realizes that Ehrlich intends to self-experiment. In the middle of the night, Ehrlich arrives in the laboratory only to fi nd that Morgenrath has performed the test on himself. He is a bit feverish but the cure works.
Good News
259
Commentary: Despite the complexity of Ehrlich’s life story and the impact of his research, the movie provides a reasonably accurate thumbnail sketch. Some discrepancies include that he was never a dermatologist, a number of individuals were conflated into one fictional character, the timing for several events was altered to simplify the narrative, his institutes were never in much fi nancial trouble, and he was never brought to trial for any deaths related to Salvarsan, the syphilis cure he developed (Baumler 1984; Lederer and Parascandola 1998). Ehrlich earned his medical degree from the University of Leipzig in 1878. For the next 10 years, he worked at the Charité Hospital in Berlin. While developing the basic fuschin stain for tubercular mycobacteria in 1887, he contracted pulmonary tuberculosis. He was cured during his travels to Egypt and other countries in 1888 and 1889. Upon his return, he was invited to join Koch at his new Institute for Infectious Diseases (now the Robert Koch Institute) in Berlin. While there, he began working and arguing with Emil Behring (later von Behring), who won the inaugural Nobel Prize in Medicine in 1901 for developing serum cures for tetanus and diphtheria. Ehrlich developed methods to boost the production of Behring’s anti-diphtheria serum, as well as methods to standardize the activities of different batches, and, fi nally, his own “side chain theory” of toxin–antitoxin interaction. He would later share the 1908 Nobel Prize in Medicine for his immune response work. In 1899, Ehrlich moved to Frankfurt, where he eventually transferred his “side chain theory” from the antibody–antigen interaction to the receptor–drug interaction. He was influenced in this by Emil Fisher’s idea of the “lock and key” scheme for molecular interactions. In 1904, he began searching for dyes that would cure trypanosomal sleeping sickness and, in 1905, syphilis. That year, Dr. Wolferstan Thomas of the Liverpool School of Tropical Medicine reported that the arsenic compound atoxyl was an effective cure for the sleeping sickness caused by Trypanosoma brucei gambiense (Riethmiller 2005). Ehrlich was very interested and immediately tested atoxyl. The compound had been synthesized many years earlier but then immediately forgotten until a report in 1902 described its effectiveness against certain types of cancers. It was also reported to be 40 times less toxic than inorganic arsenic (As2O3), a fact that would lead to the theory of differential toxicity in which the toxic dose had to be substantially higher than the effective dose. In late 1905, young organic chemist Alfred Bertheim joined Ehrlich’s group, and within three weeks of his arrival, Bertheim determined the correct structure of atoxyl (figure 9.3). This gave the team an edge over all the others because it allowed them to modify its structure in the search for improved properties. After three years, the 606th compound proved very effective against syphilis in rabbits and patients with late-stage dementia. Ehrlich named it Salvarsan, meaning “saved by arsenic.” In 1910, Hoechst chemical company created the fi rst batch of Salvarsan, but Ehrlich wanted to make sure it was as safe as possible and had it tested on 30,000 patients before allowing its commercial release. Even so, the team never stopped
260
ReAction! Chemistry in the Movies
H2N
As
NH2•HCl
HCl•H2N
O OH
As
HO
OH
Atoxyl
Arsphenamine (Ehrlich 606; Salvarsan)
ClH•H2N
H2N As
Oxophenarsamine HCl (Ehrlich 599; Mapharsen)
HO
O S
HN
O HO
OH
As
As As
O-Na+
OH
Neoarsphenamine (Ehrlich 914; Neosalvarsan)
Figure 9.3. Atoxyl was the first arsenical used in the search for a syphilis cure. Arsphenamine was the first commercial chemical antibiotic, also known as 606, Ehrlich 606, and Salvarsan. Neoarsphenamine has greater solubility, so lower doses can be used, resulting in lower toxicity. Oxophenarsamine replaced Ehrlich’s compounds in the United States because it was safer and more effective.
searching for a better compound and eventually found one with the 914th compound Neosalvarsan. Ehrlich died after a sudden illness in 1915, at a time when he was likely to receive a second Nobel Prize. In the 1930s in the United States, Mapharsen (figure 9.3) proved to be much safer (one death in three million injections) than either Salvarsan or Neosalvarsan (about one death in 25,000–50,000 injections). This work really began in the early 1920s when Carl Voegtlin of the U.S. Public Health Service studied the rate of trypanosome clearing from the blood in response to many arsenicals, including atoxyl, Salvarsan, and Mapharsen (Voegtlin 1925; Swann 1985). He noted that Mapharsen cleared the blood in less than an hour, whereas the others took as long as six hours. He proposed that most arsenicals were slowly oxidized to the effective Mapharsen-like compounds during the “latent period.” He, like Ehrlich and others before him, rejected using Mapharsen because he believed it was too toxic. This would change in the 1930s because of a unique collaboration between two academic researchers and a pharmaceutical company. After Cliff S. Hamilton earned his bachelor’s degree, he joined the Army’s Chemical Warfare Service during WWI where he may have worked with organic arsenicals for the fi rst time. In 1923, he earned a doctoral degree while working with arsenical synthesis expert Winford Lewis at Northwestern University. After Hamilton became a faculty member at the University of Nebraska, he and his students synthesized organic arsenicals that were tested for their effect on trypanosomal and syphilitic infections by Arthur Loevenhart at the University of Wisconsin Medical School. To obtain funding for their project, they entered into a contract with Parke, Davis and Company of Detroit in 1929. Parke-Davis
Good News
261
R R
R
As
As R
As As
As
As
R=
As As R
R
NH2•HCl
R
OH
R
Figure 9.4. Salvarsan is actually a mixture of structures involving arsenic rings. About one-third have three arsenic atoms in a ring, another third have five arsenic atoms, and the remaining third is a mixture, with four and six arsenic atoms being the next most abundant.
had been working in this area since 1910, when Ehrlich and Hoechst announced their important discovery. When Loevenhart died suddenly in late 1929, he was replaced by the recently hired Arthur Tatum. Because Tatum was the least familiar with this area of research, he began a series of preliminary tests and found that Mapharsen was actually the most effective agent. Instead of rejecting the agent as so many before, he proposed a theory stating that the important issue was not the absolute toxicity but the ratio of the curative dose to the toxic dose (Tatum and Cooper 1932). Hamilton helped Parke-Davis devise an efficient large-scale synthesis of Mapharsen, and by 1931, human trials proved it was very effective and even safer than Salvarsan or Neosalvarsan. By 1937, Mapharsen was on the market. Tatum and Hamilton split 5% of the royalties over the next 10 years. During WWII, sales reached their peak, and each was earning about $4,800 per year, enough to fund about four graduate students. By 1947, penicillin had replaced arsenicals in the treatment of syphilis, but Hamilton never stopped preparing and testing compounds. He eventually discovered therapeutic treatments for a number of tropical diseases such as sleeping sickness and malaria that weren’t displaced from the market until very recently. Even though Salvarsan doesn’t have any current economic value, its fame still attracts attention from the academic community. The greatest interest concerns the nature of the arsenic–arsenic bond (Lloyd et al. 2005). After two years of working with arsenicals, Ehrlich started drawing an arsenic–arsenic double bond in analogy to the nitrogen–nitrogen double bonds found in azo compounds. The assignment was reasonable because arsenic, phosphorus, and nitrogen are all members of fi fth periodic group and share many properties. Since 1907, however, many other nitrogen, phosphorus, and arsenic compounds have been prepared and studied such that the As=As bond has become less reasonable. To settle the matter, in 2005 a group of chemists at the University of Waikato in New Zealand were able to show that Salvarsan forms arsenic ring structures involving only single bonds (figure 9.4). The “latent period” for
262
ReAction! Chemistry in the Movies
Salvarsan’s activity must involve the slow oxidation of these arsenic single bonds. Mapharsen acts faster because the arsenic is already oxidized to prevent the formation of these oligomers.
RESEARCH CHEMISTS SOLVING PROBLEMS IN THE MOVIES Me & Isaac Newton (1999) Production company: Blue Sky Productions Director: Michael Apted Short summary: Documentary asks seven scientists why they do what they do Plot description: Gertrude Elion (figure 9.5) represents pharmaceutical chemists and their industry in this documentary about seven scientists. The other six scientists and engineers are Ashok Gadgil, who works on water purification; Michio Kaku, on string theory; Maja Mataric, on robotics; Steven Pinker, on language disorders; Karol Sikora, on gene therapy; and Patricia Wright, on lemurs. They discuss how they became interested in science, what type of work they do, their fi rst eureka moment, risk taking, and their thoughts about the future of science. When the facts from the movie are rearranged chronologically and dated, we learn the following about Elion. As a child growing up in New
Figure 9.5. Dr. Gertrude B. Elion, Nobel Prize Winner—Medicine. Photo Credit: Will and Deni McIntyre / Photo Researchers, Inc.
Good News
263
York City, she had an insatiable thirst for knowledge and read everything she could. She graduated from high school early, but her family didn’t have enough money to send her to college because of the Great Depression. With her excellent grades, she was admitted to Hunter College, which didn’t charge admission at that time. She decided to become a science major because her grandfather died of cancer. When she graduated in 1937 with a chemistry degree, she tried to get into a number of graduate programs, but none would provide a fellowship. It was only later that she realized it was because of sexual discrimination. She also had trouble getting a job as a chemist and eventually enrolled at New York University. Upon graduation with a chemistry master’s degree in 1941, she landed a job testing pickles for acidity and mayonnaise for color. When that company folded, she was hired in 1944 by George Hitchings as his assistant at the Burroughs Wellcome Company (now GlaxoSmithKline) in suburban New York. In the 2.5-minute interview at 55:45, Gertrude Elion discusses the impact of the fi rst major drug she and Hitchings developed in the early 1950s. She says that it has cured many cases of childhood leukemia, but she does not name the drug or show its structure. In 1988, she and Hitchings shared the Nobel Prize in Medicine for their discovery and methods. She was very happy to get the Nobel Prize, of course, but observes that there is no greater joy than to know your work has had an impact on people’s lives. Commentary: Each one of the stories in this documentary is inspiring, which makes it particularly useful in the chemistry classroom (Griep and Mikasen 2005). Since Elion’s drugs are never named or explained in the documentary, part of the assignment for students should be a treasure hunt on the Internet for the structure of the fi rst drug that Elion and Hitchings developed. When George Hitchings was 12 years old, he lost his father to illness (Kresge et al. 2008b). Years later, he enrolled at the University of Washington to pursue premedical training but quickly switched to chemistry. He earned a chemistry bachelor’s degree in 1927 and master’s in 1928. After earning a doctorate at Harvard in the area of analytical methods to study purine metabolism, he held temporary jobs in his area of expertise. Purines are some of the building blocks for DNA and RNA, which were becoming hot areas of research. In 1942, he was hired as fi rst head of the new biochemistry department at Burroughs Wellcome. He was impressed with the recent development of sulfa drugs, the fi rst synthetic drugs that killed bacteria by interfering with a natural metabolic pathway. Hitchings believed he could develop synthetic drugs that would interfere with DNA synthesis, his area of expertise. After two years on the job, Hitchings hired Gertrude Elion as his research assistant (McGrayne 1993; Avery 2000; Kresge et al. 2008a, 2008b). Six years later in 1950, Elion had her “WOW!” year during which she synthesized two effective cancer treatments (figure 9.6). They used
264
ReAction! Chemistry in the Movies
SH H N
N N
N
6-Mercaptopurine (6-MP; Mercaleukin)
Figure 9.6. 6-Mercaptopurine, the antileukemic compound developed by Elion and Hitchings in the 1950s.
bacteria to study how these two nucleoside analogs were incorporated into DNA. This was the fi rst step that allowed them to work out the order in which DNA building blocks were incorporated into the DNA inside the bacteria. Their pioneering methods would become known as “rational” drug design. They used these methods to fi nd a number of synthetic compounds that have since become some of the most celebrated discoveries in medicine. Their methods earned them the Nobel Prize in Medicine more than 35 years later. Elion and Hitchings developed the fi rst successful anticancer drug, called 6-mercaptopurine, an immune suppressant useful during organ transplants called azathioprine, and an effective treatment for gout called allopurinol. In 1967, Hitchings retired and Elion became head of the Department of Experimental Therapy. Her team began working on antiviral compounds and developed acyclovir, still the preferred treatment for herpesvirus. Even though Elion retired in 1983, she continued working as a Burroughs Wellcome Emerita Scientist and consultant. The next year, her team developed azidothymidine (zidovudine, AZT), the only drug licensed to treat AIDS until 1991. She died in 1999, and Me & Isaac Newton is dedicated to her. Medicine Man (1992) Production company: Hollywood Pictures Director: John McTiernan Screenwriters: Tom Schulman and Sally Robinson Short summary: Ethnobotanist Robert Campbell and biochemist Rae Crane use a gas chromatograph to fi nd a cancer-fighting drug in the Amazon forest Plot description: During the opening credits, “Dr. Crane” is hidden from the camera, but we know he or she is on a plane to South America with boxes labeled Aston Company. The forests are burning below and major strips of land have been cleared due to logging. A voice-over tells us that Robert Campbell, botanist, has been in the forest for three years and has requested an assistant and a gas chromatograph. The still unseen Crane
Good News
265
HO
O O
Figure 9.7. Imaginative anticancer agent eluting as “peak 37,”described as “Mother Nature’s kitchen,” but otherwise never named. In addition to the “acid derivative” mentioned in the movie, it has anthracene and tricyclohexane components. It is most similar to the real alkaloid reserpine, isolated from snake root.
lands and is transported by car, canoe, and foot to a small clearing in the jungle. Only then do we see that Crane is a woman (Lorraine Bracco). Robert Campbell (Sean Connery) appears out of nowhere and yells in disgust, “They sent a girl?” Her response establishes her credibility; she is a biochemist and she knows how to use a gas chromatograph (GC). In the 1.75-minute scene from 22:15, it is an idyllic morning when Crane fi res up her GC and injects the sample Campbell has been saving for so long. The instrument’s computer screen shows baseline separation of 49 peaks, all of which are named by the computer. As she examines the peaks, she says “No” to the salts. When she clicks on the peak numbered 37 (figure 9.7), she says, “Looks like an acid derivative. This one’s Mother Nature’s kitchen. No one will ever be able to make this. End of story.” When Campbell won’t tell her why the sample is important, she gets angry and tells him she’s been sent to replace him. He softens and reveals that one of the local boys was cured of stage two posterior lymphoma by the mixture. Crane is still skeptical and even willing to believe he faked the data. To prove it, they inject two guinea pigs with a carcinoma cell line, and later, one is injected with the compound. The next morning, Crane performs the biopsies and this convinces her. He stopped sending reports because he has been unable to reproduce “the serum.” They enter the jungle in search of the bitter-tasting flower the natives chew to prevent cancer. They fi nd one in the treetops, and he tells her it is a bromeliad. The forest burns in the distance. After gathering a number of flowers, they enlist the help of the locals to prepare the sample in as many ways as possible. Crane suggests that Campbell didn’t follow the medicine man’s procedure exactly. In the 1.5minute scene from 1:03:00, a boy gets lymph cancer and Campbell wants to use the last bit of serum, but Crane argues that the drug belongs to the world, not just one sick kid. In one last attempt to learn the secret, they
266
ReAction! Chemistry in the Movies
search for and fi nd the medicine man. He says the bromeliad is called sky flower and that it is only a house for bugs. When they return to the village, the boy is near death, so Crane injects the last of the serum while the natives dance around the fi re. The next morning, the boy is cured. Two white men then enter the camp to say that they are going to bulldoze the area. Crane has only been in the camp one week, and they still don’t have the compound. Commentary: Even though this movie’s narrative and molecule are pure fiction, it is probably safe to assume that any screenplay about a South American ethnobotanist is based in part on the life of Richard Schultes (Davis 1997). In 1938, Schultes was 23 years old and living in Mexico when he reported the existence of the hallucinogenic psilocybin mushroom in the scientific literature for the fi rst time. From 1941 to 1953, he lived among indigenous people of the Amazon rain forest while mapping rivers, collecting more than 27,000 botanical samples, and learning about natural medicines from shamans. For nearly 50 years after that, Schultes was a professor of botany at Harvard University. He was Wade Davis’s thesis adviser in the 1980s when Davis traveled to Haiti in search of the zombie powder (see description of The Serpent and the Rainbow, below). Lorenzo’s Oil (1992) Production company: Universal Pictures Director: George Miller Screenwriters: George Miller and Nick Enright Short summary: When Lorenzo Odone develops a strange disease at age 7, his parents learn everything they can about it and then develop a dietary cure that slows the disease; based on a true story Plot description: The actors Nick Nolte and Susan Sarandon appear on camera before the movie begins. Sarandon says Lorenzo’s parents Augusto and Michaela Odone fought courageously for their son’s life and are now writing a second chapter. “It is called The Myelin Project, and it’s providing real hope to thousands suffering from diseases such as multiple sclerosis.” Then, she provides a toll-free number and website (www.myelin. org) and urges viewers to pledge their support. The movie opens with a beach scene on the Comoros Islands, off East Africa. It is July 1983 and four-year-old Lorenzo Odone (Zack O’Malley) is having fun in the tropical paradise. When he returns to Washington, D.C., that fall with his parents, Augusto and Michaela Odone (Nick Nolte and Susan Sarandon), he has trouble adjusting to his new school. After some confusion over his increasingly angry outbursts, he is diagnosed with adrenoleukodystrophy (ALD), a rapid neurodegenerative disease for which there is no treatment. He has less than two years to live, so his parents learn everything they can about the disease. When
Good News
267
Dr. Gus Nikolai (Peter Ustinov) tells them it is an X-linked genetic disease, Michaela experiences profound guilt because it means her son inherited it from her. A discussion with her sister reveals that their family has a history of sons dying young. They fi nd no comfort in the ALD family support association. When they pull Lorenzo out of a clinical trial because his fatty acids levels were rising rather than falling, the husband and wife leaders of the association tell them they should not outguess the clinicians since valid results require completion of the trial. Augusto’s response is to spend months reading the biomedical and genetics literature in the U.S. Library of Congress. In one of the early shots, he is reading Lubert Stryer’s Biochemistry, an advanced introductory textbook. Augusto and Michaela convene the fi rst international ALD symposium, where researchers from around the globe share their ongoing research with each other. They learn that one of the biochemical symptoms is a high concentration of very-long-chain fatty acids (VLCFAs) in the blood. Even though it may be the cause of the problem or the result of some other effect, it is their best lead, so they pursue it. VLCFAs enter the body from two sources: diet and biosynthesis. Since there are only a few foods that are high in VLCFA (fatty meat, nuts, and the waxy coating of fruits and vegetables), it is easy to restrict dietary intake. The more difficult problem is turning off the natural biosynthetic pathway. Eventually, the Odones hypothesize the target enzyme has two fatty acid binding sites, and if they block one site with “Lorenzo’s Oil,” they will inhibit the activity of the other site that produces VLCFAs. Their fi rst supplier is an American company that had prepared the oil for a project since discontinued. When the supply runs out, the Odones search the globe for a new supplier. They fi nd a cosmetics chemist at Croda Universal in Hull, England, who agrees to make some. By regular measurements of Lorenzo’s blood VLCFAs, they prove that the special oil halts the progression of the disease but does not reverse the demyelination that has already taken place. Augusto says this gives them a new mission, and the movie ends with real boys saying they were helped by Lorenzo’s Oil. Commentary: The chemist Don Suddaby (who plays himself in the movie [Jones 2000]) is one of the biggest heroes of this story. He succeeded in producing Lorenzo’s Oil (figure 9.8), a 4:1 mixture of the triglycerides of C22:1 (erucic acid with cis-C13) and C18:1 (oleic acid with cis-C9). As a biochemist at Croda Universal in Hull, England, his lifetime of experience working with triglyceride oils made him ideally suited for this job. His most difficult task was to isolate large quantities of erucic acid from rapeseed oil. Dr. Hugh W. Moser is the ALD expert upon whom the fi lm’s Dr. Gus Nikolai (Peter Ustinov) is primarily based. After the movie was released, he wrote an editorial (Moser 1993) in which he said, “As a work of fiction, Lorenzo’s Oil is an excellent fi lm. However, as a factual documentary it has three main flaws: it overstates the success that can be achieved with the oil, it invents confl icts between the parents and the medical
268
ReAction! Chemistry in the Movies O C O C
O CH2 O
O C
CH CH2
O
Erucic Triglyceride Figure 9.8. Erucic Triglyceride is also called Lorenzo’s Oil.
establishment, and it presents an inaccurate and malicious portrayal of a valued parents’ organization [the United Leukodystrophy Foundation].” In that same year, 1993, the results of a controlled clinical trial lasting two years with young ALD patients were published. On the positive side, a diet of 20% Lorenzo’s Oil normalized the VLCFA blood levels. On the negative side, it did not stop progression of the disease (Aubourg et al. 1993). In the hope that Lorenzo’s Oil could act as a preventative, a 10-year study was undertaken by Dr. Moser and coworkers (Moser et al. 2005). This study focused on patients who had one of the gene’s disease alleles but who had not yet developed any symptoms. In 2005, the study reported that Lorenzo’s Oil substantially reduced the development of symptoms in these at-risk asymptomatic patients. Science can be messy and frustrating, and medical science even more so. Since this movie demonstrates many important issues about medical science but with the added twist of a nutritional chemical cure, its pedagogical utility for reaching students in the classroom has been noted (Wink 2001; Griep and Mikasen 2005). On the other hand, bioethicists are divided as to whether the movie creates problems for physicians. Bioethics professor Anne Hudson Jones points out that physicians deal with misinformation from the media all the time, so the falsehoods presented in this movie are not overly problematic (Jones 2000). Despite the serious factual flaws, she believes the fi lm is valuable to physicians in presenting its main point about the inherent confl ict between the goals and needs of researchers and patients. In contrast, physician poet M. Roy Jobson and bioethicist Donna Knapp van Bogaert are more troubled by the “fi lm’s close association with actual events and real people” (Jobson and Knapp van Bogaert 2005). They feel the misrepresentations and the falsely raised hopes could be considered unethical. The earliest symptom of ALD is attention deficit disorder, followed by rapid impairment of cognition, vision, hearing, and motor functions. Any one of 200 different mutations to the ABCD1 gene can cause ALD. The gene codes for ALD protein and is located on one end of the X
Good News
269
chromosome, making this a maternally inherited genetic disease. Certain mutated forms of the gene (called disease alleles) make a dysfunctional version of the protein that can’t transport VLCFA into liver peroxisomes, the organelles that degrade fatty acids two carbons at a time to generate energy. The result is that blood levels of saturated VLCFA increase, namely, C24:0 (lignoceric acid) and C26:0 (cerotic acid). Somehow, this causes demyelination of nerves and destruction of the adrenocortex. The myelin sheath allows nerves to conduct action potentials, and the adrenocortex produces some hormones. Lorenzo Odone died in 2008, one day after his 30th birthday. He outlived his two-year prognosis by 22 years. The Serpent and the Rainbow (1988) Production company: Universal Pictures Director: Wes Craven Screenwriters: Richard Maxwell and Adam Rodman, loosely based on Wade Davis’s 1985 same-titled popular ethnography Short summary: Ethnobiologist Dr. Dennis Alan searches for the molecule that causes zombinism in Haiti Plot description: The movie opens with text informing us, “In ancient voodoo, the serpent is the symbol of Earth and the rainbow is Heaven.” In 1985, the pharmaceutical company Biocore asks young Harvard ethnobiologist Dr. Dennis Alan (Bill Pullman) to retrieve a sample of the Haitian zombification drug. They think it might be useful as an anesthetic in surgery. In Port au Prince, Haiti, Alan travels to the hospital that pronounced Christophe Durand (Conrad Roberts) dead 15 years ago. He has since returned to life. Alan meets nurse Marielle Duchamp (Cathy Tyson), who agrees to help fi nd Durand. Once they do, he says a man used a powder to enslave his soul and then beat him to keep him under control. He eventually escaped but roamed the country anonymously for many years because he feared being caught again. Durand and Duchamp then help Dr. Alan fi nd a bokor, or voodoo sorcerer, named Louis Mozart (Brent Jennings). He agrees to sell them some powder and shows how it will put a goat to sleep. Alan says that when he sees the goat rise from the dead, he’ll pay. Soon afterward, the radio reports, “Things are unstable in Haiti.” It seems the government is disintegrating into chaos. When they revisit Mozart the next day, they see a goat but it doesn’t have the mark that Alan put on it, so he refuses to pay. Mozart then agrees to make some real powder with Dr. Alan’s help. That night, they dig up a coffi n and remove a recently dead woman’s head for use in creating the powder. When they return home that night, the chief of police, Lucien Celine (Paul Winfield) arrests them and tortures Alan to discover his reasons for being there. It takes him three days to recover.
270
ReAction! Chemistry in the Movies
In the 3.75-minute night scene from 1:01:15, Mozart shows Alan how to prepare the powder. In voice-over narration, Alan describes how Mozart prepares a poisonous sea toad Bufo marinus, puffer fi sh for tetrodotoxin, herbs, and other things. The process takes three days and nights and involves heating with fi re. Finally, he adds the brains of the dead woman and buries the mixture for one more day. Alan tells Mozart he wants to use the poison for good and Mozart laughs. Despite another run-in with the chief of police, Alan is able to bring some of the powder back to Harvard, where it causes the appearance of death in baboons. But they discover the baboons are conscious, and its effects wear off 12 hours later. This means human victims would wake up screaming in their coffi ns. At a cocktail party, they joke about their new anesthetic “zombinol.” Alan returns to Haiti for more adventures but no more chemistry. The end credits read, “The zombie powder and its active ingredient tetrodotoxin is currently under active study in Europe and the US. To this day, the process by which it works remains a mystery.” Commentary: Tetrodotoxin is not suitable for use as an anesthetic (figure 9.9). It binds to sodium ion channels in nerve axons to block the flow of sodium ions across the nerve membrane and, therefore, the transmission of the nerve impulses. The intoxicated person is fully conscious but isn’t able to move. In contrast, anesthetics reduce consciousness and induce forgetfulness. Davis proposed that the zombie powder’s key ingredient was tetrodotoxin due to the puffer fish ingredient. Consumption of this fi sh has caused many deaths and a few near deaths. There is one recorded case of intoxication in Japan, where a victim woke just as his body was about to be placed in a crematorium (Halstead 1988). He had been conscious throughout. The biology of tetrodotoxin has been clarified since the time this movie was released and Wade Davis earned his graduate degree (Simidu et al. 1987; Lee et al. 2000). Tetrodotoxin-producing Vibrio bacteria live,
OHO O
+H2N NH HN
OH O
HO
Tetrodotoxin
OH OH
Figure 9.9. Tetrodotoxin is the potent neurotoxin produced by certain marine Vibrio bacterial species that live symbiotically in puffer fish and other living creatures.
Good News
271
eat, and reproduce within a variety of hosts, including several species of puffer fi sh, the Costa Rican harlequin frog, and the Australian blueringed octopus. The hosts use the toxin for defense and/or predation. Of all the movies “based on a true story” in this chapter, this one takes the most liberties by far. Anyone watching this movie would conclude that zombinism is a fact and that tetrodotoxin is the cause. Both of these points require a bit of explanation. The script was inspired by Wade Davis’s popular account of his real search for the zombie powder (Davis 1985) and not by his subsequent less sensational account that provided more scientific details (Davis 1988a). His Harvard graduate studies adviser was Richard Schultes, one of the leading ethnobotanists (see the Medicine Man commentary). Schultes had learned about a Haitian man named Clarvius Narcisse who was pronounced dead in 1962 at the Albert Schweitzer Hospital in Deschapelles, Haiti, but who showed up alive in 1980 claiming to be an escaped zombie. Schultes sent Davis to meet Narcisse and to learn about the ethnography of zombinism. In his thesis, Davis proposed that zombies occur because some Haitians believe they can occur and because the zombie powder makes the occasional person appear to be dead. The poud zombi (zombie powder) is created by voodoo sorcerers in the Bizango secret society and only used on those who violate its laws. The point of Davis’s thesis that has attracted the most criticism concerns the amount of tetrodotoxin in the zombie powders he collected (Davis 1988b). It was present in low quantity in only one of two samples analyzed by C. Y. Kao of the Downstate Medical Center in Brooklyn, New York. Davis provided a number of reasons for this disparity, including the powder is not made by a pharmaceutical laboratory and its components are variable, tetrodotoxin is sensitive to pH and may have degraded before or during analysis, some of the powder’s other ingredients may enhance transport of tetrodotoxin across the blood–brain barrier, and “one success in dozens of attempts [by the sorcerer] would be sufficient to support the cultural belief in the zombie phenomenon.” Clambake (1967) Production company: United Artists Director: Arthur H. Nadel Screenwriter: Arthur Browne, Jr. Short summary: Petroleum chemist Scott Heyward creates a superhard, super-fast-drying varnish that allows him to win the boat race, the heart of the girl he loves, and his father’s respect Plot description: Scott Heyward (Elvis Presley) pulls his hot car into a gas station/hamburger stand where an old geezer fi lls the gas tank and Heyward orders a hamburger. While chatting with Tom Wilson (Will Hutchins), he says he is a chemist and son of oil magnate Duster Heyward
272
ReAction! Chemistry in the Movies
(James Gregory). He’s in Florida to fi nd a girl who will love him for himself and not for his millions. Wilson thinks having girls run after you just because you’re worth millions can’t be all bad, so they trade places. They head to the Shores Hotel on Miami Beach, where Wilson was going to be the summer waterskiing instructor. At the hotel, Diane Carter (Shelley Fabares) and all the other girls are listening intently as James J. Jamison III (Bill Bixby) describes last year’s boat race, which he won. Sam Burton (Gary Merrill) of Burton Boat Company lost last year’s race because the “resin on the boat was no good.” It just so happens that Heyward was developing a special fast-drying, superhard varnish that he calls “GOOP, the technical name is a foot long.” He asks Burton whether he can make a go of it using Burton’s boat. Burton says OK but the race is only seven days away. For the next six days, Heyward gives waterskiing instructions during the day, sings to Carter in the evening, and works on the boat all night. At 52:20, Heyward tells Burton, “I was trained as an engineer.” (We know he meant chemical engineer because he is surrounded by distillation apparatus fi lled with bubbling, colored solutions.) Back at Heyward Oil Company, they sell off the rights to the GOOP formulation and the existing experimental stock. Its full name is glycol oxyoctanoic phosphate. Buster learns that his secretary knows where his son is—she just sent the experimental stock to him. The 1.5-minute scene from 1:00:30 begins with Heyward asleep at the bench behind a centrifuge with spinning test tubes. It is his sixth night of work in a row. Burton wakes him so he can hand over the box of materials he just received. Heyward sniffs them enthusiastically and wants to get started right away. The 3.5-minute scene from 1:02:00 begins with Heyward varnishing the boat, and then Wilson arrives to help. When he claps his hands, six girls enter dancing from all directions and Heyward sings “Hey Hey Hey,” a song about helping to varnish the boat with “glyooxytonic-phosphate.” When the song is over the boat is ready, but there is no time to test the boat before the race. Heyward takes the risk and self-experiments. Commentary: Varnish is often called goop because of its high viscosity. In this movie, GOOP is an acronym for a varnish with the technical name glycol oxyoctanoic phosphate. Glycol, octanoic acid, and phosphate are real compounds (figure 9.10) and can be used to devise a structure for GOOP. To solve the structure, we also need to assume that “oxy” is short for “epoxy” and that it should be placed at the omega-3 location (three carbons from the reduced end) in an analogy to linoleic acid, which is part of the traditional linseed oil varnish. An epoxy is an activated form of oxygen and would result in rapid cross-linking of adjacent molecules. This would yield a fast-drying, superhard surface. It is analogous to linseed oil varnish, which cross-links slowly when molecules of oxygen from the atmosphere react with its double bonds.
Good News
273
OH
O O -O C
HO
OH
Glycol
Octanoic Acid
P
O-
O-
Phosphate
O O
O
O
O
O
P
O-
O-
C OH
5,6-Epoxyoctanoic Acid = Oxyoctanoic Acid?
O
GOOP
Figure 9.10. The parts of glycol oxyoctanoic phosphate can be joined together to form GOOP.
Forever, Darling (1956) Production company: Metro-Goldwyn-Mayer Director: Alexander Hall Screenwriter: Helen Deutsch Short summary: Susan Vega’s guardian angel convinces her to join her husband Lorenzo on his insecticide 383 field test to save her marriage Plot description: The movie opens with Susan and Lorenzo Vega’s wedding dance (Lucille Ball and Desi Arnaz). Susan comes from a moneyed family, whereas chemist Lorenzo is a partner in the Finlay-Vega Chemical Company. After five years of marriage, Susan no longer kisses Lorenzo as he leaves for work in the morning. In the 3.0-minute scene at 6:15, Lorenzo’s partner Bill Finlay (John Hoyt) brings Mr. Oliver Clinton (Willis Bouchey) from the D.C. office to meet Lorenzo. Finlay says, “To Larry, [insecticide] 383 is a crusade. He hasn’t been home to dinner for a month,” and “We think it will make DDT look like talcum powder.” When Clinton leaves, Lorenzo admits to Finlay that he would like to do the field tests himself. When Lorenzo arrives home that night to tell Susan, she is dining with her cousin (Natalie Schaefer). During dessert, Lorenzo becomes so angry at the meddling cousin that he blurts out that he and Susan will be gone for the next two years performing field tests on mosquitoes in South America, the tsetse fly in Africa, and cockroaches in Puerto Rico. The cousin is disgusted by the thought of all those dead roaches. The outburst comes as a surprise to Susan, who likes her cousin and doesn’t mind her meddling. In a short scene in the garden at 41:30, Susan angrily tells Lorenzo, “Go smoke a test tube. You and your science. One of these days
274
ReAction! Chemistry in the Movies
you scientists are going to blow up the world.” He responds, “That’s a sore point with scientists. We just want to make the world a better place to live in,” to which she adds “For everybody except your wives.” That night, Susan is visited by her guardian angel (James Mason), who eventually convinces her to join Lorenzo on his trip. Commentary: This is one of the few movies to explore a chemist’s balance between work and home. A brief history of DDT was given in chapter 4, but this movie confi rms it was still considered fully humanitarian to develop an insecticide in 1956. Monkey Business (1952) Production company: Twentieth Century Fox Director: Howard Hawks Screenwriters: Ben Hecht, Charles Lederer, and I. A. L. Diamond Short summary: Chemist Barnaby Fulton self-experiments with a youth formula he created Plot description: Chemist Barnaby Fulton (Cary Grant) works for Oxley Chemical Company. His wife Edwina (Ginger Rogers) explains, “He’s not often the absent-minded professor but when he is, he goes all the way.” At work, Fulton examines secretary Lois Laurel’s legs (Marilynn Monroe) at her urging because she is wearing the M41 acetate nylons he invented. They can’t tear or snag. Fulton’s current project is an antiarthritis compound that his boss Mr. Oliver Oxley (Charles Coburn) wants to market as B-4, a sort of youth formula. When Oxley sees an active chimp in Fulton’s lab and thinks that its arthritis was cured by formula X57, Barnaby tells him he is looking at young chimp Esther and not old chimp Rudolph. In the 6-minute scene from 26:45, Fulton prepares formula X58, containing sodium ascorbate and sodium molybdate, among other things. He will heat this sample to compare its properties with cooled and ambient samples. Esther watches his every move, and when he leaves, she gets out of her cage and mixes up her own formula. After a quick cut to Fulton saying, “Ooh ooh ooh” like a monkey and complaining about his bursitis, the action returns to Esther, who gargles her solution but pours most of it into the water cooler just before the jug is replaced. Fulton returns to his laboratory to say, “Self-experimentation is against all the rules,” but “History was made by people who didn’t follow the rules.” He drinks some of his formula and washes it down with some water from the cooler, saying “It tastes bitter.” Commentary: Even though the chimp Esther’s role was to serve as a control for tests on the older chimp Rudolph, she successfully creates a youth formula when the real chemist could not. The experimentalist’s and test subject’s roles are further inverted when, by chance, the chemist drinks some of Esther’s formula.
Good News
275
Strange Impersonation (1946) Production company: Republic Pictures Director: Anthony Mann Screenwriter: Mildret Lord, from a story by Lewis Herman and Anne Wigton Short summary: Pharmaceutical chemist Nora Goodrich develops new anesthetic, self-experiments, and then her life goes noir Plot description: The movie opens with sequential shots of New York City, an imposing building, a glass door to the Wilmott Institute for Chemical Research, and a conference room in which a woman wearing horn-rimmed glasses is giving a presentation. In the 4.25-minute scene from 3:00, Dr. Nora Goodrich (Brenda Marshall) describes the anesthetic she’s developing. She says the chemical tests are almost complete. It will soon be tested in humans to look for side effects, such as whether it causes hallucinations. When her lecture is over, everyone congratulates her, including her fiancé, coworker Steven Lindstrom (William Gargan), and her assistant, Arlene (Hillary Brooke). Cut to the lab, where Goodrich is working at a bench surrounded by equipment. Lindstrom sneaks in to surprise her, which causes her to mix the chemicals too fast. The solution bubbles over and catches fi re. When the brief crisis ends, Lindstrom sends Arlene on a false search for a book on osmosis so he can try to persuade Goodrich to marry him earlier than she had promised. She wants to wait until after she’s tested her formula. After Lindstrom leaves, Arlene says she wouldn’t allow an experiment to prevent her from getting married. Goodrich says every woman has to decide for herself. That night at her apartment (she has chemical glassware in her kitchen cabinets), Arlene shows up to monitor her pulse and breathing rate. Goodrich injects herself with the anesthetic, lays down on the couch, and then her life goes noir. Commentary: General anesthesia for surgery was discovered in the 1840s (see the description of The Great Moment, below) and the following compounds were soon shown to be effective: chloroform, diethyl ether, nitrous oxide, and ethyl chloride (figure 9.11) (Robinson 1946).
H3C H2C H3C
O
C
CH2 CH3
F3C
H2C
O
CH CH2
CH
O
CH2F
F3C
Diethyl Ether Isopropenyl Vinyl Ether
Sevoflurane
Figure 9.11. General anesthetics discovered in the mid-1800s, the 1930s, and 1990s. Sevoflurane is the most common general anesthetic in use today because it has nearly ideal properties (Delgado-Herrera et al. 2001).
276
ReAction! Chemistry in the Movies
Of these, only nitrous oxide is still used today because the others are too flammable or cause nausea. Research for anesthetics began anew in the 1930s and 1940s when the following anesthetics were discovered: ethylene, cyclopropane, trichloroethylene, and isopropenyl vinyl ether. All of these were flammable, and any one of them could have been the one that Nora Goodrich was supposed to have developed in the movie. Its identity can be narrowed further by noting that she intravenously injected her anesthetic rather than inhaled it. Injection would cause it to act quickly (within 30 seconds) but also suggests the compound was a liquid at room temperature. Therefore, Goodrich’s anesthetic is most likely modeled on isopropenyl vinyl ether. The nonflammable inhalation anesthetics that we use today were discovered between the 1950s and 1990s: halothane, isoflurane, enflurane, sevoflurane, and desflurane. Replacing a compound’s hydrogens with halogens (fluorine, chlorine, and bromine) generally retards combustion. According to the American Society of Anesthesiologists, the ideal inhalation anesthetic would have the following physical and chemical properties at room temperature: (1) nonflammable and nonexplosive, (2) vaporizable, (3) stable with long shelf life, (4) nonreactive with plastics or metals, (5) environmentally friendly, and (6) cheap and simple to manufacture. The ideal biological properties would be (1) pleasant to inhale and nonirritating, (2) fast onset (low solubility in blood), (3) high potency (high solubility in oil), (4) minimal side effects, (5) excreted from lungs unchanged, and (6) nontoxic to operating personnel. The Great Moment (1944) Production company: Paramount Pictures Director: Preston Sturges Screenwriter: Preston Sturges, based on René Fülöp-Miller’s 1938 book Triumph over Pain Short summary: Bemused biography of Boston dentist William T. G. Morgan, who discovered the fi rst anesthetic—sulfuric ether Plot description: Behind the opening credits, a crowd cheers for William Thomas Green Morton (Joel McCrea) of Boston, the inventor of painless dentistry. A written prologue alerts us that the story concerns sulfuric ether as well as personal struggles and sacrifices. (Sulfuric ether is an archaic name for diethyl ether, which is made by using concentrated sulfuric acid to dehydrate two molecules of ethanol so they will form diethyl ether.) The story begins when elderly Eben Frost (William Demerest) redeems a medal “to the benefactor of mankind, with the gratitude of humanity” from a pawnshop and brings it to the elderly widow Lizzie Morton (Betty Field). Her reminiscence leads to two flashbacks. The fi rst flashback lasts only six minutes but covers the 20 years after Morton’s fi rst surgical use of ether. He is a poor but content farmer
Good News
277
because he hasn’t earned any royalties from his patent. When the Senate passes a bill to award him $100,000 for it, President Franklin Pierce won’t sign and asks Morton to sue a Navy doctor using his method. He wants to prove in court that Morton has legal priority to the discovery of anesthesia. Unfortunately, public sentiment turns against him, and then the court rules that his discovery is not patentable. In the 1-minute scene from 10:00, he enters a medical supply shop to smash the glass vaporizers they are selling without paying him royalties. After Morton dies a broken man, the story returns to his widow, who reminds Frost of the other men who claimed to have invented anesthesia, dentist Horace Wells (Louis Jean Heydt) and chemistry professor Charles T. Jackson (Julius Tannen). At dinner, Lizzie Morton begins her longer reminiscence that lasts the rest of the movie and that actually took place over a matter of months. Morton becomes a dentist because he is too poor to go to medical school. He soon discovers that toothaches have to be more painful than the cure before people will visit him. So, he visits his former Harvard Medical School chemistry professor Jackson for advice. Jackson suggests “something with a low boiling point” such as ethyl chloride to cool the nerve. Morton visits a pharmacist, gets confused, and purchases some sulfuric ether instead. He brings it home, sets it on a table next to the fi replace, and falls asleep as its boiling fumes fi ll the room. Next, his old dental friend Wells pays him a visit. He wants to use the well-known nitrous oxide, or laughing gas, as an anesthetic. Together, they administer enough nitrous to a patient to make him laugh uncontrollably and then more to knock him out. When they pull his tooth, the patient wakes up and struggles. They almost kill their second patient, but Morton is inspired to read his old medical texts and decides to try sulfuric ether inhalation. In the scene at 34:30, Morton fi rst works out the proper dose on a rabbit and scales it up for a patient. The operation is a success, but he worries that he asphyxiated her. When she wakes up hours later, she is very pleased with the results. He realizes he needs to learn more about dosing, so he experiments with the pet goldfi sh and then himself. Next, he decides to call it Letheon for “the river that banishes sorrows” to protect its identity. Finally, he visits Jackson’s laboratory to learn the difference between sulfuric ether and chloric ether and is so grateful he tells Jackson he’ll include him in any patent arrangement. On September 30, 1846, Morton performs the fi rst successful painless tooth extraction on a patient named Eben Frost. His business grows quickly, and he spends his spare time testing various ether doses on Frost. He thinks it could be used during surgery, so they visit Massachusetts General Hospital. They watch professor Warren (Harry Carey) perform surgery after offering the patient alcohol to dull the pain and then placing him in restraints. On the next day, October 16, 1846, Morton administers ether to a patient, allowing Warren to perform the surgery without the patient feeling pain. Everyone is happy except two snooty
278
ReAction! Chemistry in the Movies
physicians. They don’t want the medical society to be overshadowed by a mere dentist. Some time passes, but suddenly everyone is excited to observe the second demonstration of Letheon anesthesia. While waiting for Morton to arrive, the snooty surgeons remind Warren that dentists keep their patents secret to make money. There is no Dental Hippocratic Oath. They insist that physicians can’t use or promote patent medicines. When Morton fi nally arrives, he is reluctant to reveal the secret because his patent still hasn’t been approved. He offers to provide Letheon free of charge to all surgeons who wish to use it, but the snooty surgeons stand their ground. As Morton begins to leave, he sees the young girl whose leg is about to be amputated. He returns to the amphitheater accompanied by a musical fanfare. Commentary: Morton’s efforts did change the nature of surgery, but exactly what did he discover: anesthesia, ether anesthesia, or the inhaler to deliver the ether to the patient? Of these three, he patented an inhaler (U.S. patent 4848 on November 12, 1846), but he was never satisfied with its design and made several changes to it (Desbarax 2002). In fact, within six months of the fi rst public demonstration, he no longer used any apparatus at all, according to his own letter to Lancet medical journal (Morton 1847). He said he now preferred a hand-sized sponge soaked in ether. Since the inhaler wasn’t essential to the delivery of ether anesthesia, there weren’t grounds for Morton to press his suit of patent infringement. The discovery of general anesthesia has a complicated history (Smith and Daniels 1998). Though many people had the intelligence and opportunity, there is no doubt that Boston dentist William Morton was responsible for its fi rst public demonstration during surgery. Upon reading about that fi rst demonstration, Oliver Wendell Holmes wrote a letter to Morton suggesting that the procedure be called anesthesia, which is Greek for “without feeling” (Holmes 1847). Many general anesthetics have been developed in the years since 1846 (see description of Strange Impersonation, above), but it is still interesting that there is no fully satisfactory theory for their mechanisms of action (Antkowiak 2001). The earliest clue appeared in about 1900, when physicians Meyer and Overton discovered a linear relationship between an anesthetic’s potency and its ability to dissolve in olive oil. Since then, the potency tests have become more sophisticated. The relationship still holds, except we also know there are many oil-soluble compounds that are not anesthetics. This indicates that anesthetics do not act on a single type of nerve receptor but on many, which corresponds with the current view that the effect called “general anesthesia” is composed of several physiological and psychological effects, such as unconsciousness, amnesia, analgesia, loss of sensory processing, and reduction of motor reflexes. Each anesthetic compound induces a different constellation of effects, indicating that different nerve receptors are responsible for each effect. Likely targets are ion channels
Good News
279
(some mentioned in chapter 5), such as GABA receptors complexed with chloride ion channels, sodium ion/potassium ion channels, and potassium ion channels. The Love Test (1935) Production company: Fox-British Pictures Director: Michael Powell Screenwriter: Selwyn Jepson, based on a story by Jack Celestin Short summary: When Mary Lee becomes chief chemist, the other chemists conspire to make her fall in love so she will neglect her job and fail Plot description: Solutions bubble in chemical glassware behind the credits, which ends with a zoom onto a round-bottom flask that fades to become a man’s round face. It is the company president talking to his advisers about their celluloid improvement project. A quick camera pan shows the adjacent laboratory fi lled with men and women working at their benches. One of the workers lights a flame to a plastic cupie doll, which rapidly combusts. Everyone breaks for lunch except Mary Lee (Judy Gunn), who stays at her bench. We return to the president reading a letter from the Union Celluloid Company asking for quick delivery on the fi reproof celluloid. The president calls in Chief Chemist Mr. Hosiah Smith (Gilbert Davis), who hiccups with nervousness and says he’ll have to resign because the job is too much for him. He says Mary Lee should succeed him because she is the best chemist. Chemist Thompson (Louis Hayward) was listening at the door during the entire conversation. When he tells the other male chemists, John Gregg (Dave Hutcheson) says, “She’s a very good chemist,” to which Thompson responds, “I don’t want to take orders from a woman.” They continue to listen as Thompson devises a plan. One of them will have to cause her to neglect her duties by making her fall in love. To select the volunteer, each man picks up a test tube and empties its contents into a large beaker. “If it fi zzes, you’re elected.” Gregg is surreptitiously but deliberately chosen. Thompson keeps track of their relationship with a “Reaction Chart” and, after a slow start, stirs up the relationship by sending flowers and a love letter to Lee in Gregg’s name. Lee visits her neighbor to learn how to dress like a woman. The president’s secretary (Googie Withers) teaches Gregg how to kiss like you mean it. In a quick fade at 11:15, the bottles of (NH4)2SO4, Na2HPO4, and amyl alcohol in Gregg’s apartment turn into perfume bottles in Lee’s neighbor’s room. At 40:00, Lee becomes chief chemist and Thompson decides she is now attractive enough that he wants to pursue her. He shows her the “Reaction Chart,” claiming Gregg was playing a dirty game. There is a quick reaction shot of Lee standing outside her new office with “Chief Chemist” written on the glass door above her head.
280
ReAction! Chemistry in the Movies
Fire Proofing Celluloid Process -- Solutions KCN MgSO4 AgNO3 C.D. (NH4)Ox 3g Na2NH4(PO4) 6g KNO3 Na2B4O4 K4Fe(CN)6 2g
Figure 9.12. A portion of John Gregg’s celluloid fireproofing formula.
The next day, Thompson convinces everyone not to work on the formula, so the men distill booze and play solitaire while one woman knits. Gregg is the only one still working, and when he steps away from his bench, they sabotage his Bunsen burner so that it explodes upon his return. Lee rushes to help him, but he’s OK. When she thinks this has been a trick, she dismisses him. At home, Gregg empties his pockets in anger and tosses a kewpie doll into the fi replace. It doesn’t combust. That night, he sneaks into the laboratory to write the formula on a piece of paper (figure 9.12) and place it in Lee’s office along with a freshly dipped kewpie doll. Commentary: This is the earliest existing movie to feature a woman chemist. The fi lm posits a world in which there are equal numbers of men and women chemists, making it seem less unusual when a woman is selected to be chief chemist. Even though it is a light romantic comedy with characters who are chemists, it makes very creative use of chemical imagery. Nitrocellulose fi lm (figure 9.13) is notorious for its high flammability. This movie supposes that a coating of salts will render it fi reproof. In reality, safety fi lm involves a change in the chemistry of the fi lm itself. Photographic fi lm is a transparent strip coated with gelatin embedded O2NOH2C
*
O2NO
O ONO2
Cellulose Nitrate (or Nitrate Film)
AcOH2C O
* n
*
O
AcO
O
OAc
* n
Cellulose Triacetate (or Safety Film)
Figure 9.13. Nitrocellulose and cellulose acetate have very similar structures and many similar properties. The parentheses indicate that the modified glucose is a monomer in a repeating unit. The asterisks show the connection point to the next monomer within the cellulose polymer.
Good News
281
with silver salts for black-and-white images (or colored dyes for color images). When the fi lm is exposed to light, the salts (or dyes) precipitate to create a region of density. Processing the fi lm removes the unreacted salts (or dyes) and fi xatives hold the precipitates in place. Movie fi lm has to withstand high-speed and high-temperature fluctuations as it is projected past an intense light beam. When movie making was in its infancy, there was only one choice—nitrocellulose fi lm. Nitrocellulose fi lm (also called nitrate fi lm, or celluloid) was introduced by Eastman Kodak for still photography in 1899 (Slide 1992). It was used to make movies until the 1949, when Eastman Kodak introduced cellulose triacetate fi lm (also called safety fi lm or acetate fi lm). Overnight, safety fi lm became the industry standard and has been used in all Hollywood movies since 1950. Despite its desirable properties (hardness, flexibility, transparency), nitrocellulose is prone to spontaneous combustion and was linked to fi res in many theaters and fi lm storage vaults over the years. For instance, a spontaneous explosion at the Lubin Film Manufacturing Company in Philadelphia on June 13, 1914, destroyed its vault and contents. It is the reason that so few Lubin fi lms exist today. The earliest type of safety fi lm was developed by Kodak in 1909. It was a combination of mono- and diacetylcellulose. It was immediately adopted for home movies and almost immediately rejected by the movie studios. It was more brittle because of the plasticizers, allowing for fewer showings per fi lm. It was thicker than nitrocellulose to compensate for its brittleness, causing it to break more often because it didn’t fit as well in the projectors. All types of safety fi lm shrink and tear more easily than nitrocellulose, causing the image quality to decrease noticeably during every movie run. For instance, nitrate fi lm can be shown 200 times, whereas acetate can handle only 50 runs. The distributor has to make four times as many copies of an acetate fi lm. The problem for fi lm preservationists is that the half-life for triacetate fi lm is about 50 years, and possibly a bit longer for nitrocellulose (Smith 1981; Brems 1988; Image Permanence Institute 1991). The chemical issue is that the ester bond is not one of the strongest no matter whether it is formed with nitrate or acetate. Both are prone to hydrolysis in humid environments at high temperatures. Nitrocellulose fi lm releases NO2, HNO2, and HNO3, all of which are reactive gases and two of which are strong acids. These gases accelerate the decay of nearby fi lm and eventually reduce it to a highly explosive brown powder. Acetate fi lm decays to produce acetic acid (CH 3COOH) gas and liquid, which causes the fi lm to smell like vinegar and explains why it is called the “vinegar syndrome.” The best conditions for storage appear to be low humidity (no water for hydrolysis) at cool temperatures (slower reaction rate) in glass boxes (reasonably inert material). For this reason, many fi lm companies store their master reels in the former salt mines near Hutchison, Kansas, where the temperature is always 20°C (68°F) and the humidity always 50%.
282
ReAction! Chemistry in the Movies
The Schemers (1922) Production company: Reol Films Director: Wallace Johnson Screenwriter: Wallace Johnson Short summary: Research chemist Paul Jackson develops gasoline substitute, is kidnapped, and then thwarts criminals with the help of his girlfriend Isobel Benton Plot description: Industrial research chemist Paul Jackson (G. Edward Brown) develops a gasoline substitute. The company president’s secretary, Juan Bronson (Lawrence Chenault), and his associate Miguel Anderson (Walter Thomas) kidnap Jackson to get the formula but discover he isn’t carrying it with him. When Jackson calls his girlfriend Isobel Benton (Edna Morton) to tell her where to fi nd it in the lab, Anderson overhears the conversation. He heads out to get it, but she outwits him. After more sparring, the kidnappers are captured (figure 9.14). It seems they are known criminals, wanted by a South American government.
Figure 9.14. Lobby Card for THE SCHEMERS, described as “A thrilling dramatic photodrama with an appeal that knows no class or distinction.” Image courtesy of the Photographs and Prints Division, Schomburg Center for Research in Black Culture, The New York Public Library, Astor, Lenox and Tilden Foundations.
Good News
283
Commentary: This fi lm is probably the earliest fi lm to feature an African-American chemist and the only fi lm in this book considered to be “lost.” A fi lm is considered “lost” when it is not in the collection of any archive or prominent collector, in which case its narrative must be found in newspaper reviews or the trade magazines that catered to theater owners. The description above is based on the entry in the American Film Institute’s online catalog (www.afi.org). Reol Productions was one of about six companies making fi lms for African-American audiences during the late silent era (Sampson 1977; Butters 2002). They were formed to counter the offensive black stereotypes in many early fi lms, but especially from the highly influential The Birth of a Nation (1915), a sympathetic portrait of the Ku Klux Klan. Reol was based in New York City, used actors from a troupe called the Lafayette Players, and produced about 10 fi lms in 1921 and 1922. Their stars were Lawrence Chenault and Edna Morton, leaders of the Harlem Renaissance (Aberihani et al. 2003). In 2005, a set of eight lobby cards for The Schemers were donated to the New York Public Library as part of the Edna Morton Collection. A poster for another lost Reol fi lm, The Sport of the Gods, appeared on a U.S. Postal stamp in 2008 as part of its Black Cinema series.
10 First, Do No Harm (but Before That, Self-Experiment)
DRUG DISCOVERY, TARGET IDENTIFICATION, AND DRUG DEVELOPMENT “The physician must . . . have two special objects in view with regard to disease, namely, to do good or to do no harm.” Hippocrates (Of the Epidemics, 400 B.C.E.) (Adams 1891). This phrase from Hippocrates is more often quoted in its shortened version: “First, do no harm.” In the movies, this sentiment lies at the heart of the bioethical dilemmas in the drug discovery and development process (table 10.1). In horror movies, any step in the drug discovery process can go terribly wrong. This reveals public fears about human fallibility, or even malice, subverting even the best protocols. In the dramatic movies, each new drug or medical protocol is one more step into the bright light of a better future. The goal for these noble scientists is to reduce human suffering. The most common fi rst phase of drug discovery for compounds in this book has been the result of happy accidents and ethnobiological ventures. In chapter 5, the properties of LSD and Thorazine were both discovered while searching for other effects, although the discovery of these drugs has not been dramatized cinematically. Ethnobiology entered the picture in chapter 9 in the form of two movies from about 1990. Dr. Dennis Alan in The Serpent and the Rainbow searches for the zombie powder and discovers that it requires both puffer fish toxin and a cultural belief in zombies. Dr. Robert Campbell in Medicine Man searches for botanical pharmaceuticals and fi nds a cure for lymphoma. An ethical dilemma is presented in that movie with regard to who should benefit from the compound he discovers. Within the movie, we have to believe that no chemist would be able to synthesize the compound called “Mother Nature’s kitchen” and that Campbell is unable to replicate its isolation from the Amazonian flower. When one vial of the compound remains, a local boy gets cancer and Dr. Crane asks who is more important: one boy, or the rest of the world. Later, she chooses the boy. The movie does not give voice to the interesting question of the pharmaceutical company’s compensation to the locals’ discovery of the anticancer extract. Even though serendipity and ethnobiology still play important roles in discovering new chemicals 284
First, Do No Harm
285
*
√
√ *
*
Hollow Man (2000)
√ √
Nutty Professor (1998)
*
Kids in the Hall: Brain Candy (1996)
√
Nutty Professor (1963)
√
√
√
√
√
√
√ √
√ *
Monkey Business (1952)
Scanners (1980)
√ √ — √
Invisible Man (1933)
√ √ — * √
*
Strange Impersonation (1946)
√
√ * √
Jekyll and Hyde (1941)
√ √
√ * √
Jekyll and Hyde (1931)
√
The Great Moment (1944)
√
Lorenzo’s Oil (1992)
√
Dr. Ehrlich’s Magic Bullet (1940)
Medicine Man (1992)
Ethnobiology Target identification Known compounda Drug discovery and development Preclinical phase Self-testing Phase I: safety Phase II: effectiveness Phase III: large, ran domized studies Marketing and postproduction
The Serpent and The Rainbow (1988)
Table 10.1. Drug development phases shown in the movies
√
√
√
√ √
*
* √
√ * — — *
a “Known compound” means to extract or synthesize a known compound, as opposed to discovering or synthesizing a new compound. The use of a known compound implies that the compound is safe and that some of its favorable properties are already known, * Bioethical dilemmas presented; √, phase shown or discussed; —, phase was deliberately skipped by the characters in the movie, possibly because they self-experimented.
with new properties, it is just as important to modify existing chemical structures and to create novel structures by synthetic means. After you’ve collected or discovered therapeutic compounds, your next step is to identify the target protein or enzyme to which it binds in the cell. In real drug discovery, it is now more common to identify the target fi rst than to collect biologically active compounds fi rst. If you want to kill syphilis bacteria, you will identify essential proteins and enzymes in that bacteria and then isolate and purify them so you can search for small molecules that inhibit their activities. It is also true that if you don’t know the target, you really don’t know how the drug works. This esoteric fi rst phase is rarely shown in the movies, with the exception of Lorenzo’s Oil. Augusto Odone hypothesizes that he must inhibit the enzyme that transports very-long-chain fatty acids across a membrane. Since the enzyme appears to perform two functions, he has the idea to fi nd a small molecule that will interfere with one of them in the hopes it will also shut down the other function. Another movie that gets close to discussing the
286
ReAction! Chemistry in the Movies
enzyme target is Dr. Ehrlich’s Magic Bullet. Dr. Paul Ehrlich launched the field of modern drug discovery and formulated his “side chain theory.” He imagined that certain chemical structures were able to bind to a target in the syphilis bacteria to allow the arsenic to kill the bacteria. He also imagined that some of these chemical structures bound to a brain receptor to cause insanity in his test mice. He searched for the specific compounds that preferentially cured but didn’t cause insanity. Ehrlich was right that antibiotics either kill or arrest the growth of bacteria by entering the bacterial cell and binding to an essential enzyme in a way that inhibits it from carrying out its role. Known compounds are the starting point for the most therapeutic programs in the movies (table 10.1). In The Great Moment, it is actually a dilemma for W. T. G. Morton that the compound he uses as his inhalation anesthetic is so well known. He can make money by keeping its identity secret, but he reveals it when he realizes its incredible benefit to humanity. Two generations of moviemaking later, Scanners hypothesizes a secret government agency in which a known teratogenic compound is used to create an army of telekinetic people. A number of movies mention a drug development stage during which the physician or chemist carries out a series of tests. Three movies show this phase the best. In Hollow Man, Dr. Sebastian Caine adds a stabilizing bridge to his compound so that it can withstand serial irradiation. In Monkey Business, Dr. Barnaby Fulton heats his 58th formula to create the next version of his series. In Dr. Ehrlich’s Magic Bullet, 606 compounds were tested for their ability to bind and kill the syphilis bacteria and not to bind to the brain, where they can cause insanity. In real drug development, the relationship of the compound’s structure to several properties is extensively tested for: its adsorption into the body, its distribution throughout the body and to the target site, its speed of metabolism to waste products and the structures and properties of the waste products, the excretion of the waste products, and its toxicity. These are called the ADMET properties.
PRECLINICAL PHASE The use of animals as precursors to human experimentation has been a mainstay from the earliest guidelines until today (Altman 1987; Lederer 1995; Jonsen 1998). Even though cell cultures are replacing animals in preliminary tests, the animals are still required for defi nitive tests. Shortterm animal studies establish the safety and effectiveness of a treatment and are carried out before human studies begin. If a lead compound is safe and effective for many species, the rationale is that it must be safe and effective for humans. For this same reason, the effective and safe dose in animals is used to determine the starting dose in human trials. Some drugs and effects are common to the entire animal kingdom, such
First, Do No Harm
287
as general anesthesia. There are also compounds that have different effects in different animals, such as thalidomide (see Scanners, chapter 3; and Kids in the Hall: Brain Candy, chapter 4). Long-term animal studies begin before, but continue throughout, human trials. They are designed to identify reproductive problems and the effects of chronic use, both of which were required by the Kefauver-Harris Amendment of 1962, which was a direct response to the thalidomide problem. In fact, the amendment specifically called for reproductive tests in two animal species at two doses. Quite a number of movies show animal testing as a critical but mundane precursor to self or human experimentation. In The Great Moment, Dr. Morton anesthetizes the house pets, including three goldfi sh. Only then does he deliberately test it on himself. In the 1941 version of Dr. Jekyll and Mr. Hyde, the Hyde formula causes a rabbit to become vicious and a rat to become gentle. In this case, the compound transforms “good” or “evil” to its opposite. Jekyll knows he’s “good” so he also knows he will become “evil” when he takes the formula. In Monkey Business, an old and a young chimpanzee are kept in cages in the same laboratory where the chemicals are used. The roles of investigator and animal model are inverted, though, when Dr. Fulton drinks some of the chimp Esther’s effective formula. In The Serpent and the Rainbow, the zombie powder is used on a young goat as proof that it is effective. It was a ruse, and the goat dies. Later in the movie, Dr. Alan is deliberately zombified by someone with a grudge against him. This event and his subsequent burial alive don’t have a corollary in the book upon which the movie is based. In Hollow Man, invisibility and reversion formulas are tested on a menagerie of vertebrates before being tested on a human under false pretences.
SELF-EXPERIMENTATION Dr. Jekyll’s continued popularity ensures that all self-experimentation in the movies carries a tinge of the scientist out of control. Jekyll doesn’t have to ask anyone for permission to take his own medicine, with the result that society holds no sway over his actions. His story shows us, however, that Jekyll needs to reexamine his ethics. His Hyde has a negative impact on society, and Jekyll’s addiction to the formula ensures Hyde will be back. Today, self-experimentation is one of the great clichés. As an added homage to Frankenstein, it is often accompanied by a clap of thunder and a crack of lightning. A list of the real and fictional self-experiments described in this book (table 10.2) reveals that Jekyll appeared just as medical research was becoming more scientific, although not yet including drug discovery. Until well into the 1900s, physicians would use complex treatments proven to work through repeated experience rather than by scientific testing. In the mid-1800s, pure compounds with reliable efficacy were
288
ReAction! Chemistry in the Movies
Table 10.2. Self-experimentation relating to drug discovery mentioned in this book Researcher
Substance
Effect
Year
Morton Mantegazza Richardson Freud Koller Jekyll Morgenrath Hofmann Goodrich Fulton Jessup
Diethyl ether Coca Mandrake extract Cocaine Cocaine Hyde Formula Compound 606 LSD Unnamed Formula X58 Atropine-like
General anesthetic Psychomotor General anesthetic Psychomotor Local anesthetic Personality separator Antisyphilitic Hallucinogen General anesthetic Anti-arthritic Hallucinogen
1846 1859 1874 1886 1886 1886 1909 1943 1946 1951 1980
The italicized entries indicate fictional events. Only the original Jekyll appearance is listed.
just becoming available. In the modern period of drug discovery (1909 to 1960s), chemists would often use their own sensations to know they had purified the active ingredients from plants, animals, or minerals. This tradition remains strongest in the area of general and local anesthetics, where the chance of taking a toxic dose is low. Since all animals respond to anesthetics at similar doses, their dose is nearly perfectly scalable. Immunology and bacteriology are other fields in which selfexperimentation was common (Beecher 1970; Altman 1987; Lederer 1995; Jonsen 1998). In fact, one could argue that self-motivated, selfexperimenting physicians are most able to make accurate and valuable observations about their experience. For instance, in 1767, physician John Hunter inoculated himself with a patient’s gonorrheal pus to prove the disease was transmissible. It was—he contracted the somewhat treatable gonorrhea but also the less easily treated syphilis (he didn’t realize the patient had both). Robert Koch pushed bacteriology in a more scientific direction when he published his bacterial theory of disease in 1877. According to his postulates, it was necessary to isolate the bacteria in culture and then use it to infect a test subject to prove that the symptoms were replicated. This was fi ne when studying animal diseases, but research soon turned to human diseases that lacked good animal models. Bacteriologists apparently tired of self-experimenting, and some performed human experiments without the subject’s consent, with two cases cited most frequently (Lederer 1995; Jonsen 1998). In 1897, the Italian researcher Sanarelli injected bacilli into five unknowing subjects, who then contracted “typical yellow fever” that he described in a publication. In 1898, German researcher Neisser infected four unknowing prostitutes with Treponema pallidum bacteria, and they contracted syphilis that he described in a publication. The response to Neisser’s research was swift and led to the fi rst
First, Do No Harm
289
Figure 10.1. Dr. Paul Ehrlich (Edward G. Robinson, farthest left) summarizes the team’s successful animal experiments and says the next step is to test in humans. Soon after this scene, young Morgenrath (farthest right) performs the first self-experiment. Doctor Ehrlich’s Magic Bullet © Turner Entertainment Co. A Warner Bros. Entertainment Company. All Rights Reserved. Collection of the authors.
human testing standards by the Prussian Ministry of Health in 1900. These standards were further strengthened by the Weimar Republic in 1931. Among the requirements were that animals must be used fi rst and all patients must consent. German health care, treatment, and medical science were the best in the world in the 1920s and 1930s. Dr. Paul Ehrlich carried out the fi rst modern chemotherapeutic drug development program, which ended with a chemical cure for syphilis in 1909 (see chapter 9). In Dr. Ehrlich’s Magic Bullet, self-experimentation is presented as being entirely noble (figure 10.1). When young Morgenrath realizes that Ehrlich intends to infect himself with syphilis and then the toxic drug, he decides to carry out the self-experiment fi rst. Luckily for him, it worked. A great deal was written about the ethics of human experimentation in the period between 1947 and the 1980s. This period began with the Nuremberg trials and was transformed in the United States by the National Research Act of 1974. During WWII, German physicians and nonphysicians carried out unethical and unscientific procedures they called “experiments” on political prisoners. The evidence was documented
290
ReAction! Chemistry in the Movies
in the Nuremberg trials of 1947 and resulted in the Nuremberg Code, an attempt to articulate the ethics of human experimentation. Article 5 states, “No experiment should be conducted where there is a priori reason to believe that death or disabling injury will occur; except, perhaps, in those experiments where the experimental physicians also serve as subjects.” In other words, physician self-experimentation was used as a smell test: If it is too dangerous for the doctor, it is too dangerous for the patient. This article of the Nuremberg Code is problematic, however, in that unstable physicians might be willing to perform a dangerous experiment on themselves and, if they survive, be willing to perform the same procedure on others (Beecher 1970). Today, it is clear that public oversight of biomedical research by institutional review boards (IRBs) is a much better way of handling the ethics of human experimentation. The inclusion of physicians in their own clinical trials is now considered to be a source of undesirable bias.
HUMAN EXPERIMENTATION AND THE INSTITUTIONAL REVIEW BOARD The drugs that reach the market today are safer and more effective than any in the past and have passed the strongest ethical scrutiny by a public body. These achievements are due to rigorous double-blind clinical trials and IRBs that became the standard in the 1960s and 1970s. Before then, it was common in the Unites States and around the world for medical researchers to test new protocols and drugs on patient volunteers and also on military personnel, prisoners, mental patients, and aged people who were about to die in the hospital (Lederer 1995). The use of these subjects became a concern in the 1960s (Beecher 1966), two decades after medical research exploded in the aftermath of WWII. There was a desire among the researchers to help society by eliminating its diseases, but there were too many cases where the rights of some patients were ignored. In many documented cases, groups of biomedical researchers were the ones who balanced a given participant’s perceived contribution to society with the supposed benefit to society if that participant became an unknowing test subject. The competing roles of researcher/participant and physician/patient were clarified in the Belmont Report of 1979, possibly the most important document in bioethics (Jonsen et al. 1998). This is the same document that also provided guidance for creating informed consent forms. It was the last of a series of important documents created by the National Commission for the Protection of Human Subjects of Biomedical and Behavioral Research, composed of 12 commissioners who were scientists, lawyers, ethicists, and citizens. The commission was formed by the National Research Act of 1974, which Congress was stimulated to pass when the Washington Post broke a story in 1972 involving a 40-year,
First, Do No Harm
291
government-sponsored research into the effects of untreated syphilis (Heller 1972). The Tuskegee Syphilis Study began in 1932 in rural Macon County, Georgia, where African-American men were routinely denied simple cures for syphilis. The project eventually involved about 400 men, none of whom ever knew they were part of a study. The revelation of this unethical study during the most confrontational period of the civil rights movement led to a rather speedy response by the government. The result was a set of rules for human experimentation and a full understanding that social progress cannot be achieved at the expense of any individual’s human rights. The commission also required that all biomedical research be subjected to IRBs. IRBs had been created in 1966 by the National Institutes of Health (NIH) to provide independent, on-site review of the ethical portions of human clinical experiments that they funded. IRBs are composed of at least five individuals that include scientists, nurses, social workers, ethicists, and at least one lay person. This meant that nonphysicians would now scrutinize biomedical research projects and pay attention to the welfare and rights of its participants. Biomedical researchers design their proposals and carry out their research with this social constraint in mind. NIH created the IRBs for a variety of reasons, including the 1964 adoption of the Nuremberg Code by the World Medical Association. There was general agreement that ethical research required the consent of the subject, the prerequisite of animal experiments, and proper medical supervision. British statistician Austin Bradford Hill created the randomized (also called double-arm) clinical trial to eliminate bias (Bradford Hill 1951; Bowden et al. 2003). He published an important summary of his method in 1963 in which he emphasized its ethical dimensions (Bradford Hill 1963). The patients are split into two groups: One receives the new treatment, and the other receives a placebo or the common treatment. In double-blind trials, neither the physician nor the patient knows who is receiving which treatment, which eliminates subtle or unconscious biases on the part of the physician because every patient receives the best interim care regardless of their treatment. A placebo is only given when it would not cause harm. If it becomes clear during a study that one group of patients is doing particularly well or not, the project can be terminated. In addition, any participant may withdraw from the study at any time. Single-arm studies are used when there are small numbers of participants, such as for rare and debilitating diseases like adrenoleukodystrophy (ALD). In the recently completed project involving 89 asymptomatic ALD patients (Moser et al. 2005), all participants received Lorenzo’s Oil because there is no ethically suitable alternate treatment. Bradford Hill performed the fi rst randomized clinical trials in 1946 for a whooping cough vaccine and streptomycin antibiotic effectiveness, both of which involved large numbers of individuals. Bradford Hill continued to develop his methods throughout the
292
ReAction! Chemistry in the Movies
1950s, and even though their adoption was slow in coming, his methods became the gold standard for clinical trials in 1962 when Congress passed the Kefauver-Harris Amendment specifying that federal drug approval required “adequate and well-controlled investigations, including clinical investigations.”
HUMAN CLINICAL TRIALS: PHASES I, II, III, AND IV Today’s clinical trials consist of four phases, each created by law during the past century. Phase I is a small-scale human safety trial. This requirement was created by the Food, Drug, and Cosmetics Act of 1938. Prior to that, it was only necessary to prove purity and to have all ingredients listed (Pure Food and Drug Act of 1906). Today, phase I trials involve 20–80 healthy volunteers and establish the safety of various doses of the proposed new drug. Phase II involves a medium-scale test for effectiveness on patient volunteers. Phases II and III were added by the Kefauver-Harris Amendment of 1962. The safety, effectiveness, and side effects of several doses are determined during this phase on 100–300 patients with the disease in question. In Dr. Ehrlich’s Magic Bullet, the protagonist and Behring agree to a test in which 20 hospitalized children dying from diphtheria will receive the diphtheria serum and 20 will not. After Ehrlich fi nishes injecting the twentieth child, he decides to give his compound to the remaining 20 children over the strong objections of the hospital chief. Ehrlich feels it would be unethical to withhold the serum even if it only had a remote chance to cure them. Phase III is a large-scale test or tests to confi rm drug effectiveness and to monitor adverse reactions during long-term treatment. This phase involves 1,000–3,000 patient volunteers at multiple locations. If this phase is successful, the FDA can approve the compound for the purpose it was developed, and the compound is now properly called a “drug.” This phase is not represented in the movies even though there are several ways to inject drama into confi rmatory tests. In a step sometimes called phase IV, pharmaceutical companies are required to collect information about problems that are encountered for many years after the drug’s release to the public.
MARKETING AND POSTPRODUCTION PHASES During the marketing and postproduction phase, physicians or the public are informed about the new drug’s curative properties (figure 10.2). In The Great Moment set in 1840, the marketing phase for Letheon is very similar to Snake Oil hucksterism. In fact, it wasn’t until the 1906
Figure 10.2. “Mileposts in Local Anesthesia: Monocaine” 1946 advertisement by the Novocol Chemical Manufacturing Co., of Brooklyn, NY, in the Journal of the American Medical Association. Collection of the authors.
294
ReAction! Chemistry in the Movies
Pure Food and Drug Act that manufacturers were required to list the ingredients in their products. The Food, Drug, and Cosmetics Act of 1938 strengthened that provision by allowing prosecution of false claims and by requiring over-the-counter drugs to include warning labels and directions for use. Most recently, the 1962 Kefauver-Harris Amendment required both risks and benefits to be stated in any promotion.
Conclusion Chemistry in the Movies
The authors’ common reaction to chemistry in the movies is encapsulated in the archetype movies (see table 11.1). These are, fi rst and foremost, great movies that present certain facets of chemistry especially well. They were selected from a much larger group of movies by ranking according to four criteria: (1) contemporary (meaning released after 1970), (2) available on VHS or DVD, (3) included women or other underrepresented groups in significant roles, or (4) was especially favored by one or both of the authors. It became clear from the ranking exercise that older fi lms overcame the criterion of not being recent when they were favored by both authors. We felt they represented the archetype for that chapter and merited special attention. The oldest archetype movie is the 1931 Dr. Jekyll and Mr. Hyde, making it the book’s de facto archetype and reiterating its importance as the book’s overarching theme. Considered as a whole, the five chapters on the “dark side” show chemists, sociopaths, chemical companies, and pleasure seekers making one-sided decisions that ultimately harm themselves and society. After Jekyll becomes addicted to his Hyde formula, he commits acts of personal terrorism and then murder. Griffi n works alone to isolate his invisibility formula because he seeks fame, wealth, and power. Once he knows those things are within his grasp, it drives him mad to the point that he commits mass murder. Dr. Mabuse isn’t a chemist, but he is already insane when he commands his army of thugs to engage in acts of chemical sabotage. He wants to begin a “reign of terror.” Reporter Jason Brady learns that a president knows his chemical company produces a toxin that kills his workers and the children living near the plant. He won’t stop production because it would deprive the community of employment. Finally, television director Paul Groves takes his fi rst LSD trip to get in touch with his feelings. While under the influence, he flees the apartment of a guide who was there to ensure he had a good experience. “Bright side” chemists usually work in teams and rely on other people for critical input—they are engaged with society. The oldest archetype in the bright second half of the book is Dr. Ehrlich’s Magic Bullet from 1940. Ehrlich’s altruistic desire to reduce human suffering without being judgmental is noble. The movie is also important because it essentially serves as the archetype for chapters 9 and 10. Paul Ehrlich made seminal 295
296
ReAction! Chemistry in the Movies
Table 11.1. Archetypical chemistry in the movies Chapter and Title (Year)
Chemical Theme
1. Dr. Jekyll and Mr. Hyde (1931) 2. The Invisible Man (1933) 3. The Testament of Dr. Mabuse (1933) 3. Dr. Strangelove (1964) 4. One Man (1977) 5. The Trip (1967) 6. The Man in the White Suit (1951) 7. Kid Glove Killer (1942) 8. The Nutty Professor (1963) 9 and 10. Dr. Ehrlich’s Magic Bullet (1940)
Chirality and mirrors Cocaine local anesthesia Nerve gas Fluoridation paranoia Lead poisoning LSD and Thorazine Synthetic fiber Vanadium trace detection Androgenic-anabolic steroids Arsenic-containing antisyphilitic
contributions to diagnostic cell staining and immunology before he led his team to create the field of modern drug discovery. Julius Kelp in the 1963 Nutty Professor relies upon theory gathered from reading books to solve his personal problems, but not in a way that harms society. As an instructor, he is educating students for the future. Forensic chemist Jane Mitchell knows how to use a spectrograph to assess clues. She and her partner discover and connect as many clues as possible to fi nd a killer. She is the only woman chemist among the archetypes. It is interesting to note that most prominent women’s roles fall on the bright side. The few found in the fi rst five chapters work against the dark elements such as bad chemical companies, which places their characters on the bright side. Two of these portrayals earned Academy Awards (Rachel Weisz and Julia Roberts). Finally, we have Sidney Stratton, whose fiber can’t be stained, dirtied, or torn. As an inventor chemist, he has a goal in mind but doesn’t know how to reach it, so he uses trial and error to discover the right set of synthetic conditions. These 10 archetype fi lms distill the essence of chemistry in the movies.
Appendix 1 How to Use This Material in the Classroom
Chemistry instructors can use either entire movies or movie clips as part of their lecture strategy. Only a small subset of the movies is best suited for viewing in their entirety, as described in the next section. On the other hand, nearly all movies in this book have short “scientific explanation scenes” within their narratives that can be used in the chemical classroom to illustrate a chemical point or provoke a discussion. These 3- to 5-minute movie clips can be used for all the same reasons as lecture demonstrations.
USING ENTIRE MOVIES IN THE CLASSROOM OR LABORATORY The following 11 movies are suitable for use in their entirety. These movies are based on true chemical stories: Dr. Ehrlich’s Magic Bullet (1940), Edison, the Man (1940), Madame Curie (1943), The Great Moment (1944), Silkwood (1983), Lorenzo’s Oil (1992), Apollo 13 (1995), Me and Isaac Newton (1999), An Inconvenient Truth (2006), and Who Killed the Electric Car? (2006). Two of these movies, Apollo 13 and An Inconvenient Truth, have not been described in this book. Even though they contain teachable chemical moments, they don’t fall as easily into the book’s ten major themes (in Apollo 13, the 5.5-minute scene starting at 1:27:00 starts when the CO2 lamp illuminates to cause all attention to focus on the CO2-scrubbing LiOH canisters and ends when the lamp turns off; the fi rst half of An Inconvenient Truth provides a highly visual summary of global warming). In all cases, instructors should watch the selected movie before using it in the classroom. You must judge for yourself whether and when to show it to your students. The story line for each movie can be found on the Internet Movie Database (www.imdb.com). Since it is highly desirable for students to learn how to write about chemical topics (Beall 1993; Herreid 1994, 2004; Koprowski 1997; Kovac and Sherwood 1999; Jones and Miller 2001; Shibley et al. 2001; Cornely 2003), and because the pedagogical utility of movies to teach science has been noted (Dubeck et al. 1988, 2004; Resch and Schicker 1992; Borgwald and Schreiner 1994; Wink 2001; Rose 2003), it is only natural to combine these two ideas to use movies based on true chemical stories 297
298
Appendices
as the engaging source material for such writing assignments. Griep used this approach with great success in his general chemistry course, and our description of the project was featured in the Journal of Chemical Education (Griep and Mikasen 2005). The essence of the project involved choosing two fi lms from the list in the previous paragraph (Griep chose Dr. Ehrlich’s Magic Bullet and Me and Isaac Newton), arrange for rooms and times so that large numbers of students could view the movies outside the classroom, and then set a deadline for completion of the writing project. The students were required to write a 600-word report about the chemistry in it, to use citations (at least one of which is not from the Internet), and to include an image of the molecule described or implied in the movie. Reports were scored using a standardized scoring sheet, also described in the article, that the students received before they began writing. The most important justification for including a laboratory component with undergraduate chemistry courses is that it provides hands-on experience. The argument is that students can’t possibly understand matter and its transformation unless they carry out the manipulations themselves. As one laboratory exercise out of 12 or 13, it is also reasonable to include one in which the students watch a movie based on a true chemical story. In the movies listed above, the students learn that chemists encounter many experimental and nonexperimental hurdles in their lives, that they are human, and that they are motivated by a desire to do good. As with all laboratory exercises, the students answer a series of questions (12–15 in this case), write a short essay in response to one question, and fi nd the structure of the molecule mentioned in the movie. Griep has had success with this approach in his liberal arts chemistry course using The Great Moment (1944) about the discovery of ether anesthesia and Dr. Ehrlich’s Magic Bullet (1940) about the development of synthetic chemical therapies such as antibiotics.
USING MOVIE CLIPS IN THE CLASSROOM The location and duration for the most pedagogically useful clips within each movie are provided in each movie description in chapters 1–10. Many of these clips relate to material presented in an introductory chemistry course. The simplest way to use the clips is to purchase a new or used VHS copy, cue up the desired scene, and then show it during class linked to a normal lecture topic. (DVDs are problematic with regard to cueing.) Prior to showing the clip, it is desirable to describe the characters the students are about to see, although this is not necessary for those scenes in which the characters describe their motivation. After showing the clip, it is essential to describe the chemistry in the clip and the manner in which it relates to real chemistry.
Appendices
299
Movie clips are useful in the chemistry classroom for the same reasons as the perennially popular lecture demonstrations. The justifications given for lecture demonstrations are that they (1) teach chemistry, (2) are fun to do, (3) are enjoyed by students, (4) grab students’ attention, and (5) provide concrete examples of abstract concepts (Bodner 2001). In fact, Carleton College professor Richard Ramette suggested that demonstrations are “exocharmic” (Ramette 1980) when they are kinetically and thermodynamically favorable. In common language, that means they “exude charm.” As should be obvious from the movies described in this book, exocharmic reactions are also favored by moviemakers, but with the addition of an encompassing plot. It seems appropriate to end with a quote from Hubert Alyea, the inspiration for Disney’s The Absent-Minded Professor. In 1954 he wrote, “Surprise, humor, and truth are the servants of a good lecturer” (Alyea 1955; emphasis original). Alyea was a master of these elements and, in 1955 and 1956, published a series of 24 articles describing low-cost and effective demonstrations in the Journal of Chemical Education. It was compiled and reprinted as a book in 1957 and continuously amended, printed, and reprinted by co-compiler Frederic Dutton until at least 1969 (Alyea 1957; Alyea and Dutton 1969). Several generations of chemistry instructors used these tersely worded descriptions to great effect. On the other hand, today’s reader will be struck by Alyea’s lack of explanations and the inadequate safety “cautions.” These omissions were both corrected in Shakhashiri’s later compilations of demonstrations (Shakhashiri 1983–1989).
This page intentionally left blank
Appendix 2 About the Back Cover Art
Marjorie Mikasen, artist and co-author, created the painting Jekyll & Hyde especially for this book (see color image on back cover and black and white image on title page). It is not only a two-dimensional painting; it is also a stereopair painting. The design comprises two side-by-side images that can be optically converged for a three-dimensional effect. Since twosidedness is the most salient feature of the Jekyll and Hyde story, stereo treatment provides a fitting conceptual approach. The significant molecule in the story, ergotamine (see chapter 1), is depicted here as a stick model—itself a stereopair. This type of image has a long association with chemistry. Often found in textbooks, stereopairs convey three-dimensional structural information about molecules and proteins. Sometimes small hand-held viewers are included with the textbooks, enabling readers to see the images in stereo. Due to the large number of stereopairs chemists encounter, however, many dispense with the viewers and train themselves to see the images with free vision, or cross-eyed vision, using no mechanical aids. This is the intended viewing method for Jekyll & Hyde. Conceived with an understanding of the function of convergence and disparity in the vision process, free vision uses the knowledge that we focus where lines of sight cross. Crossing the eyes to view the images is just doing intentionally what the eyes do naturally. In addition to providing depth cues, the stereo technique is used to explore visual syncopation, optical color mixing, and kinetic effects. The suggestion of being at the movies is fundamental to Jekyll & Hyde’s design. The fi lmstrip elements run vertically on both sides and in the center. Their oval perforations alternate upward, evoking the ebullition of Dr. Jekyll’s effervescent formula. When optically converged, these fi lmstrip elements rhythmically vibrate as though they are running through a fi lm projector. Jekyll & Hyde’s design also features a precise set of colors. They correspond in hue and saturation to the color changes occurring in Dr. Jekyll’s formula (see chapter 1). When colors are “mixed” by stereo viewing, a color quality like watery green can be perceived especially well. And where would a painting about Dr. Jekyll be without a mirror symbol? In the upper left corner of each segment is a black-and-white square split on the diagonal. The architectonic, abstract figural form is a signature motif used by Mikasen in previous stereopair paintings. When
301
302
Appendices
the diagonally divided squares optically overlap, white meets black and black meets white, creating a shimmering effect that is reminiscent of a shining mirror. Finally, Jekyll & Hyde has an added dimension. Instead of a single focal point to the stereopair, there are two focal points that mirror the divided Jekyll and Hyde character. When viewing it with one of the focal points, the ergotamine molecule structure comes into three-dimensional focus while other elements shift and split. Viewing with the other focal point, Dr. Jekyll’s “mirror” comes into focus and the ergotamine molecule structure begins to pull apart. This visually encapsulates the problem of the character’s oppositional personalities never fitting together easily. A seamless, easy integration is not possible.
References
Introduction Bronowski J (1965) Science and Human Values. New York: Perennial Library/ Harper and Row. Emsley B (2006, May 26) “Jekyll and Hyde Spooky? A Real Nightmare Is a World with No Chemistry!” [press release]. Royal Society of Chemistry. Available: www.rsc.org/AboutUs/News/PressReleases/2006/JekyllandHyde.asp. Indick W (2004) Psychology for Screenwriters. Studio City, CA: Michael Wiese Productions. Jung CG (1970) Collected Works, Vol. 10. Civilization in Transition. Princeton, NJ: Princeton University Press. Jung CG (1980) Collected Works, Vol. 9. Part 1: The Archetypes and the Collective Unconscious. Princeton, NJ: Princeton University Press. National Academy of Sciences (1995) On Being a Scientist: Responsible Conduct in Research. National Academy of Sciences; National Academy of Engineering; Institute of Medicine. Available: www.nap.edu/openbook.php? record_id=4917. National Science Foundation (2006) “Doctorates in the 20th Century.” NSF Report 06–319. Washington, DC: National Science Foundation. Tudor A (1989) “Seeing the Worst Side of Science.” Nature 340: 589–592. Zurer P (1998, November 30) “Science Film Festival Capitalizes on ‘Teachable Moments.’ ” Chemical & Engineering News, 30.
Chapter 1 Avens R (1977) “The Image of the Devil in C. G. Jung’s Psychology.” Journal of Religion and Health 16: 196–222. Berg JM, Tymoczko SL, and Stryer L (2007) Biochemistry. New York: Freeman. Bernal I, Hamilton WC, and Ricci JS (1972) Symmetry: A Stereoscopic Guide for Chemists. New York: Freeman. Bonjean J (1842) Histoire Physiologique, Chimique, Toxicologique et Medicale du Seigle Ergote [Physiological History, Chemistry, Toxicology and Medicine of Rye Ergot]. Paris: Crochard. Booth BA, and Mehew E, Eds. (1994–1995) The Letters of Robert Louis Stevenson. New Haven, CT: Yale University Press. Brown TL (2003) Making Truth: Metaphor in Science. Urbana: University of Illinois Press. Close F (2000) Lucifer’s Legacy: The Meaning of Asymmetry. Oxford: Oxford University Press.
303
304
References
Dixon WW (1991) The Charm of Evil: The Life and Films of Terence Fisher. Metuchen, NH: Scarecrow Press. Dury R (2006) “Film Versions of ‘Dr Jekyll and Mr Hyde.’ ” Available: dinamic02. unibg.it/rls/fi lms-jh.htm. Faraday M (1857) “The Bakerian Lecture: Experimental Relations of Gold (and Other Metals) to Light.” Philosophical Transactions of the Royal Society of London 147: 145–181. Forepaugh L, and Fish GF (1897) Dr. Jekyll and Mr. Hyde, Or, A Mis-spent Life. New York: Samuel French. Goodwin K (2005, March 20) “Drug Took Stevenson Face to Face with Hyde.” Times (London). Available: www.timesonline.co.uk/tol/news/uk/scotland/ article432698.ece. Guttmacher AE, and Callahan JR (2000) “Did Robert Louis Stevenson Have Hereditary Hemorrhagic Telangiectasia?” American Journal of Medical Genetics 91: 62–65. Hannah B (1971) Striving Towards Wholeness. New York: Putnam Press. Haynes RD (1994) From Faust to Strangelove: Representations of the Scientist in Western Literature. Baltimore, MD: Johns Hopkins University Press. Houghton W (1957) The Victorian Frame of Mind. New Haven, CT: Yale University Press. Hu M, Chen J, Li Z-Y, Au L, Hartland GV, Li X, Marquez M, and Xia Y (2006) “Gold Nanostructures: Engineering Their Plasmonic Properties for Biomedical Applications.” Chemical Society Reviews 35: 1084–1094. Huxtable RJ, and Schwarz SKW (2001) “The Isolation of Morphine: First Principles in Science and Ethics.” American Society of Pharmacology and Experimental Therapy 1: 189–191. Klepper RK (1999) Silent Films, 1877–1996: A Critical Guide to 646 Movies. Jefferson, NC: McFarland. Reid J (2006) Robert Louis Stevenson, Science, and the Fin de Siecle. Houndmills, Hampshire, UK: Palgrave Macmillan. Robinson D (2007) Consciousness and Its Implications, Lecture Two: Self Consciousness [Course Guidebook]. Chantilly, VA: Teaching Company. Rose BA (1996) Jekyll and Hyde Adapted: Dramatizations of Cultural Anxiety. Westport, CT: Greenwood Press. Rylance R (2000) Victorian Psychology and British Culture, 1850–1880. Oxford: Oxford University Press. Salman S (1997) “The Creative Psyche: Jung’s Major Contributions” in The Cambridge Companion to Jung, P Young-Eisendrath and T Dawson, Eds. Cambridge: Cambridge University Press. Schardl CL, Panaccione DG, and Tudzynski P (2006) “Ergot Alkaloids: Biology and Molecular Biology.” The Alkaloids 63: 45–86. Sertürner F (1817) “Über das Morphium, eine Neue Salzfaehige Grundlage, und die Makonsaeure, als Hauptbestandtheile des Opiums” [On Morphine, a New Salt-Forming Base, and Makon Acids, as the Main Therapeutic Components of Opium]. Annalen der Physik 55: 56–89. Shubnikov AV, and Koptsik VA (1974) Symmetry in Science and Art. New York: Plenum Press. Skal DJ (1998) Screams of Reason: Mad Science and Modern Culture. New York: Norton.
References
305
Stachel SJ, Habeeb RL, and Van Vranken DL (1996) “Formation of Constrained, Fluorescent Peptides via Tryptophan Dimerization and Oxidation.” Journal of the American Chemical Society 118: 1225–1226. Stevenson F (1885) “Letter from Fanny Stevenson to William Henley.” Yale Archives, New Haven, CT. Stoll A (1920) “Zur Kenntins der Mutterkorn Alkaloide” [On the Isolation of the Rye Ergot Alkaloid]. Verhand der Schweiz Naturs Gesselschaft 6: 190–200. Tanret C (1875) “Sur la Presence d’une Nouvelle Alkaloide, L’Ergotinine, Dans Seigle Ergote” [Concerning the Presence of a Novel Alkaloid Ergotine In Rye Ergot]. Comptes Rendus des Seances de L’Academie des Sciences [Reports of the Meetings of the Academy of Science] 81: 896–897. Thanhouser Company (1912) “Dr. Jekyll and Mr. Hyde” [Advertisement]. The Moving Picture World 33. Thanhouser Company Film Preservation Inc. (1997) “Dr. Jekyll and Mr. Hyde (1912).” Available: www.thanhouser.org/fi lms/jekyll.htm. Tudzynski P, Correia T, and Keller U (2001) “Biotechnology and Genetics of Ergot Alkaloids.” Applied Microbiology and Biotechnology 57: 593–605. Tweney RD (2006) “Discovering Discovery: How Faraday Found the First Metallic Colloid.” Perspectives on Science 14: 97–121. U.S. Pharmacopeia (2006) “About USP—An Overview.” Available: www.usp. org/aboutUSP/index.html. Whyte LL (1975) “Chirality.” Leonardo 8: 245–248.
Chapter 2 Anonymous (1979) “Cocaine.” British Medical Journal 1: 971–972. Armstrong Davison MH (1965) “The Evolution of Anaesthesia” in The Evolution of Anaesthesia. Baltimore, MD: William and Wilkins, 171–177. Becker HK (1963) “Carl Koller and Cocaine.” Psychoanalytic Quarterly 32: 309–373. Brain PF, and Coward GA (1989) “A Review of the History, Actions, and Legitimate Uses of Cocaine.” Journal of Substance Abuse 1: 431–451. Brandt R (2007) “What Can Neuroscience Tell Us about Evil?” Technology Review. Available: technologyreview.com/Biotech/18573/. Braun H (1905) “Über einige neue örtliche anaesthetica (Stovain, Alypin, Novocain)” [On a New Local Anesthetic]. Deutsche Medizine Wochenschrift [German Medical Weekly] 31: 1667–1671. Calatayud J, and Gonzalez A (2003) “History of the Development and Evolution of Local Anesthesia since the Coca Leaf.” Anesthesiology 98: 1503–1508, 100: 1321–1323. Curtis J (1982) James Whale. Metuchen, NJ: Scarecrow Press. Einhorn A, and Uhlfelder E (1909) “Über den p-Aminobenoesaurediathylaminound -Piperidoathylester.” Justis Liebig’s Annalen der Chemie 371: 131–142. Freud S (1884) “Ueber Coca” [On Coca]. Zentralblatt fuer die Gesammte Therapie [Central Journal for Therapy] 2: 289–314 (translated by S. Pollak [1884], St. Louis Medical and Surgical Journal 47: 502–505). Freud S (1957) “On Narcissim (1914)” in The Complete Psychological Works of Sigmund Freud, Standard Edition, Vol. 4. London: Hogarth Press, 73–102.
306
References
Gatiss M (1995) James Whale: A Biography; Or The Would-be Gentleman. London: Cassell. Hadda SE (1964) “Procaine: Alfred Einhorn’s Ideal Substitute for Cocaine.” Journal of the American Dental Association 64: 841–845. Hoffman M (2000) Empathy and Moral Development: Implications for Caring and Justice. Cambridge: Cambridge University Press. Ingelman-Sundberg M, Oscarson M, and McLellan RA (1999) “Polymorphic Human Cytochrome P450 Enzymes: An Opportunity for Individualized Drug Treatment.” Trends in Pharmacological Science 20: 342–349. Knapp H (1884) “On Cocaine and Its Use in Ophthalmic and General Surgery.” Archives of Ophthalmology 13: 3–4. Knight Z (2006) “Some Thoughts on the Psychological Roots of the Behavior of Serial Killers as Narcissists: An Object Relations Perspective.” Social Behavior and Personality 34: 1189–1206. Koller C (1884a) “On the Use of Cocaine for Producing Anaesthesia on the Eye.” Lancet 2: 990–992. Koller C (1884b) “Über die Verwendung des Cocaien zur Anaesthesirung am Auge” [On the Use of Cocaine for Producing Anaesthesia on the Eye]. Wiener Medizinische Wochenschrift [Viennese Weekly Medical Journal] 34: 1276– 1278, 1309–1311. Koller C (1928) “Historical Notes on the Beginning of Local Anesthesia.” Journal of the American Medical Association 90: 1742–1743. Kuehl P, Zhang J, and Lin Y (2001) “Sequence Diversity in CYP3A Promoters and Characterization of the Genetic Basis for Polymorphic CYP3A5 Expression.” Nature Genetics 27: 383–391. Lezec HJ, Dionne JA, and Atwater HA (2007) “Negative Refraction at Visible Frequencies.” Science 316: 430–432. Liljestrand G (1971) “The Historical Development of Local Anesthesia” in International Encyclopedia of Pharmacology and Therapeutics: Section 8: Local Anesthetics, P Lechat. Oxford: Pergamon Press. Löfgren N (1948) Studies on Local Anesthetics: Xylocaine. Stockholm: Hoeggstroms. Lossen W (1865) “Ueber das Cocain” [On Cocaine]. Annalen der Chemie und Pharmacie [Annals of Chemistry and Pharmacy] 133: 351–371. Mantegazza P (1859) Sulle Virtu Igieneche e Medicinale della Coca, a Sugli Alimenti Nervosi in Generale [On the Hygienic and Medicinal Virtues of Coca, and about Nervous Digestion in General]. Milan. Mariani A (1886) Coca Erythroxylon, Vin Mariani, Its Uses in the Treatment of Disease. New York: Mariani. Mariani A (1890) Coca and Its Therapeutic Application. New York: Jaros, J.N. Mariani A (1894–1910) Figures Contemporaines. Paris: Henri Floury. Mariani A (1902) Eminent Physicians’ Portraits. Paris: Mariani. Mayer E (1924) “The Toxic Effects Following the Use of Local Anesthetics.” Journal of the American Medical Association 82: 876–885. Meikle D (2003) Vincent Price: The Art of Fear. London: Reynolds and Hearn. Milgram S (1974) Obedience to Authority. New York: Harper and Row. Mortimer WG (1901) Peru. History of Coca, “The Divine Plant” of the Incas. New York: Vail. Nathanson JA, Hunnicutt EJ, Kantham L, and Scavone C (1993) “Cocaine as a Naturally Occurring Insecticide.” Proceedings of the National Academy of Sciences U.S.A. 90: 9645–9648.
References
307
Niemann A (1860) “Über eine neue organische Base in den Cocablaettern” [On a New Organic Base in Coca Leaves]. Archiv des Pharmazie 153: 129–155, 291–308. Noyes HD (1884) “A Few Cursory Notes on the Proceedings of the German Ophthalmological Society in the Middle of September, This Year.” Medical Record 26: 417. Pendergrast M (1993) For God, Country and Coca-Cola: The Unauthorized History of the Great American Soft Drink and the Company That Makes It. New York: Charles Scribner’s Sons. Pendry JB, Holden AJ, Stewart WJ, and Young I (1996) “Extremely Low Frequency Plasmons in Metallic Megostructures.” Physics Review Letters 81: 4773–4776. Pendry JB, Holden AJ, Robbins DJ, and Stewart WJ (1999) “Magnetism from Conductors and Enhanced Nonlinear Phenomena.” IEEE Transactions on Microwave Theory and Techniques 47: 2075–2084. Ruetsch YA, Boeni T, and Borgeat A (2001) “From Cocaine to Ropivacaine: The History of Local Anesthetic Drugs.” Current Topics in Medicinal Chemistry 1: 175–182. Senn B (1996) Golden Horrors: An Illustrated Critical Filmography, 1931–1939. Jefferson, NC: McFarland. Sherriff RC (1968) No Leading Lady: An Autobiography. London: Victor Gollancz. Siodmak C (2001) Wolf Man’s Maker: Memoir of a Hollywood Writer. Lanham, MD: Scarecrow Press. Smith DG (2002) H. G. Wells on Film: The Utopian Nightmare. Jefferson, NC: McFarland. Solomon R (2006) The Passions: Philosophy and the Intelligence of Emotions [Course Guidebook]. Chantilly, VA: Teaching Company. Spillane JF (2000) Cocaine from Medical Marvel to Modern Menace in the United States, 1884–1920. Baltimore, MD: Johns Hopkins University Press. Street JP (1917) The Composition of Certain Patent and Proprietary Medicines. Chicago: American Medical Association. Tancredi L (2005) Hardwired Behavior: What Neuroscience Reveals about Morality. Cambridge: Cambridge University Press. Van Dyke C, and Byck R (1982) “Cocaine.” Scientifi c American 246: 128–141. Veselago VG (1968) “The Electrodynamics of Substances with Simultaneously Negative Values of H and µ.” Soviet Physics USPEKHI 10: 509–514. von Scherzer K (1861) Die Reise der Novara [The Trip of the Novarra]. Wien: Karl Grehold’s Sohn. Weinshilboum R (2003) “Inheritance and Drug Response.” New England Journal of Medicine 348: 529–537. Wells HG (1934) Seven Famous Novels by H. G. Wells. New York: Alfred A. Knopf. Willstätter R (1898) “Über die Constitution der Spaltungsproducte von Atropin und Cocaien” [On the Structure of the Hydrolytic Product of Atropine and Cocaine]. Berlin Deutsches Chemisches Gesselschaft 31: 1534–1553. Willstätter R, Wolfes D, and Maeder H (1923) “Synthese des Natuerlichen Cocains” [Synthesis of the Natural Product Cocaine]. Justis Liebig’s Annalen der Chemie 434: 111–139.
308
References
Chapter 3 Crenshaw M (1990) “Questions to Be Answered, Research to Be Done” in Origins of Terrorism, W Reich, Ed. Cambridge: Cambridge University Press. Crone HD (1992) Banning Chemical Weapons: The Scientifi c Background. Cambridge: Cambridge University Press. Denny D (2005) Risk and Society. London: Sage. Foulkes CH (1934) Gas! The Story of the Specialist Brigade. Edinburgh: Blackwood Press. Galston AW (1969) “Defoliants” in CBW: Chemical and Biological Warfare, S Rose, Ed. Boston: Beacon Press. Goodman A (2007a, August 23) “Psychologists in Denial about Torture.” Seattle (WA) Post-Intelligencer. Available: seattlepi.nwsource.com/opinion/328652_ amy23.html. Goodman A (2007b, August 28) “Renowned Psychologist, Author Returns APA Award over Interrogation Policy.” Democracy Now. Available: www. democracynow.org/article.pl?sid=07/08/28/1527210. Haber LF (1986) The Poisonous Cloud: Chemical Warfare in the First World War. Oxford: Clarendon Press. Hammond C (2003) “Emotional Rollercoaster—Series 1.” BBC Radio. Available: www.bbc.co.uk/radi04/science/rollercoaster1.shtml. Haqqani H, Isham C, Aslan R, and Mogahed D (2007) “The Face of Terrorism: Global Movement or Global Network?” [Panel Discussion]. Aspen Ideas Festival. Available: www.aifestival.org/index2.php?menu=3&sub=1&title= 168&action=full_info. Hoenig SL (2002) The Handbook of Chemical Warfare and Terrorism. Westport, CT: Greenwood Press. International Spy Museum and History Associates Inc. (2004) The Enemy Within: Terror in America—1776 to Today. Washington, DC: International Spy Museum. Jones GW (1991) Black Cinema Treasures: Lost and Found. Denton: University of North Texas Press. LaChance D (2007) “Space Crafts.” Reach, CLA Magazine, University of Minnesota 17. Lessing D (1987) Prisons We Choose to Live Inside. New York: Harper and Row. Los Alamos National Laboratory (2004) “Little Boy and Fat Man.” History of the Atomic Bomb. Available: www.lanl.gov/history/atomicbomb/ littleboyandfatman.shtml. Lytle MH (2007) The Gentle Subversive: Rachel Carson, Silent Spring, and the Rise of the Environmental Movement. New York: Oxford University Press. Melchert M, and List A (2007) “The Thalidomide Saga.” International Journal of Biochemistry and Cell Biology 39: 1489–1499. Miles D (2004) “Despite Not Finding WMDs, Rumsfeld Says World Safer Without Saddam” [Press Release]. American Forces Press Service News Articles. Available: www.defenselink.mil/news/newsarticle.aspx?id=25149. Morgenstern A (2007, December 1) “Small Film, Big Heart.” Wall Street Journal, W3. Newtan SU (2007) Nuclear War I and Other Major Nuclear Disasters of the 20th Century. Bloomington, IN: AuthorHouse.
References
309
Noyes R (1996) Chemical Weapons Destruction and Explosive Waste/Unexploded Ordnance Remediation. Westwood, NJ: Noyes Publications. Peterson J (1997) “Solving the Mystery of the Colorado Brown Stain.” Journal of the History of Dentistry 45: 57–61. Rogers M (2003) “The Psychology of Cyber-Terrorism” in Terrorists, Victims and Society, A Silke, Ed. Chichester, UK: Wiley. Rubenstein L (2007) “TORTURE: National Security and American Values.” Aspen Ideas Festival. Available: www.aifestival.org/index2.php?menu=3&sub =1&title=233&action=full_info. Rumsfeld DH (2002) “DoD News Briefi ng—Secretary Rumsfeld and Gen. Myers” [Transcript]. Available: www.defenselink.mil/transcripts/transcript. aspx?transcriptid=2636. Runciman D (2006) The Politics of Good Intentions. Princeton, NJ: Princeton University Press. Seger KA (2003) “Deterring Terrorists” in Terrorists, Victims and Society, A Silke, Ed. Chichester, UK: Wiley. Solomon R (2006) The Passions: Philosophy and the Intelligence of Emotions [Course Guidebook]. Chantilly, VA: Teaching Company. Soyinka W (2004) “The Changing Mask of Fear” [Lecture]. Climate of Fear, Reith Lecture 1. Available: www.bbc.co.uk/radi04/reith2004/lecture1.shtml. Stoltzenberg D (2004) Fritz Haber: Chemist, Nobel Laureate, German, Jew. Philadelphia: Chemical Heritage Foundation. Thomas K (2007, April 13) “Filmmaker Jost Chronicles America.” Los Angeles Times. Available: articles.latimes.com/2007/apr/13/entertainment/et-jost13. Wink DJ (2001) “ ‘Almost Like Weighing Someone’s Soul’: Chemistry in Contemporary Film.” Journal of Chemical Education 78: 481–483. Young J (2007, August 23) “Pipher Returns Award in Protest.” Lincoln (NE) Journal Star. Available: journalstar.com/articles/2007/08/23/news/local/ doc46ccf4d458103603645050.txt.
Chapter 4 Anonymous (1962, September 8) “Are We Poisoning Ourselves?” Business Week, 36–38. Anonymous (1969, August 1) “The Cities: The Price of Optimism.” Time, 41–43. Available: www.time.com/time/magazine/article/0,9171,901182,00.html. Corwin M (1989, January 28) “The Oil Spill Heard ’Round the Country!” Los Angeles Times. www.es.ucsb.edu/general_info/eshistory.php. Crabb C (2004) “Revisiting the Bhopal Tragedy.” Chemical and Engineering News 306: 1670–1671. Creger TL, and Peryea FJ (1992) “Lead and Arsenic in Two Apricot Cultivars and in ‘Gaia’ Apples Grown on Lead Arsenate-Contaminated Soils.” Hortscience 27: 1277–1281. Darby WJ (1962, October 1) “Silence, Miss Carson.” Chemical and Engineering News, 60–63. Devine K, and Allison S (2006, March) “Strengthening Support for Society’s Underdogs: A Psychological Approach.” Available: (engage.richmond.edu/ projects/AllisonResearch06.pdf). Duchemin M (2007) “Riders of the Whistling Pines,” introduction. Los Angeles, CA: Gene Autry Museum of the West.
310
References
Dunlap TR (1981) DDT: Scientists, Citizens, and Public Policy. Princeton, NJ: Princeton University Press. Fernald CH (1898) “Arsenate of Lead as an Insecticide.” Report of the Massachusetts State Board of Agriculture 45: 353–356. Fitzgerald B (2007, March) “A ‘Rags-to-Riches’ Story: Faculty, Students Study Why We Root for then Abandon the Underdog.” Richmond Now. Available: oncampus.richmond.edu/news/richmondnow/2007/03/rags.html. Harvey M (1962, September 29) “Using a Plague to Fight a Plague.” Saturday Review, 8–10. International Programme on Chemical Safety (1995) “Methomyl: Health and Safety Guide.” Geneva: World Health Organization. Kolata G (2000, April 11) “Reflections; A Hit Movie Is Rated ‘F’ in Science.” New York Times. Lange JH (1994) “Sampling and Evaluation of Soil for Lead Contamination: Criteria for Determining Permissible Concentrations in Soil.” Fresenius Environmental Bulletin 3: 751–756. Lapierre D, and Moro J (2002) Five Past Midnight in Bhopal. New York: Warner Books. Lauger P, Martin H, and Muller P (1944) “Ueber Konstitution und Toxische Wirkung von Natuerlichen und Neuen Synthetischen Insektentoetenden Stoffen” [On the Structure and Toxicity of Natural and New Synthetic Insecticides]. Helvetica Chimica Acta 27: 892–928. Lopate P (2000, April 4) “The Corporation as Fantasy Villain.” New York Times, 24. Magers B (2007) Gene Autry Westerns. Madison, NC: Empire. Markham B (1892–1893) “On the First Use of Paris Green for the Potato-Beetle.” Insect Life 5: 44. Matthiessen P (1999, March 29) “Rachel Carson.” Time. Available: www.time. com/time/time100/scientist/profi le/carson.html. Merwin I, Pruyne PT, Ebel JG, Manzell KL, and Lisk DJ (1994) “Persistence, Phytotoxicity, and Management of Arsenic, Lead, and Mercury Residues in Old Orchard Soils of New York State.” Chemosphere 29: 1361–1367. Noyori R (2005) “Pursuing Practical Elegance in Chemical Synthesis.” Chemical Communications 1807–1811. O’Brien TJ, Ceryak S, and Patierno SR (2003) “Complexities of Chromate Carcinogenesis: Role of Cellular Response, Repair, and Recovery Mechanisms.” Mutation Research 533: 3–36. Ohio Historical Society (2005) “Cuyahoga River Fire.” Ohio History Central: Online Encyclopedia of Ohio History. Available: www.ohiohistorycentral. org/entry.php?rec=1642. Patnaik P (1992) A Comprehensive Guide to the Hazardous Properties of Chemical Substances. New York: Van Nostrand Reinhold. Peryea FJ (1998) “Phosphate Starter Fertilizer Temporarily Enhances Soil Arsenic Uptake by Apple Trees Grown under Field Conditions.” Hortscience 33: 826–829. Ribstein LE (2005) “Wall Street and Vine: Hollywood’s View of Business.” University of Illinois Law and Economics Research Paper No. LE05–010. Available: ssrn.com/abstract=563181. Robin C (2004) Fear: The History of a Political Idea. Oxford: Oxford University Press.
References
311
Sales K (1993) The Green Revolution: The American Environmental Movement, 1962–1992. New York: Hill and Wang. Schlosberg D (2005) “Networks and Mobile Arrangements: Organisational Innovation in the US Environmental Justice Movement” in Debating the Earth, J Dryzak and D Schlosberg, Eds. Oxford: Oxford University Press. Shepard HH (1939) The Chemistry and Toxicity of Insecticides. Minneapolis, MN: Burgess. Union Carbide (2001) “Statement of Union Carbide Corporation Regarding the Bhopal Tragedy.” Available: www.bhopal.com/ucs.html. Van Aken K, Strekowski L, and Patiny L (2006) “EcoScale, a Semi-Quantitative Tool to Select an Organic Preparation Based on Economical and Ecological Parameters.” Beilstein Journal of Organic Chemistry 2: online article 3. Available: bjoc.beilstein-journals.org/content/2/1/3. Welsh N (1989, January 26) “The Big Spill.” Santa Barbara Independent. Available: http://www.es.ucsb.edu/general_info/eshistory.php. Whorton J (1974) Before Silent Spring: Pesticides and Public Health. Princeton, NJ: Princeton University Press. Ziedler O (1874) “Verbindug von Chloral mit Brom- und Chlorbenzol” [Studies of Chloroacetal with Bromo- and Chlorobenzene]. Berichte der deutschen Chemischen Gesellschaft zu Berlin [Reports of the German Chemical Society to Berlin] 8: 1180–1181.
Chapter 5 Considine S (1994) Mad as Hell: The Life and Work of Paddy Chayefsky. New York: Random House. Funk CK, and Koob GF (2007) “A CRF(2) Agonist Administered into the Central Nucleus of the Amygdala Decreases Ethanol Self-Administration in Ethanol-Dependent Rats.” Brain Research 1155: 172–178. Gaoni Y, and Mechoulam R (1964) “Isolation, Structure, and Partial Synthesis of an Active Constituent of Hashish.” Journal of the American Chemical Society 86: 1646–1648. Gottfried M (1990) All His Jazz: The Life and Death of Bob Fosse. New York: Da Capo Press. Gray MT (2005) “The Shifting Sands of Self: A Framework for the Experience of Self in Addiction.” Nursing Philosophy 6: 119–130. Grens K (2007, June) “Addictive Research.” The Scientist, 48–55. Heilig M, and Koob GF (2007) “A Key Role for Corticotropin-Releasing Factor in Alcohol Dependence.” Trends in Neuroscience 30: 399–406. Heilig M, Koob GF, Ekman R, and Britton T (1994) “Corticotropin-Releasing Factor and Neuropeptide Y—Role in Emotional Integration.” Trends in Neuroscience 17: 80–85. Hofmann A (1980) LSD—My Problem Child. New York: McGraw-Hill. James W (1890) “The Stream of Thought” in The Principles of Psychology, Vol. 1. New York: Holt, chapter 9. Koob GF, and Le Moal M (2006) Neurobiology of Addiction. Amsterdam: Academic Press. Leeb-Lundberg F, Snowman A, and Olsen RW (1980) “Barbiturate Receptor Sites Are Coupled to Benzodiazepine Receptors.” Proceedings of the National Academy of Sciences U.S.A. 77: 7468–7472.
312
References
Lieberman JA, Stroup TS, McElvoy JP, Swartz MS, Rosenheck RA, Perkins DO, Keefe RS, Davis SM, Davis CE, Lebowitz BD, Severe J, and Hsiao JK (2005) “Effectiveness of Antipsychotic Drugs in Patients with Chronic Schizophrenia.” New England Journal of Medicine 353: 1209–1223. Lopez-Munoz F, Alamo C, Cuenca E, Shen WW, Clervoy P, and Rubio G (2005) “History of the Discovery and Clinical Introduction of Chlorpromazine.” Annals of Clinical Psychiatry 17: 113–135. O’Brien CP (1997) “A Range of Research-Based Pharmacotherapies for Addiction.” Science 278: 66–70. Robinson D (2007) Consciousness and Its Implications, Lecture Two: Self Consciousness [Course Guidebook]. Chantilly, VA: Teaching Company. Robinson V (1946) Victory over Pain: A History of Anesthesia. New York: Henry Schuman. Satel S, and Goodwin F (1998) Is Drug Addiction a Brain Disease? Washington, DC: Ethics and Public Policy Center. Scheff J (2007, December 1) “Schizophrenia: The Treatments.” The Scientist, 49–57. Available: www.the-scientist.com/2007/12/01/s48/1/. Sellman D (2007) “What Is Addiction?” New Zealand Family Physician 34: 77–81. Shulgin A (1986) “The Background and Chemistry of MDMA.” Journal of Psychoactive Drugs 18: 291–304. Starks M (1982) Cocaine Fiends and Reefer Madness: An Illustrated History of Drugs in the Movies. New York: Cornwall. Weiden PJ (2007) “Discontinuing and Switching Antipsychotic Medications: Understanding the CATIE Schizophrenic Trial.” Journal of Clinical Psychiatry 68 (Supplement): 12–19. Winter A (2003) “The Chemistry of Truth and the Literature of Dystopia” in Literature, Science, Psychoanalysis, 1830–1970: Essays in Honour of Gillian Beer, H Small and T Tate, Eds. Oxford: Oxford University Press. Wright D (1994) “ ‘The Prisonhouse of My Disposition’: A Study of the Psychology of Addiction in ‘Dr. Jekyll and Mr. Hyde.’ ” Studies in the Novel 26: 254–267.
Chapter 6 Ackerman D (1999) Deep Play. New York: Random House. Anonymous (1877) “The Talking Phonograph.” Scientifi c American 36: 2. Anonymous (2006, November 30) “The Edison of Our Age?” The Economist. Available: www.economist.com/printedition/PrinterFriendly.cfm?story_id= 8312367. Balderfroid K (1997, January 24) “Seaborg an Ever-Popular Resource.” Berkeley Lab Currents. Available: www.lbl.gov/publications/currents/archive/jan-24– 1997.html. Ball P (2006, June) “Chancing upon Chemical Wonders.” Chemistry World. Available: www.rsc.org/chemistryworld/Issues/2006/June/ ChancingUponChemicalWonders.asp. Bardeen J, Cooper LN, and Schrieffer JR (1957) “Microscopic Theory of Superconductivity.” Physics Review 108: 162–164. Bowser E (1990) The Transformation of Cinema: 1907–1915. New York: Charles Scribner’s Sons. Buntar V, and Weber HW (1996) “Magnetic Properties of Fullerene Superconductors.” Superconductor Science and Technology 9: 599–615.
References
313
Cooper LN (1956) “Bound Electron Pairs in a Degenerate Fermi Gas.” Physics Review 104: 1189–1190. Corak WS, Goodman BB, Satterthwaite CB, and Wexler A (1956) “Atomic Heats of Normal and Superconducting Vanadium.” Physics Review 102: 656–661. Csikszentmihalyi M (1996) Creativity: Flow and the Psychology of Discovery and Invention. New York: HarperCollins. Debre P (1996) Louis Pasteur. Baltimore, MD: Johns Hopkins University Press. Flicker E (2003) “Between Brain and Breasts—Women Scientists in Fiction Film: On the Marginalization and Sexualization of Scientific Competence.” Public Understanding of Science 12: 307–318. Florida R (2002) The Rise of the Creative Class: And How It’s Transforming Work, Leisure, Community and Everyday Life. New York: Basic Books. Harmon J (1996) “Hubert Alyea, Innovator in Science Teaching, Dies at 93” [press release]. Princeton University. Available: www.princeton.edu/pr/ news/96/q4/1022alye.html. Internet Movie Database (2008) “Internet Movie Database.” Available: www. imdb.com. Jehl F (1936 and 1938) Menlo Park Reminiscences. Dearborn, MI: Edison Institute. Kauffman GB, and Myers RD (1975) “The Resolution of Racemic Acid.” Journal of Chemical Education 52: 777–781. Kroto HW, Heath JR, O’Brien SC, Curl RF, and Smalley RE (1985) “C60: Buckminsterfullerene.” Nature 318: 162–163. Livesay HC (2007) American Made: Shapers of the American Economy. New York: Pearson-Longman. McCarthy RE (1992) Secrets of Hollywood Special Effects. Boston: Focal Press. Musser C (1991) Before the Nickelodeon: Edwin S. Porter and the Edison Manufacturing Company. Berkeley: University of California Press. Musser C (1997) Edison Motion Pictures, an Annotated Filmography, 1890–1900. Washington, DC: Smithsonian Institution Press. Noonan B (2005) Women Scientists in Fifties Science Fiction Films. Jefferson, NC: McFarland. Peiss K (1998) Hope in a Jar: The Making of America’s Beauty Culture. New York: Metropolitan Books. Roberts RM (1989) Serendipity: Accidental Discoveries in Science. New York: John Wiley and Sons. Shakhashiri BZ (1997) “Hubert N. Alyea, 1903–1996.” Journal of Chemical Education 74: 20. Stout JW, and Jones WM (1947) “Calorimetric Determination of the Energy Produced by Plutonium (239).” Physics Reviews 71: 582–585. Vanderbilt BM (1971) Thomas Edison, Chemist. Washington, DC: American Chemical Society. Walt Disney Pictures (1997) “Flubber Production Notes.” Available: www.cineramareview.com/static-htm/11917mp.htm. Wilson EO (1998) Consilience: The Unity of Knowledge. New York: Alfred A. Knopf. Wu MK, Ashburn JR, Torng CJ, Hor PH, Meng RL, Gao L, Huang ZJ, Wang YQ, and Chu CW (1987) “Superconductivity at 93 K in a New Mixed-Phase Y-Ba-Cu-O Compound System at Ambient Pressure.” Physics Review Letters 58: 908–910.
314
References
Chapter 7 Baird D (1993) “Analytical Chemistry and the ‘Big’ Scientific Instrumentation Revolution.” Annals of Science 50: 267–290. Barbour AG (1970) Days of Thrills and Adventures. New York: MacMillan. Bertillon A (1893) Identifi cation Anthropometrique: Instructions Signaletiques. Paris: Melun Imprimerie Administrative. Bisbing RE (2004) “Fractured Patterns: Microscopical Investigations of Real Physical Evidence.” Modern Microscopy Journal. Available: www. modernmicroscopy.com/main.asp?article=11. Chisum WJ, and Turvey BE (2000) “Evidence Dynamics: Locard’s Exchange Principle and Crime Reconstruction.” Journal of Behavioural Profiling 1. Available: www.profi ling.org/journal/v011_n01/jbp_ed_january2000_1–1. html. Coel M (1988, Winter) “Keeping Time by the Atom.” White House Historical Foundation. Available: AmericanHeritage.com/articles/magazine/it/1988/3/ 1988_3_42.shtml. Cole SA (2001) Suspect Identities: A History of Fingerprinting and Criminal Investigation. Cambridge, MA: Harvard University Press. Crouch EAC, and Wilson R (1982) Risk/Benefit Analysis. Cambridge, MA: Ballinger. Crummett WB (2002) Decades of Dioxin: Limelight on a Molecule. Philadelphia: Xlibris. Currie LA (1968) “Limits for Quantitative Detection and Quantitative Determination.” Analytical Chemistry 40: 586–593. Gerber SM (1983) “A Study in Scarlet: Blood Identification in 1875” in Chemistry and Crime: From Sherlock Holmes to Today’s Courtroom, SM Gerber, Ed. Washington, DC: American Chemical Society. Girard G (1994) “The Third Verification of National Prototypes of the Kilogram (1988–1992).” Metrologia 31: 317–336. Haycraft H (1984) Murder for Pleasure: The Life and Times of the Detective Story. New York: Carroll and Graf. Hingham C (1976) The Adventures of Conan Doyle: The Life of the Creator of Sherlock Holmes. New York: Norton. Irwin JT (1994) The Mystery to a Solution: Poe, Borges, and the Analytic Detective Story. Baltimore, MD: Johns Hopkins University Press. Labianca DA (1998) “Rohypnol: Profi le of the ‘Date-Rape Drug.’ ” Journal of Chemical Education 75: 719–722. Lahue KC (1968) Bound and Gagged: The Story of the Silent Serials. Cranberry, NJ: A. S. Barnes. Locard E (1930) “The Analysis of Dust Traces.” American Journal of Police Science 1: 276–298. Marsh J (1836) “Account of a Method of Separating Small Quantities of Arsenic from Substances with Which It May Be Mixed.” Edinburgh New Philosophical Journal 21: 229–236. Mavis P (2001) The Espionage Filmography: United States Releases, 1898 Through 1999. Jefferson, NC: McFarland. McCarthy RE (1992) Secrets of Hollywood Special Effects. Boston: Focal Press. Pitts MR (1979) Famous Movie Detectives. Metuchen, NJ: Scarecrow Press. Pyrhoenen H (1999) Mayhem and Murder: Narrative and Moral Problems in the Detective Story. Toronto: University of Toronto Press.
References
315
Rainey B (1999) Series and Serials: A World Filmography. Jefferson, NC: McFarland. Rhodes H (1968) Alphonse Bertillon: Father of Scientifi c Detection. New York: Greenwood Press. Silverman K (1991) Edgar A. Poe: Mournful and Never-Ending Remembrance. New York: HarperCollins. Snyder LJ (2004) “Sherlock Holmes: Scientific Detective.” Endeavour 28: 104–108. Steinbrunner C, and Penzler O, Eds. (1976) Encyclopedia of Mystery and Detection. New York: McGraw-Hill. Webster SH (1947) “The Development of the Marsh Test for Arsenic.” Journal of Chemical Education 24: 487–490. Wilson R (1984) “Commentary: Risks and Their Acceptability.” Science, Technology, and Human Values 9: 11–22. Zweig G (1970) “The Vanishing Zero—The Evolution of Pesticide Analyses” in Essays in Toxicology, Vol. 2, FR Blood, Ed. 157–198. Zweig G (1978) “The Vanishing Zero—Ten Years Later.” Journal of the Association of Offi cial Analytical Chemists 61: 229–248.
Chapter 8 Anderson WC (1984) Between the Library and the Laboratory: The Language of Chemistry in Eighteenth-Century France. Baltimore, MD: Johns Hopkins University Press. Ball P (2005) Elegant Solutions: Ten Beautiful Experiments in Chemistry. Cambridge: Royal Society of Chemistry. Beynon JH, and Morgan RP (1978) “The Development of Mass Spectrometry: An Historical Account.” International Journal of Mass Spectrometry and Ion Physics 27: 1–30. Carson R (1956, July) “Help Your Child to Wonder.” Woman’s Home Companion, 46–48. Available: training.fws.gov/history/carson/carsonwonder.pdf. Carson RL (1965) The Sense of Wonder. New York: Harper and Row. Clute J, and Nicholls P (1993) The Encyclopedia of Science Fiction. New York: St. Martin’s Press. Conley T (1991) Film Hieroglyphs: Ruptures in Classical Cinema. Minneapolis: University of Minnesota Press. Curie E (1937) Madame Curie. New York: Literary Guild of America. De Bievre P, and Peiser HS (1992) “ ‘Atomic Weight’—The Name, Its History, Defi nition, and Units.” Pure and Applied Chemistry 64: 1535–1543. Dunaway DK (1989) Huxley in Hollywood. New York: Harper and Row. Elena A (1997) “Skirts in the Lab: Madame Curie and the Image of the Woman Scientist in the Feature Film.” Public Understanding of Science 6: 269–278. Eriksen KK (2002) “The Future of Tertiary Chemical Education.” HYLE: International Journal for Philosophy of Chemistry 8: 35–48. Felter HW, and Lloyd JU (1898) King’s American Dispensary. Cincinnati: Ohio Valley Co. Flicker E (2003) “Between Brain and Breasts—Women Scientists in Fiction Film: On the Marginalization and Sexualization of Scientific Competence.” Public Understanding of Science 12: 307–318. Freemantle M (2003, August 25) “Chemistry at Its Most Beautiful.” Chemical and Engineering News, 27–30.
316
References
Gardner H (2006) Five Minds for the Future. Boston: Harvard Business School Press. Gruber HE (1981) “On the Relation Between ‘Aha Experiences’ and the Construction of Ideas.” History of Science 19: 41–59. Hill EA (1900) “On a System of Indexing Chemical Literature; Adopted by the Classification Division of the U.S. Patent Office.” Journal of the American Chemical Society 22: 478–494. Holmes FL (1998) Antoine Lavoisier—The Next Crucial Year; Or, The Source of His Quantitative Methods of Chemistry. Princeton, NJ: Princeton University Press. Jensen WB (2005) “The Symbolism of Chemical Equations.” Journal of Chemical Education 82: 1461. Klein U (2003) Experiments, Models, Paper Tools: Cultures of Organic Chemistry in the Nineteenth Century. Stanford, CA: Stanford University Press. Lafleur LJ (1956) “Introduction” in Discourse on Method by René Descartes. Indianapolis: Bobbs-Merrill. Latham A (1971) Crazy Sundays: F. Scott Fitzgerald in Hollywood. New York: Viking Press. LeRoy M, and Kleiner R (1974) Mervyn LeRoy: Take One. New York: Hawthorn Books. Lewis ES (1998) A Biography of the Distinguished Gilbert Newton Lewis. Lewiston, NY: Edwin Mellen Press. Lewis GN (1923) Valence and the Structure of Atoms and Molecules. New York: Dover Publications. Mattauch J (1958) “The Rational Choice of a Unified Scale for Atomic Weights and Nuclidic Masses.” Journal of the American Chemical Society 80: 4125–4126. Melhado EM (1981) Jacob Berzelius: The Emergence of His Chemical System. Madison: University of Wisconsin Press. Mills I, Cvitas T, Homan K, Kallay N, and Kuchitsu K (1993) Quantities, Units and Symbols in Physical Chemistry. Oxford: Blackwell Science. Nebergall WH (deceased), Holtzclaw HF Jr, and Robinson C (1980) General Chemistry with Qualitative Analysis. Lexington, MA: D. C. Heath. Neibaur JL, and Okuda T (1995) The Jerry Lewis Films: An Analytical Filmography of the Innovative Comic. Jefferson, NC: McFarland. Perrin C (1989) “Document, Text, Myth: Lavoisier’s Crucial Year Revisited.” British Journal of the History of Science 22: 3–25. Poirier J-P (1996) Lavoisier: Chemist, Biologist, Economist. Philadelphia: University of Pennsylvania Press. Reed JW (1989) American Scenarios: The Uses of Film Genre. Middletown, CT: Wesleyan University Press. Reid R (1974) Marie Curie. New York: Saturday Review Press/E. P. Dutton. Robinson V (1946) Victory over Pain: A History of Anesthesia. New York: Henry Schuman. Rocke AJ (2008) “Berzelius, Joens Jacob” in Complete Dictionary of Scientifi c Biography, Vol. 19. Detroit: Charles Scribner’s Sons, 266–269. Savisky ZM (2006) “Portrait: Tom Conley.” Harvard Crimson. Available: www. thecrimson.com/article.aspx?ref=511945. World Anti-Doping Agency (2002) “A Brief History of Anti-Doping.” Available: http://www.wada-ama.org/en/dynamic.ch2?pageCategory.id=312.
References
317
Yesalis CE, Ed. (2000) Anabolic Steroids in Sport and Exercise. Champagne IL: Human Kinetics Press. Yesalis CE, Kennedy NJ, Kopstein AN, and Bahrke MS (1993) “AnabolicAndrogenic Steroid Use in the United States.” Journal of the American Medical Association 270: 1217–1221.
Chapter 9 Aberihani, West SL, and Price CA (2003) “Lawrence Chenault” in Encyclopedia of the Harlem Renaissance. New York: Checkmark Books, 217–219. Abrams I (1988) The Nobel Peace Prize and the Laureates: An Illustrated Biographical History, 1901–1987. Boston: G. K. Hall. Anonymous (2007, July 2) “Gains in Chemical Output Continue.” Chemical and Engineering News, 55–64. Antkowiak B (2001) “How Do General Anesthetics Work?” Naturwissenschaften 88: 201–213. Atkins PW (1978) Physical Chemistry. San Francisco: W. H. Freeman. Aubourg P, Adamsbaum C, Lavallard-Rousseau M-C, Rocchiccioli F, Cartier N, Jambaque I, Jakobezak C, Lemaitre A, Boureau F, Wolf C, and Bougneres P-F (1993) “A Two-Year Trial of Oleic and Erucic Acids (‘Lorenzo’s Oil’) as Treatment for Adrenomyeloneuropathy.” New England Journal of Medicine 329: 745–752. Avery ME (2000) “Biographical Memoir of Gertrude B. Elion.” Proceedings of the National Academy of Sciences U.S.A. 78: 16–29. Baumler E (1984) Paul Ehrlich: Scientist for Life. New York: Holmes and Meier. Bergengren E (1962) Alfred Nobel: The Man and His Work. London: Thomas Nelson and Sons. Brems KAH (1988) “The Archival Quality of Film Base.” Society of Motion Picture and Television Engineering 97: 991–993. Butters GR Jr (2002) Black Manhood on the Silent Screen. Lawrence: University Press of Kansas. Davis W (1985) The Serpent and the Rainbow. New York: Simon and Schuster. Davis W (1988a) Passage of Darkness: The Ethnobiology of the Haitian Zombie. Chapel Hill, NC: UNC Press. Davis W (1988b) “Zombification” [Letter]. Science 240: 1715–1716. Davis W (1997) One River: Explorations and Discoveries in the Amazon Rain Forest. New York: Simon and Schuster. Delgado-Herrera L, Ostroff RD, and Rogers SA. (2001) CNS Drug Review 7: 48–120. Desbarax P (2002) “Morton’s Design of the Early Ether Vaporisers.” Anaesthesia 57: 463–469. Dolan JE (1997) “The Nobel Era” in Explosives in the Service of Man: The Nobel Heritage, JE Dolan and SS Langer, Eds. Cambridge: Royal Society of Chemistry. Evlanoff M, and Fluor M (1969) Alfred Nobel: The Loneliest Millionaire: Ward Ritchie Press. Gray T (1976) Champions of Peace: The Story of Alfred Nobel, the Peace Prize, and the Laureates. New York: Paddington Press. Griep MA, and Mikasen ML (2005) “Based on a True Story: Using Movies as Source Material for General Chemistry Reports.” Journal of Chemical Education 82: 1501–1503.
318
References
Halstead BW (1988) Poisonous and Venomous Marine Animals of the World. Princeton, NJ: Darwin Press. Holmes OW (1847) Some Account of the Letheon; Or Who Was the Discoverer? Boston: Dutton and Wentworth. Image Permanence Institute (1991, February 1) “Brief Summary Report of Research, Storage Conditions, and Deterioration in Acetate and Nitrate Film, to the Film Foundation Archival Council.” Rochester, NY. Jobson MR, and Knapp van Bogaert D (2005) “Just a Story or a ‘Just Story’?: Ethical Issues in a Film with a Medical Theme” in Signs of Life: Medicine and Cinema, G Harper and A Moor, Eds. London: Wallflower Press. Jones AH (2000) “Medicine and the Movies: Lorenzo’s Oil at Century’s End.” Annals of Internal Medicine 133: 567–571. Kresge N, Simoni RD, and Hill RL (2008a) “Developing the Purine Nucleoside Analogue Acyclovir: The Work of Gertrude B. Elion.” Journal of Biological Chemistry 283: e11–e13. Kresge N, Simoni RD, and Hill RL (2008b) “The Rational Design of Nucleic Acid Inhibitors to Treat Leukemia: The Work of George H. Hitchings.” Journal of Biological Chemistry 283: e10–e11. Lederer SE, and Parascandola J (1998) “Screening Syphilis: Dr Ehrlich’s Magic Bullet Meets the Public Health Service.” Journal of the History of Medicine and Allied Sciences 53: 345–370. Lee M-J, Jeong D-Y, Kim W-S, Kim H-D, Kim C-H, Park W-H, Park Y-H, Kim K-S, Kim H-M, and Kim D-S (2000) “A Tetrodotoxin-Producing Vibrio Strain, LM-1, from the Puffer Fish Fugu vermicularis radiatus.” Applied and Environmental Microbiology 66: 1698–1701. Lloyd NC, Morgan HW, Nicholson BK, and Ronimus RS (2005) “The Composition of Ehrlich’s Salvarsan: Resolution of a Century-Old Debate.” Angewandte Chemie International Edition 44: 941–944. McGrayne SB (1993) Nobel Prize Women in Science: Their Lives, Struggles, and Momentous Discoveries. New York: Birch Lane Press. Morton WTG (1847) “Letter from Dr. Morton, of Boston, U.S.A.” Lancet 2: 80. Moser HW (1993) “Lorenzo’s Oil” [letter]. Lancet 341: 544. Moser HW, Raymond GV, Lu S-E, Muenz LR, Moser AB, Xu J, Jones RO, Loes DJ, Melhem ER, Dubey P, Bezman L, Brereton NH, and Odone A (2005) “Follow-up of 89 Asymptomatic Patients with Adrenoleukodystrophy Treated with Lorenzo’s Oil.” Archives of Neurology 62: 1073–1080. Murrell W (1879) “Nitro-Glycerine as a Remedy for Angina Pectoris.” Lancet 1: 80–81, 113–115, 151–152, 225–227. National Science Foundation (2006) “Investing in America’s Future: Strategic Plan FY 2006–2011.” NSF Report 06–48. Available: www.nsf.gov/statistics/ infbrief/nsf08301/. Noyes R (1996) Chemical Weapons Destruction and Explosive Waste/Unexploded Ordnance Remediation. Westwood, NJ: Noyes Publications. Patnaik P (1992) A Comprehensive Guide to the Hazardous Properties of Chemical Substances. New York: Van Nostrand Reinhold. Riethmiller S (2005) “From Atoxyl to Salvarsan: Searching for the Magic Bullet.” Chemotherapy 51: 234–242. Ritter SK (2004, November 29) “Redefi ning Chemistry.” Chemical and Engineering News, 31.
References
319
Robinson V (1946) Victory over Pain: A History of Anesthesia. New York: Henry Schuman. Sampson HT (1977) Blacks in Black and White: A Source Book on Black Films. Metuchen, NJ: Scarecrow Press. Simidu U, Noguchi T, Hwang D-F, Shida Y, and Hashimoto K (1987) “Marine Bacteria Which Produce Tetrodotoxin.” Applied and Environmental Microbiology 53: 1714–1715. Slide A (1992) Nitrate Won’t Wait: Film Preservation in the United States. Jefferson, NC: McFarland. Smith A (1981, May) “Mixing Chemistry with Culture: Preserving Film and Television.” Screen Digest, 87–91. Smith EB, and Daniels S, Eds. (1998) Gases in Medicine: Anaesthesia. Cambridge: Royal Society of Chemistry. Swann JP (1985) “Arthur Tatum, Parke-Davis, and the Discovery of Mapharsen as an Antisyphilitic Agent.” Journal of the History of Medicine and Allied Sciences 40: 167–187. Tatum AL, and Cooper GA (1932) “Meta-Amino Para-Hydroxy Phenyl Arsine Oxide as an Antisyphilitic Agent.” Science 75: 541–542. Voegtlin C (1925) “Arsphenamine and Related Arsenicals.” Physiological Reviews 5: 63–94. Whitesides GM (2004) “Assumptions: Taking Chemistry in New Directions.” Angewandte Chemie International Edition 43: 3632–3641. Wink DJ (2001) “ ‘Almost Like Weighing Someone’s Soul’: Chemistry in Contemporary Film.” Journal of Chemical Education 78: 481–483.
Chapter 10 Adams F (1891) The Genuine Works of Hippocrates. New York: William Wood. Altman LK (1987) Who Goes First? The Story of Self-Experimentation in Medicine. New York: Random House. Beecher HK (1966) “Ethics and Clinical Research.” New England Journal of Medicine 274: 1354–1360. Beecher HK (1970) Research and the Individual: Human Studies. Boston: Little, Brown. Bowden ME, Crow AB, and Sullivan T (2003) Pharmaceutical Achievers: The Human Face of Pharmaceutical Research. Philadelphia: Chemical Heritage Foundation. Bradford Hill A (1951) “The Clinical Trial.” British Medical Bulletin 7: 278–282. Bradford Hill A (1963) “Medical Ethics and Controlled Trials.” British Medical Journal 1: 1043–1049. Heller J (1972, July 26) “U.S. Testers Let Many Die of Syphilis.” Washington Post, A1, A6. Jonsen AR (1998) “The Ethics of Research with Human Subjects: A Short History” in Source Book in Bioethics, AR Jonsen, RM Veatch, and L Walters, Eds. Washington, DC: Georgetown University Press. Jonsen AR, Veatch RM, and Walters L, Eds. (1998) Source Book in Bioethics. Washington, DC: Georgetown University Press. Lederer SE (1995) Subjected to Science: Human Experimentation in America before the Second World War. Baltimore, MD: Johns Hopkins University Press.
320
References
Moser HW, Raymond GV, Lu S-E, Muenz LR, Moser AB, Xu J, Jones RO, Loes DJ, Melhem ER, Dubey P, Bezman L, Brereton NH, and Odone A (2005) “Follow-up of 89 Asymptomatic Patients with Adrenoleukodystrophy Treated with Lorenzo’s Oil.” Archives of Neurology 62: 1073–1080.
Appendix 1 Alyea HN (1955) “Tested Demonstrations in General Chemistry.” Journal of Chemical Education 32: 28–29. Alyea HN (1957) Tested Demonstrations in General Chemistry. Easton, PA: American Chemical Society. Alyea HN, and Dutton FB (1969) Tested Demonstrations in General Chemistry. Easton, PA: American Chemical Society. Beall H (1993) “Literature Reading and Out-of-Class Essay Writing in General Chemistry.” Journal of Chemical Education 70: 10–11. Bodner GM (2001) “Why Lecture Demonstrations Are ‘Exocharmic’ for both Students and Their Instructors.” University Chemical Education 5: 31–35. Borgwald JM and Schreiner S (1994) “Science in the Movies: The Good, the Bad, and the Ugly: A Novel Interdisciplinary Course for Teaching Science to Nonscience Majors.” Journal of College Science Teaching 23: 367–371. Cornely K (2003) “Content and Confl ict—the Use of Current Events to Teach Content in a Biochemistry Course.” Biochemistry and Molecular Biology Education 31: 173–176. Dubeck LW, Moshier SE, and Boss JE (1988) Science in Cinema: Teaching Science Fact Through Science Fiction Films. New York: Teachers College Press. Dubeck LW, Moshier SE, and Boss JE (2004) Fantastic Voyages: Learning Science Through Science Fiction Films. New York: Springer. Griep MA and Mikasen ML (2005) “Based on a True Story: Using Movies as Source Material for General Chemistry Reports.” Journal of Chemical Education 82: 1501–1503. Herreid CF (1994) “Case Studies in Science—a Novel Method of Science Education.” Journal of College Science Teaching 23: 221–229. Herreid CF (2004) “Case Collection.” Available: ublib.buffalo.edu/libraries/ projects/cases/ubcase.htm. Jones MB, and Miller CR (2001) “Chemistry in the Real World.” Journal of Chemical Education 78: 484–487. Koprowski JL (1997) “Sharpening the Craft of Scientific Writing: A Peer-Review Strategy to Improve Student Writing.” Journal of College Science Teaching 27: 133–135. Kovac J, and Sherwood DW (1999) “Writing in Chemistry: An Effective Learning Tool.” Journal of Chemical Education 76: 1399–1403. Ramette RW (1980) “Exocharmic Reactions.” Journal of Chemical Education 57: 68–69. Resch KE, and Schicker VD (1992) Using Film in the High School Curriculum: A Practical Guide for Teachers and Librarians. Jefferson, NC: McFarland. Rose C (2003) “How to Teach Biology Using the Movie Science of Cloning People, Resurrecting the Dead, and Combining Flies and Humans.” Public Understanding of Science 12: 289–296.
References
321
Shakhashiri BZ (1983–1989) Chemical Demonstrations. Madison: University of Wisconsin. Shibley IA Jr, Milakofsky LK, and Nicotera CL (2001) “Incorporating a Substantial Writing Assignment into Organic Chemistry: Library Research, Peer Review, and Assessment.” Journal of Chemical Education 78: 50–53. Wink DJ (2001) “ ‘Almost Like Weighing Someone’s Soul’: Chemistry in Contemporary Film.” Journal of Chemical Education 78: 481–483.
This page intentionally left blank
Index
MOVIE INDEX actors Adams, Edie, 176 Addams, Dawn, 30 Affleck, Ben, 88 Anderson, Stanley, 87 Arkin, Alan, 215 Arnaz, Desi, 273 Arnold, Edward, 179 Autry, Gene, 131 Backus, Jim, 59 Bacon, Kevin, 57 Baggott, King, 25 Baldwin, William, 201, 210 Ball, Lucille, 160, 164, 183–184, 273 Barry, Raymond J., 169 Barrymore, John, 32, 61 Basinger, Kim, 211 Baskin, Elya, 212 Bassermann, Albert, 246, 257 Bates, Alan, 88 Bates, Ralph, 29 Bedelia, Bonnie, 91, 94 Benham, Harry, 33 Beregi Sr., Oscar, 82 Bergman, Ingrid, 31 Berry, Halle, 164 Beswick, Martine, 29 Betts, Jack, 86 Bixby, Bill, 271 Bouchey, Willis, 273 Bowman, Lee, 204 Bracco, Lorraine, 265 Britton, Pamela, 219 Brolin, Josh, 57 Brooke, Hillary, 275 Brown, Clancy, 169 Brown, G. Edward, 282 Bruce, Virginia, 61 Butler, Dan, 127 Cabot, Susan, 164 Cage, Nicholas, 89 Caine, Michael, 85 Campbell, Julia, 173 Carey, Harry, 277
Cariou, Len, 117, 118 fig Casey, Bernie, 26 Cash, Rosalind, 26 Catillon, Brigitte, 207 Chenault, Lawrence, 282–283 Chiles, Lois, 213 Cobb, Lee J., 218 Coburn, Charles, 181, 274 Coburn, James, 218, 242 Connery, Sean, 89, 99, 206, 212, 215, 265 Cooper, Chris, 238 Cotton, Joseph, 130 Courtney, Inez, 183 Crisp, Donald, 31 Cruze, James, 33 Cumming, Alan, 172 Curran, Tony, 206 Cusack, John, 91–92 Cushing, Peter, 28 Dafoe, Willem, 86 Daly, Arnold, 220 Davis, Gilbert, 279 Day, Doris, 160, 174, 175 DeHaven, Gloria, 179 Demerest, William, 276 DeNiro, Robert, 210 Dern, Bruce, 138, 142 Dern, Laura, 238 Donat, Robert, 101 Dowling, Doris, 164 Duke, Patty, 152 Dutronc, Jacques, 207 Duvall, Robert, 150, 215 Eckhart, Aaron, 143 Elizondo, Hector, 236 Fabares, Shelley, 271 Field, Betty, 276 Fiennes, Ralph, 119 Finlay, Frank, 99 Finney, Albert, 123 Flemyng, Jason, 206 Flynn, Joe, 58 Foldi, Erzsebet, 149
323
324 actors (continued) Fox, Michael J., 173 Francis, Kay, 164 Fraser, Brendan, 85 Freeman, Kenneth, 100 Freeman, Morgan, 88 Frome, Milton, 231 Gaddis, Marshall, 95, 95 fig Gargan, William, 275 Garofalo, Janeane, 172 Garson, Greer, 246, 248 fig Gilbert, Earle, 249 Glover, Crispin, 174 Golan, Gila, 218 Gordon, Ruth, 257 Grant, Cary, 274 Greaves, William, 100 Greenwood, Joan, 166 Gregory, James, 271 Grodin, Charles, 129 Guinness, Alec, 155, 160, 165 Gunn, Judy, 279 Guyse, Shelia, 100 Guzman, Luis, 208 Gyllenhaal, Jake, 238 Hall, Jon, 60 Hamilton, Margaret, 61 Hamilton, Murray, 154 Harden, Marcia Gay, 169 Hardwicke, Cedric, 60, 63 Harrigan, William, 56 Harris, Ed, 89 Harris, Richard, 99, 174 Harris, Rosemary, 86 Harvey, Paul, 180 Hatch, Riley, 219 Hayden, Sterling, 83 Hayward, Louis, 279 Hefl in, Van, 186, 204 Herbert, Holmes, 22 Heston, Charlton, 130 Heydt, Louis Jean, 277 Hiken, Gerald, 90 Hill, Arthur, 215 Hiller, Wendy, 132 Hinds, Ciaran, 88 Hobart, Rose, 22 Hobbes, Holliwell, 22 Homolka, Oscar, 61 Hopkins, Anthony, 24 Hopkins, Miriam, 22 Hopper, Dennis, 151 Howard, John, 61 Hoyt, John, 273 Hudson, Rock, 154, 160, 175
Index Hunt, Marsha, 186, 204 Huppert, Isabelle, 207 Hurst, Brandon, 32 Hurt, William, 148, 148 fig Hutcheson, Dave, 279 Hutchins, Will, 271 Inescort, Frieda, 183 Ironside, Michael, 97 Jackson, Samuel L., 145 Jennings, Brent, 269 Jens, Salome, 154 Jolie, Angelina, 208 Kay, Kay, 120 Keaton, Buster, 249 Keaton, Michael, 211 Kellaway, Cecil, 62 Kennedy, Madge, 163 Kent, April, 99 Kidd, Jonathan, 164 Kilmer, Val, 144, 239 Kirk, Tommy, 178 Klein-Rogge, Rudolf, 82 Knowles, Patric, 183 Koundé, Hubert, 119 Kruger, Otto, 258 Kruschen, Jack, 175, 176 Kudrow, Lisa, 172 LaBadie, Florence, 33 Lack, Stephen, 97 Lane, Charles, 61 Lange, Jessica, 149 Latifah, Queen, 208 Lazare, Carole, 117 Lee, Bernard, 213 Lee, Christopher, 28, 30 Levine, Ted, 169 Lewis, Jerry, 24, 231, 233 Lewis, Sheldon, 219 Livesey, Roger, 133 Llewelyn, Desmond, 213 Lloyd, Christopher, 174 Lockhart, Gene, 181 Lonsdale, Michael, 213 Lorre, Peter, 60 Mably, Luke, 235 MacDonald, Kelly, 146 MacMurray, Fred, 177, 179 Maguire, Tobey, 86 Malkovich, John, 25 Mansfield, Martha, 32 March, Fredric, 10, 22–24, 23 fig, 232–233 Marshall, Brenda, 275 Mason, James, 273 Massey, Ilona, 60
Index Massie, Paul, 30, 35 McCormick, J. Patrick, 57 McCrea, Joel, 276 McCulloch, Bruce, 124–125 McDonald, Christopher, 169 McDonald, Kevin, 124–125 McGoohan, Patrick, 97 McGreevey, Michael, 58 McGregor, Ewan, 146 McKinney, Mark, 124–125 McOmie, Maggie, 150 Meatloaf, 145 Meixner, Karl, 81 Menges, Joyce, 58 Merrill, Gary, 271 Milland, Ray, 244 Miller, Larry, 242 Montroe, Marilynn, 274 Moore, Del, 231 Moore, Julianne, 126 Moore, Roger, 213, 214 fig Morse, Barry, 117 Morton, Edna, 282–283 Mouglais, Anna, 207 Mulner, Martin, 152 Murphy, Eddie, 241–242 Newman, Paul, 90 Nolte, Nick, 127, 266 Norris, Edward, 258 Oakie, Jack, 176 O’Brien, Edmond, 218 Offley, Hilda, 100 Olson, Nancy, 177 O’Malley, Zack, 266 O’Neill, Jennifer, 97 Ouspenskaya, Maria, 258 Ovshinsky, Iris, 167–168 Ovshinsky, Stanford, 167–168 Owen, Chris, 238 Palance, Jack, 211 Pauly, Rodolphe, 207 Peña, Elizabeth, 237 Peters, Jean, 244 Pidgeon, Walter, 246, 248 fig Presley, Elvis, 3, 250, 271 Price,Vincent, 62–63 Pullman, Bill, 269 Radcliffe, Daniel, 235 Raghuraman, Nethra, 120 Rains, Claude, 56 Randolph, John, 153 Raven, Mike, 28 Reid, Elliott, 177 Reid, Kate, 215 Reinking, Ann, 149
325 Reynolds, Debbie, 221, 243 Rickman, Alan, 94 Roberts, Conrad, 269 Roberts, Julia, 25, 123, 127, 296 Robertson, Cliff, 86 Robinson, Edward G., 130, 257 Rodriguez, Paul, 237 Rogers, Ginger, 274 Romero, Cesar, 59 Rosselini, Isabella, 164 Roxburgh, Richard, 206 Russell, Kurt, 58, 210 Sarandon, Susan, 266 Schaefer, Natalie, 273 Scheider, Roy, 149 Schellenberg, August, 117 Scott, George C., 83 Sellers, Peter, 83 Serbedzija, Rade, 239 Shah, Naseeruddin, 120, 206 Shue, Elizabeth, 57, 239 Sim, Gerald, 29 Simonet, Matthieu, 207 Skala, Lilia, 164, 175 Skelton, Red, 160, 179 Slater, Christian, 212 Smith, Jada Pinkett, 242 Sorvino, Mira, 172 Spector, Phil, 151 Spencer, John, 89 Stevens, Stella, 231 Stiles, Julia, 235 Strasberg, Susan, 142 Streep, Meryl, 128 Streisand, Barbara, 164 Stuart, Gloria, 56 Suddaby, Don, 267 Sutherland, Donald, 201 Sutton, John, 62 Tannen, Julius, 277 Thomas, Walter, 282 Thompson, Leah, 174 Thompson, Scott, 124–125 Tomlin, Lily, 129 Townsend, Stuart, 206 Tracy, Spencer, 31, 181 Travers, Henry, 56 Tsu, Irene, 175 Turner, Lana, 31 Tyson, Cathy, 269 Ustinov, Peter, 267 Van, Bobbie, 243 Walston, Ray, 175 Washington, Denzel, 208 Watson, Emma, 235
326 actors (continued) Weisz, Rachel, 119–120, 296 Wernicke, Otto, 81 West, Shane, 206 White, Pearl, 195, 219 Williams, Billy Dee, 211 Williams, Grant, 99 Williams, Robin, 168, 170 fig, 172 Williamson, Nicol, 215 Willis, Bruce, 93 Wilson, Peta, 206 Windom, William, 58 Winfield, Paul, 269 Withers, Googie, 279 Woodruff Jr., Tom, 57 Woods, Donald, 183 Wynn, Keenan, 178 Wyss, Sarah, 95 Yen, Do Thi Hai, 85 consultants, science advisors, and technical advisors Cruzan, Jeff, 172 Dunne, Edward F., 219 Langer, Rudolph, 249 Mallove, Eugene, 241 Miller, Julius Sumner, 179 Simonds, William A., 183 Speiden, Norman A., 183 directors Annaud, Jean-Jacques, 212 Arnold, Jack, 99 Bacon, Lloyd, 244 Bay, Michael, 89 Bucquet, Harold S., 101, 205 Boyle, Danny, 146 Brown, Clarence, 181 Burton, Tim, 211 Butler, Robert, 58 Caruso, D. J., 144 Chabrol, Claude, 207 Christie, Al, 249 Columbus, Chris, 235 Coolidge, Martha, 234 Corman, Roger, 142 Crane, William, 26 Craven, Wes, 269 Cronenberg, David, 97 Dieterle, William, 257 Donohue, Jack, 179 English, John, 131 Fisher, Terence, 30 Fleischer, Richard, 130 Fleming, Victor, 30
Index Fosse, Bob, 149–150 Frankenheimer, John, 153 Frears, Stephen, 25 Gasnier, Louis, 219 Gilbert, Lewis, 213 Greengrass, Paul, 84–85 Griffith, David Wark, 53 Hall, Alexander, 273 Hawks, Howard, 274 Haynes, Todd, 125–126 Henderson, Lucius, 32 Hopper, Dennis, 151 Howard, Ron, 210 Joffé, Roland, 90 Johnson, Wallace, 282 Johnston, Joe, 238 Jost, Jon, 94, 96 Kemp, Jack, 100 Kubrick, Stanley, 83 Lang, Fritz, 81–82 LeRoy, Mervin, 246, 249 Lewis, Jerry, 24, 231, 233 Lucas, George, 150 Mackendrick, Alexander, 165 Makin, Kelly, 124 Mamoulian, Rouben, 22, 25 Mann, Anthony, 275 Mann, Daniel, 217 Mann, Delbert, 175 Marin, Edwin L., 60 Maté, Rudolph, 218 Mathai, Mahesh, 120 May, Joe, 61, 62–63 Mayfield, Les, 168 McTiernan, John, 93, 264 Meirelles, Fernando, 119 Miller, George, 266 Mirken, David, 172 Nadel, Arthur H., 271 Nichols, Mike, 128 Norrington, Stephen, 205 Noyce, Phillip, 85, 208, 239 Paine, Chris, 167 Powell, Michael, 132, 279 Pressburger, Emeric, 132 Raimi, Sam, 86 Reitman, Jason, 143 Ripoll, Maria, 236 Ritt, Martin, 98 Robertson, John S., 32 Robinson, Phil Alden, 87 Robson, Mark, 152 Ross, Herbert, 215 Russell, Ken, 147 Schumacher, Joel, 129
Index Seitz, George B., 219 Shadyac, Tom, 241 Shyer, Charles, 127 Soderbergh, Steven, 122 Spry, Robin, 117 Stevenson, Robert (director), 177 Sturges, Preston, 276 Sutherland, A. Edward, 61 Tashlin, Frank, 174 Tryon, Glenn, 183 Verhoeven, Paul, 57 Ward Baker, Roy, 29 Weeks, Stephen, 28 Weis, Don, 243 Whale, James, 36, 48, 55 Wise, Robert, 215 Yu, Ronny, 145 Zecca, Ferdinand, 63 Zemeckis, Robert, 173 Zinnemann, Fred, 204 makeup artists Anderson, David Roy, 242 Baker, Rick, 242 movies with plot descriptions Absent-Minded Professor, The (1961), 155, 168, 177–179, 299 Affairs of Dobie Gillis, The (1953), 221–224, 243–244 All that Jazz (1979), 149–150 Altered States (1980), 147–149, 148 fig Andromeda Strain, The (1971), 215–217 Apollo 13 (1995), 297 Back to the Future (1985), 173–174 Backdraft (1991), 201, 209–211 Batman (1989), 211–212 Beauty for the Asking (1939), 164, 183–184 Bell Diamond (1986), 94–96, 95 fig Bhopal Express (2001), 103, 120–122 Bone Collector, The (1999), 202, 208–209 Caprice (1966), 164, 174–175 Chemist, The (1936), 249 China Syndrome (1979), 103 Clambake (1967), 3, 250–251, 271–273 Constant Gardener, The (2005), 103, 119–120 D.O.A. (1949), 218–219 Die Hard (1988), 93–94 Dr. Black, Mr. Hyde (1976), 26–27
327 Dr. Ehrlich’s Magic Bullet (1940), 166, 257–262, 285–286, 289, 289 fig, 292, 295–296, 297–298 Dr. Jekyll and Mr. Hyde (1912), 32–34 Dr. Jekyll and Mr. Hyde (1913), 25 Dr. Jekyll and Mr. Hyde (1920), 22, 31–32 Dr. Jekyll and Mr. Hyde (1931), 5, 22–25, 23 fig, 233, 285, 295 Dr. Jekyll and Mr. Hyde (1941), 21, 30–31, 77, 285, 287 Dr. Jekyll and Sister Hyde (1971), 29 Dr. Strangelove (1964), 83–84 Easy Rider (1969), 151–152 Eat Drink Man Woman (1994), 236 Edison, the Man (1940), 158, 162, 181–183, 219, 297 Erin Brockovich (2000), 103, 114, 122–124 Exploits of Elaine, The (1914), 195, 219–220 Fat Man and Little Boy (1989), 90–93, 129 Flubber (1997), 168–172, 170 fig Forever, Darling (1956), 273–274 Formula 51 (2001), 140, 145 Great Moment, The (1944), 236, 275, 276–279, 285–287, 292, 297 Harry Potter and the Chamber of Secrets (2002), 229, 235–236 Hollow Man (2000), 37, 42, 45 fig, 48, 57–58, 285, 287 I Know Where I’m Going! (1947), 104, 132–133 I Love Trouble (1994), 103, 127–128 I, Monster (1971), 20, 21, 28 Incredible Shrinking Man, The (1957), 99–100, 129, 239 Incredible Shrinking Woman, The (1981), 104, 129–130, 212 Invisible Agent, The (1942), 35, 36, 60–61 Invisible Man Returns, The (1940), 36, 48, 62–63 Invisible Man, The (1933), 35, 36, 44, 46, 47–48, 55–57, 285 Invisible Thief, The (1909), 63–64 Invisible Woman, The (1940), 35, 37, 61–62 It Happens Every Spring (1949), 244–246 Kid Glove Killer (1942), 164–165, 186, 203, 204–205 Kids in the Hall Brain Candy (1996), 103, 124–125, 285, 287
328 movies with plot descriptions (continued) League of Extraordinary Gentlemen, The (2003), 205–2207 Lorenzo’s Oil (1992), 266–269, 285, 297 Love Test, The (1935), 279–281 Lover Come Back (1961), 175–177 Madame Curie (1943), 229, 246–249, 248 fig, 297 Man in the White Suit, The (1951), 155, 165–167 Mary Reilly (1996), 25–26 Me and Isaac Newton (1999), 250, 262–264, 297–298 Medicine Man (1992), 165, 215, 264–266, 271, 284–285 Merci pour le Chocolat (2000), 207–208 Miracle in Harlem (1948), 100–101 Molly Maguires, The (1970), 74, 98–99 Monkey Business (1952), 274, 285, 286 Moonraker (1979), 87, 212–215, 214 fig Mr. Edison at Work in his Chemical Laboratory (1897), 160, 184–185 Name of the Rose, The (1986), 212 Now You See Him, Now You Don’t (1972), 42, 58–60 Nutty Professor, The (1963), 24, 231–234, 241, 253, 285, 296 Nutty Professor, The (1996), 241–243, 285 October Sky (1999), 238–239 One Man (1977), 103, 114, 117–118, 118 fig Our Man Flint (1966), 217–218 Prince and Me, The (2004), 234–235 Quiet American, The (2002), 85–86 Riders of the Whistling Pines (1949), 104, 131–132, 132 fig Rock, The (1996), 70, 89–90, 236 Romy and Michele’s High School Reunion (1995), 172–173 Sabotage Agent, aka The Adventures of Tartu (1943), 101–102 Safe (1995), 104, 125–126 Saint, The (1997), 239–241 Salton Sea, The (2002), 144–145 Scanners (1981), 87, 97–98, 285 Schemers, The (1922), 282–283, 282 fig Seconds (1966), 140, 153–154 Serpent and the Rainbow, The (1988), 266, 269–271, 284–285, 287 Seven-Per-Cent Solution, The (1976), 202, 215
Index Silkwood (1983), 103, 114, 128–129, 297 Soylent Green (1973), 104, 130–131 Spider-Man (2002), 86–87 Strange Impersonation (1946), 275–276, 278, 285 Sum of All Fears, The (2002), 87–89 Testament of Dr. Mabuse, The (1933), 81–82 Thank You for Smoking (2005), 143–144 THX-1138 (1970), 150 Tortilla Soup (2001), 225, 236–237 Trainspotting (1995), 146–147 Trip, The (1967), 138, 142–143 Two Faces of Dr. Jekyll, The (1960), 26, 30 United 93 (2006), 84–85 Valley of the Dolls, The (1967), 141, 152–153 Who Killed the Electric Car? (2006), 167–168, 297 Yellow Cab Man (1950), 160, 179–181 movies mentioned briefly Abbott and Costello Meet Frankenstein (1948), 63 Blackenstein (1973), 27 Blacula (1972), 27 Blooming Angel, The (1920), 163 Catwoman (2004), 164 Civil Action, A (1998), 103 Death Becomes Her (1992), 164 Dracula (1931), 24 Electronic Labyrinth: THX 1138 4EB (1967), 150 For His Son (1912), 53 Fugitive, The (1993), 103 Furia a Bahia pour OSS 117 (1965), 215 House of Dracula (1945), 35 Inconvenient Truth, An (2006), 297 Insider, The (1999), 103 Journey’s End (1930), 48 Living It Up (1953), 233 Main Event, The (1979), 164 Mission Impossible II (2000), 103 Mystery of the Leaping Fish, The (1916), 220 New Exploits of Elaine, The (1915), 195 Night Before the Divorce, The (1942), 102 Nothing Sacred (1937), 233 Perfume: The Story of a Murder (2006), 183–184
Index Pimple’s Clutching Hand (1916), 220 Quiet American, The (1958), 94 Romance of Elaine, The (1915), 195 Silent Running (1971), 131 Son of Flubber (1963), 179 Story of Louis Pasteur, The (1936), 229 Tarantula (1955), 165 They’re Always Caught (1938), 204–205 Trouble in Paradise (1932), 164 Tubby’s Clutching Hand (1916), 220 Undying Monster (1942), 165 Vanishing Private, The (1940), 59 Wasp Woman (1955), 164 Wink of an Eye (1958), 164 Adventures of Kathlyn, The (1916), 220 Frankenstein (1931), 5, 24, 48, 287 Metropolis (1926), 82 Old Dark House, The (1932), 48 Perils of Pauline, The (1913), 195, 220 Silence of the Lambs, The (1982), 24 Sport of the Gods, The (1921), 283 Thriller (1984), 63 Whatever Happened to Mary? (1916), 220 screenwriters Amiel, Jack, 234 Amin, Mark, 234 Anderson, Doris, 183 Arlen, Alice, 128 Arthur, Robert Alan, 149 Attansio, Paul, 87 Begler, Michael, 234 Beich, Albert, 179 Beranger, Clara, 32 Bernstein, Walter, 98 Birkin, Andrew, 212 Blasi, Vera, 236 Blaustein, Barry W., 241 Brown Jr., Arthur, 271 Burnstine, Norman, 257 Butler, Hugo, 181 Caine, Jeffrey, 119 Carlino, Lewis John, 153 Chabrol, Claude, 207 Chayefsky, Paddy, 147, 148 Clemens, Brian, 29 Cole, Lester, 62 Colick, Lewis, 238 Cook, Douglas, 89 Cronenberg, David, 97 Davie, Valentine, 244 Deutsch, Helen, 152, 273 Diamond, I. A. L., 274
329 Dighton, John, 165 Enright, Nick, 266 Ephron, Nora, 128 Fimberg, Hal, 217 Fitzgerald, F. Scott, 248 Fonda, Peter, 138, 142, 151 Foote, Bradbury, 181 Fosse, Bob, 149–150 Freedman, David, 249 Freeman, Devery, 179 Fugate, Katherine, 234 Gale, Bob, 173 Gayton, Tony, 144 Gibbing, Nelson, 215 Grant, Susannah, 122 Greenberg, Stanley R., 130 Greene, Clarence, 218 Greengrass, Paul, 84–85 Hamm, Sam, 211 Hampton, Christopher, 25, 85 Haynes, Todd, 125–126 Heath, Percy, 22 Hecht, Ben, 274 Henning, Paul, 175 Hensleigh, Jonathan, 239 Herald, Heinz, 257 Hiscock, Norm, 124 Hodge, John, 146 Hoffenstein, Samuel, 22 Hopper, Dennis, 151 Hughes, John, 168 Huston, John, 257 Iacone, Jeremy, 208 Jarrico, Paul, 183 Jayson, Jay, 174 Jepson, Selwyn, 279 Joffé, Roland, 90 Johnson, Wallace, 282 Jost, Jon, 94, 96 Kloves, Steven, 235 Koepp, David, 86 Kubrick, Stanley, 83 Lang, Fritz, 81–82 Lederer, Charles, 274 Lees, Robert, 61 Lewis, Jerry, 24, 231, 233 Lord, Mildret, 275 Lucas, George, 150 MacDougall, Roger, 165 Mackendrick, Alexander, 165 Mahin, John Lee, 31, 101 Mankowitz, Wolf, 30 Marlowe, Andrew, 57 Matheson, Richard, 99 Maxwell, Richard, 269
330 screenwriters (continued) McCulloch, Bruce, 124–125 McDonald, Kevin, 124–125 McEveety, Joseph L., 58 McKinney, Mark, 124–125 Menéndez, Ramón, 236 Meyer, Nicholas, 215 Miller, George, 266 Murch, Walter, 150 Musca, Tom, 236 Myers, Nancy, 127 Nicholson, Jack, 142, 151, 211 Oekekerk, Steve, 241 Osborn, Paul, 246 Paine, Chris, 167 Pandey, Piyush, 120 Pandey, Prasson, 120 Pavlou, Stel, 145 Pearson, Peter, 117 Powell, Michael, 132, 279 Pressburger, Emeric, 132 Purcell, Gertrude, 61 Pyle, Daniel, 87 Rameau, Paul Hans, 246 Reeve, Arthur B., 190, 194–195, 219 Reitman, Jason, 143 Renaldo, Frederic, 61 Richmond, Bill, 231 Rivkin, Allen, 204 Robinson, Bruce, 90 Robinson, James, 205 Robinson, Sally, 264 Rodman, Adam, 269 Rogers, Howard Emmett, 101 Rosner, Mark, 89 Rouse, Russell, 218
Index Schenkkan, Robert, 85 Schiff, Robin, 172 Schulman, Tom, 264 Shadyac, Tom, 241 Shapiro, Stanley, 175 Sheffield, David, 241 Sherriff, Robert Cedric, 48, 55 Shulman, Max, 243 Shyer, Charles, 127 Siodmak, Curt, 60–61, 62 Skaaren, Warren, 211 Southern, Terry, 83, 151 Spry, Robin, 117 Starr, Ben, 217 Strick, Wesley, 239 Stuart, Jeb, 93 Sturges, Preston, 276 Subotsky, Milton, 28 Thompson, Scott, 124–125 Townley, Jack, 131 Valentini, Vincent, 100 von Harbou, Thea, 81 Wagner, Jane, 129 Walsh, Bill, 177 Weisberg, David, 89 Widen, Gregory, 210 Wood, Christopher, 213 Woolner, Lawrence, 26 Zemeckis, Robert, 173 special chemical effects colored smoke, 176–177 combustion from sink drains, 239 fi re, 211 invisible ink revealed, 212 Katz, Ira, 177 safety adage, 211
SUBJECT INDEX addiction and drug abuse (see also cocaine; Jekyll and Hyde) alcohol, 27, 27 fig, 108 table, 135, 135 fig, 138, 140, 151, 153–154, 175–177 amphetamines (see neurostimulants) and anxiety, 12, 134, 136, 139, 143 and depression, 134, 136, 142 and personal identity, 141 and stream of consciousness, 141 and the self as subject of possible experience, 141–142 caffeine, 20, 135, 143, 154 chlorpromazine (see medicinals) defi nition of, 140 dopamine receptors and addiction, 134–137 heroin (see poppy) Jekyll as addict (see Jekyll and Hyde) marijuana (see addicton/ tetrahydrocannabinol) nicotine, 135, 143 olanzapine (see medicinals) OxyContin (see poppy) reward system set of nerves, 134 serotonin receptors and hallucinogens (see hallucinogens) social and genetic factors, 136 tetrahydrocannabinol (THC), 135, 135 fig, 151–152, 152 fig therapies, 141–142 Thorazine (see medicinals) versus drug dependency, 140–142 anesthetics, general chloroform, 275 cyclopropane, 276 desflurane, 276 diethyl ether, 144, 275–279, 275 fig, 288 enflurane, 276 ether (see diethyl ether) ethyl chloride, 275, 277 ethylene, 276 halothane, 276
ideal properties, 276 isoflurane, 276 isopropenyl vinyl ether, 275–276 mandrake extract (fictional anti-petrification) (see atropine), 235–236, 288 nitrous oxide, 275–277 sevoflurane, 275–276 sulfuric ether (see diethyl ether) thiopental as brainwashing drug (see medicinals) thiopental as truth drug (see medicinals) trichloroethylene, 276 anesthetics, local (see also cocaine) Einhorn, Alfred, 55 eucaine, 54, 54 fig lidocaine, 47, 54 fig, 55 Löfgren, Nils, 55 Lundquist, Bengt, 55 monocaine local anesthetic, 54 fig, 293 Novocain (see procaine) procaine (Novocain), 47, 54 fig, 293 fig arsenic trioxide (poisonous arsenic) for nervous disorders, 212 for sleeping sickness, 259 forensic toxicology, origins of, 186 Marsh, James, 196 Metzger, Johann, 196–197 Orfi la, Matthieu (see toxicology) poisoned chocolates, 101 Rose, Valentine, 196 Carson, Rachel (see insecticides) chemical companies, fictional (see also David and Goliath) as fantasy villain, 103 chemical companies, real (see also Green Chemistry) American Cyanamid, 104 consumers and consumer product manufacturers, 104, 106 Dow, 122, 127, 198–199 Du Pont, 121
331
332 chemical companies, real (see also Green Chemistry) (continued) Imperial Chemical Industries, 68 manufacturing waste, 107 Monsanto, 104, 127 Novocol, 294 fig Pharmacia, 127 Rohm and Haas, 181 Union Carbide, 120–122 Union Oil, 106 Velsicol, 104 chemical demonstrations in the classroom Alyea, Hubert, 179, 299 how to use movies in the classroom, 297–299 Ramette, Richard, 299 Shakhashiri, Bassam, 299 chemical engineers Bosch, Carl, 71 fictional George Nordyke in The Night Before the Divorce (1942), 102 fictional Jan Tartu in Sabotage Agent (1943), 101–102 fictional Scott Heyward in Clambake (1967), 3, 250–251, 271–273 Fry, Art, 173 chemical symbols and theory balanced chemical equations, 222–223, 225, 228, 237, Berzelius, Jöns Jacob, 227 blackboards as fi lm ruptures, 221–225, 222 table, 222 fig bond polarity and through bond effects, 245 chemical nomenclature, 223, 226–227 chemical notation as a chemical language, 221–228 de Morveau, Guyton, 226 double barbed arrow and equilibrium, 228 downward arrow and solid precipitation, 223 elemental periodicity, 43, 60, 159, 224, 261 Fownes, George, 228 Hill naming system for organic compounds, 228 Hill, Edwin, 228 Holtzclaw Jr., Henry, 223 Lavoisier, Antoine, 226–227 Lewis, Gilbert Newton, 245–246 Mendeleev, Dmitri, 43, 60, 159
Index organic versus inorganic compounds, 227 parenthetical letters and solids, liquids, gases, 223 Paulze Lavoisier, Marie, 226–227 upward arrow and gas evolution, 223 van’t Hoff, Jacobus, 228 chemical weapons (see also chlorine; explosives; nerve agents; nuclear weapons) 1925 Geneva Protocol, 73 1993 Chemical Weapons Convention, 70 Agent Orange, 66, 96, 127, 198 ammonium nitrate (fertilizer), 77 and Iran-Iraq War, 79 and Vietnam War, 66, 94–96, 127 and WWI, 65–66, 71–74, 79 and WWII, 65–68, 94, 100–102 cyanide gas (aka hydrogen cyanide), 72–74 ephemerol (fictional), 97–98 human performance enhancement drug, 87 jet fuel, 77, 85 mustard blistering agent, 72–74 napalm, 66, 127 peroxide plus acetone, 77 protective suits against chemical weapons, 79 thermite bombs, 89 zyklon B (see cyanide gas) chemists, fictional six scientist stereotypes, 19 stereotypes, gender and racial, 162–163, 283 women chemists with men’s names, 164–165 women inventors (see inventors) chemists, real Alyea, Hubert (see chemical demonstrations) Anastas, Paul (see Green Chemistry) Aston, Francis (see instrumental analysis) Benedictus, Edouard (see inventors) Bertheim, Alfred (see medicinals) Berzelius, Jöns Jacob (see chemical symbols) Crummett, Warren (see detectives) Dalton, John (see scientific language) de Morveau, Guyton (see chemical symbols) Ehrlich, Paul (see medicinals) Einhorn, Alfred (see anesthetics, local)
Index Elion, Gertrude (see medicinals) Fleischmann, Martin (see research chemists) Fownes, George (see chemical symbols) Ghosh, Ranajit (see nerve agents) Haber, Fritz (see nerve agents) Haid, Alfred (see inventors) Hamilton, Cliff S. (see medicinals) Hitchings, George (see medicinals) Hofmann, Albert (see ergot) Holtzclaw Jr., Henry (see chemical symbols) Kusky, Jacod (see inventors) Lavoisier, Antoine (see chemical symbols) Lewis, Gilbert Newton (see chemical symbols) Löfgren, Nils (see anesthetics, local) Lossen, Wilhelm (see cocaine) Lundquist, Bengt (see anesthetics, local) Marsh, James (see arsenic trioxide) Mattauch, J. (see scientific language) Mendeleev, Dmitri (see chemical symbols) Metzger, Johann (see arsenic trioxide) Moulton, F. C. (see insecticides) Muller, Paul (see insecticides) Niemann, Albert (see cocaine) Nobel, Alfred (see explosives) Noyori, Ryoji (see Green Chemistry) Pasteur, Louis (see chirality) Paulze Lavoisier, Marie (see chemical symbols) Pelouze, Théophile-Jules (see explosives) Pons, Stanley (see research chemists) Ramette, Richard (see chemical demonstrations) Rose, Valentine (see arsenic trioxide) Schönbein, Christian (see explosives) Schrader, Gerhard (see nerve agents) Seaborg, Glenn (see research chemists) Shakhashiri, Bassam (see chemical demonstrations) Silver, Spencer (see research chemists) Sklodowska Curie, Marie (see research chemists) Sobrero, Ascanio (see explosives) Stoll, Arthur (see ergot) Suddaby, Don (see actors) van’t Hoff, Jacobus (see chemical symbols)
333 Warner, John (see Green Chemistry) Whitesides, George (see research chemists) Willstätter, Richard (see cocaine) Wöhler, Friedrich (see cocaine) Ziedler, Othmar (see insecticides) chirality (see also mirrors) chess game, 201–203 enantiomers, 10 glove turned inside out, 203 in two movie molecules, 215 optical activity, 157 Pasteur, Louis, 157–158, 180, 229 racemic acid crystals, 157–158 right- and left-handed molecules, 11 space arrangement riddle, 10 coca (see also cocaine) Coca-Cola, 53 Fauvel, Charles, 50 Mantegazza, Paolo, 49 Mariani, Angelo, 49–50 Pizarro, Francisco, 49 Vespucci, Amerigo, 49 vin Mariani, 50 cocaine (see also anesthetics, local) 1914 Harrison Narcotics Tax Act, 53 addiction, 47, 50–54, 135–136, 142 and other stimulants, 136 fig fi rst isolated in pure form, 20 fi rst local anesthetic, 47, 49–55, 54 fig, 239 fig Freud, Sigmund, 20–21, 28, 45, 51, 54, 215, 288 Holmes, Sherlock, 190, 193, 202, 215 hydrolytic products, 38–39, 38 fig Koller, Carl, 47, 51–53, 288 Lossen, Wilhelm, 50–51 Niemann, Albert, 50–51 nitrogen bridge, 58 psychomotor stimulant, 51, 145, 151 structure determined, 55 Willstätter, Richard, 54–55 Wöhler, Friedrich, 50, 71 David and Goliath (see also insecticides) biblical story, 114–115 chemical companies as Goliath, 114–116 oppositions at play, 115–117 underdog psychology, 115–116 whistleblowers, 103, 115 DDT (see insecticides)
334 deadly nightshade compounds and derivatives atropine, 52, 52 fig, 54 fig, 148–149, 236 atropine as nerve agent antidote, 70 hygrine, 52 scopolamine, 52, 143, 148–149, 195, 220 tropane, 52, 52 fig detectives and detection (see also toxicology) arsenic (see arsenic trioxide; toxicology) Bertillon, Alphonse, 188–189, 194 chess game metaphor for criminal/ detective duality, 201–204 crime scene reconstruction, 186–187 Crummett, Warren, 198–200 dioxin, 96, 198–201, 200 table Dupin, C. Auguste (fictional detective), 190–192 fi ngerprints, 154, 187, 189–190, 220 forensic science, origins of (see also toxicology), 187–189 Gross, Hans, 190 Holmes, Sherlock (fictional detective), 190, 192–195, 215 in A Study in Scarlet by Arthur Conan Doyle [book] (1887), 190, 193 in The Hound of the Baskervilles by Arthur Conan Doyle [book] (1901), 194 in The Murders in the Rue Morgue by Edgar Allen Poe [book] (1841), 191–193 in The Mystery of Marie Rogêt by Edgar Allen Poe [book] (1842–43), 192 in The Purloined Letter by Edgar Allen Poe [book] (1844), 192 in The Sign of Four by Arthur Conan Doyle [book] (1890), 193 Kennedy, Craig (fictional detective), 190, 194–195, 219 known knowns as facts, 79, 196 known unknowns as risks, 196 Lacassagne, Jean-Alexandre, 189 limits of detection (see also arsenic trioxide; toxicology), 195–201, 200 table Locard, Edmund, 186, 189–190 Locard’s exchange principle, 186–190 mirror structure of analytic detective stories, 202 mug shots, 189 trace concentrations, 199, 199 table
Index trace dust analysis, 189 trace gadolinium in nuclear fallout, 88 trace promethium in nuclear fallout, 88 trace vanadium in gunpowder, 204 unknown knowns as potential discoveries, 196 women forensic chemists with men’s names, 164–165 zero tolerance limits, 199–201 drug discovery and development (see also anesthetics; cocaine; ergot; deadly nightshade; poppy) 1817 birth of pharmacology (see also poppy), 20 1906 Pure Food and Drug Act, 53, 292, 294 1938 Federal Food, Drug, and Cosmetic Act, 98, 292, 294 1947 Nuremberg Trials and unethical human experimentation, 289–290 1962 Kefauver-Harris Drug Amendment, 98, 287, 294 1974 National Research Act, 290 1979 Belmont Report, 290 animal testing preclinical phase, 30–31, 37, 48, 57–58, 236, 269–270, 274, 277, 285–287, 285 table Bradford Hill, Austin, 291–292 drug development phases, 120, 284–294, 285 table drug metabolism, 37–39 drug side effects, 36 ethnobiological drug discovery, 215, 264–266, 269–271, 284–285, 285 table informed consent waiver, 120 Institutional Review Boards and human experimentation, 290–292 marketing phase, 292–294, 293 fig pharmacodynamics, 35 pharmacokinetics, 35–39 randomized double-blind clinical trials, 290–291 self-experimentation, 285 table, 287–290, 287 table self-experimentation as noble deed, 288–289 serendipitous drug discovery, 156–158, 284 Tuskegee Syphilis Study, 291 U.S. Pharmacopeia (USP), 33
Index elements, ions, isotopes americium, 93, 129 antimony, 107, 182 arsenic (see arsenic trioxide) barium, 99, 247 beryllium, 92 bismuth, 182 calcium, 99, 227 carbide (IV) ion, 92 carbon as C-12 isotope, 224 carbon as charcoal, 196, 239, 252 carbon as coal, 98–99, 238 carbon fi laments, 159, 182 cerium, 227 chlorine (including as chemical weapon), 59, 67, 71, 72–73, 130, 222–225 chromium (III) ion, 123–124 chromium, hexavalent (actually a collection of compounds), 123–124 copper, 181 fluoride ion in drinking water as Communist plot, 83–84 gadolinium, 88 gold, 183 heavy hydrogen (deuterium), 165, 240–241 helium, 240 hydrogen, 18, 71, 82, 114, 168, 196, 206, 222–225, 228 iodine, 99, 240 iridium, 181, 198, 219 lead, 182 lithium, 205 magnesium, 207, 210, 227, 232 manganese, 182 mercury, 107, 117, 182 nickel, 181 nitrogen, 71, 99, 209, 232 oxygen, 222–224, 226 palladium, 240–241 phosphorus, 13–14, 18, 99, 144, 194, 206–207, 226 platinum, 159, 182, 198, 219, 241 plutonium, 87–89, 90–93, 128–129, 174 potassium, 205 promethium, 88 radium, 246–248, 248 fig rubidium, 205 seaborgium, 172 selenium, 227 sodium, 82, 205 sulfur, 107, 182, 223, 226, 238–239, 252
335 thorium, 165, 227, 246 tin, 159, 181 tungsten, 92 uranium, 89–93, 128, 246, 247 vanadium, 204 zinc, 196, 222–225, 239 environmentalism (see also Green Chemistry; insecticides) Earth Day, 113 global warming, 79, 114, 116 limits to growth, 130–131 ergot compounds and derivatives clavines, 16 elymoclavine, 17, 17 fig ergobasine, 137 ergot as pharmacological toolbox, 137 ergotamine, 16–18, 17 fig, 301–302 ergotine, 16 Hofmann, Albert, 137 LSD (lysergic acid diethylamide), 137, 137–140, 138 fig, 142, 145, 149, 151, 156, 284, 295 lysergic acid, 17–18, 17 fig Stoll, Arthur, 17, 137 ethical choices (see also drug discovery; mirrors) empathy and mimicry, 46 in One Man (1977), 118 in The Ring of Gyges by Plato (360 B.C.E.), 43, 47 explosives ammonium nitrate, 77 as fuel for model rockets, 238–239 black gunpowder (potassium nitrate, charcoal, sulfur), 252–253 blasting caps, 94, 251, 253 C-4 plastic explosives, 86, 94 dynamite (nitroglycerin in kieselguhr), 251, 253 explosivity, relative, 255 table guncotton (see nitrocellulose) HMX, 255 jet fuel, 77, 85 mechanism, 254–255 mercury fulminate (see blasting caps) nitrocellulose, 159, 180–181, 252, 279–281, 280 fig nitroglycerin, 231–232, 232 fig, 252–253, 255 nitroglycerin in kieselguhr (see dynamite) nitroglycerin plus black gunpowder (see Swedish Blasting Oil) Nobel, Alfred, 251–254
336 explosives (continued) Nobel, Immanuel, 251–253 Pelouze, Théophile-Jules, 252 RDX, 94, 255 Schönbein, Christian, 252 Sobrero, Ascanio, 252–253 Swedish Blasting Oil (nitroglycerin plus black gunpowder), 253 Tetryl, 255 TNT (trinitrotoluene), 255 fibers, plastics, and polymers acrylic plastic, 181 butadiene, 178 celluloid (see nitrocellulose) cellulose acetate (mono-, di-, and tri-) (safety fi lm), 180–181, 280–281, 280 fig cellulose nitrate (see nitrocellulose) foam rubber, 242 nitrate fi lm (see nitrocellulose) nitrocellulose, 159, 180–181, 252, 279–281, 280 fig nylon, 166–167 Plexiglass (see acrylic plastic), 181 poly(vinylbutyral), 181 polymer chain reaction, 166 safety fi lm (see cellulose acetate) siloxane (see Andromeda alien life form) (fictional) viscoelastic microspheres, 173 fictional compounds and mixtures (see also invisibility, fictional) adenine quinone triphosphate (fictional fat-reduction), 242, 242 fig Andromeda alien life form (fictional), 216–217, 217 fig anti-reflective paint (fictional), 59 astringent face cream (fictional), 183 B-4 or formula X58 (fictional anti-arthritis), 274 biacetylplumbane (fictional toxin), 117–118 businessman’s breakfast powder (fictional), 249 duonoflouriximinimum 602 (fictional anti-depressant), 124 Dypraxa (fictional antituberculosis), 119 elastiglass (fictional safety glass), 160, 180–181 Elmer Triple’s Tripling Powder (fictional), 249 ephemerol (fictional teratogen), 97–98
Index flubber (fictional flying rubber), 169, 177–179 formula X58 or B-4 (fictional anti-arthritis), 274 Gleemonex (fictional anti-depressant), 124 glyooxyoctanoic phosphate (GOOP) (fictional varnish), 251, 272–273, 273 fig GOOP (see glyooxyoctanoic phosphate) (fictional varnish) Kelp’s nitroglycerin (fictional), 231–232, 232 fig M41 acetate nylons (fictional), 274 methylbromochloroparacine (fictional performance enhancement drug), 87 nitrocene (fictional explosive), 102 nitrocyclohexane compound (fictional insect- and wood-repellent), 245 noiseless explosive (fictional), 249 non-staining fiber (fictional), 160 Novampen (fictional anti-AIDS), 119 orchid toxin (fictional), 213–215, 214 fig peak 37 (fictional), 215, 265, 265 fig POS 51 (fictional street drug), 140, 145, 145 fig Smile (fictional neurotoxin), 211 soma (fictional ideal pleasure drug), 150 SP5 etrazene (fictional depressant), 150 Toxipest (fictional herbicide), 119 trychtichlorate (fictional solvent), 210 VIP alcoholic mint (fictional), 176 water-repellent hairspray (fictional), 174–175 Green Chemistry 12 principles of, 113–114 1990 Pollution Control Act, 113 Anastas, Paul, 113 atom economy, 114 greenness of synthetic procedures, 114 Noyori, Ryoji, 114 Warner, John, 113 hallucinogens (see also ergot) 5-methoxydimethyltryptamine, 138 fig and serotonin receptors, 137–139 Ecstasy street drug (see MDMA) MDMA (street drug Ecstasy), 138 fig, 140, 145
Index mescaline, 138, 138 fig psilocybin, 138 fig, 266 insecticides American Cyanamid (see chemical companies) as mutagens, 100 DDT, 104–106, 108–113, 131–132, 132 fig, 273–274 Dow Chemical Company (see chemical companies) gypsy moth (tussock moth), 110, 131–132 in Silent Spring by Rachel Carson [book] (1964), 100, 104–106, 108–109, 112–113, 228 in WWII, 109 lead arsenate, 110 Methomyl, 120–122, 121 fig Moulton, F. C., 110 Muller, Paul, 111–112, 132 paris green pigment, 109–110 pyrethrins, 105 rotenone, 105 ryania, 105 U.S. Department of Agriculture, 110–111 U.S. Food and Drug Administration, 110 Union Carbide (see chemical companies) Velsicol (see chemical companies) World Health Organization, 112 Ziegler, Othmar, 111 instrumental analysis Aston, Francis, 224 elemental composition, 51, 204, 216–218, 246–247, 247 fig elemental line spectra, 165, 204–205, 205 fig gas chromatograph, 165, 265 gadget for every purpose, 211 instrumentation revolution since 1950, 195 mass spectral analysis, 216–217, 224–225 inventors and inventions Arden, Elizabeth, 163 Ayer, Harriet Hubbard, 184 Benedictus, Edouard, 180 copyright, 160 cosmetics inventors in the movies, 163–164
337 creativity, deep play, and flow, 160–162 Edison, Thomas, 50, 155–156, 158–160, 162, 180–185 face creams, 162–163, 175, 183–184 Factor, Max, 163 Haid, Alfred, 159 incandescent light bulb, 181–182 industrial research laboratories, 155 Jehl, Francis, 158–159 Kusky, Jacob, 163 leaded glass, 180 Nobel, Alfred (see explosives) Nobel, Immanuel (see explosives) Ovshinsky, Iris, 167–168 Ovshinsky, Stanford, 167–168 oxymoronic products, 156, 183, 249 Pasteur, Louis (see chirality) patents (see also inventors/Edison; explosives/Nobel), 160, 162, 168, 173, 180–181, 228, 246–248, 276–278 Rohm and Haas (see chemical companies) Rubinstein, Helena, 163 safety glass (see also fibers/ nitrocellulose), 180 serendipitous discovery, 156–158, 180, 284 trial-and-error discovery, 156 women inventors, 162–165 invisibility, fictional caine-125, gorilla-specific invisibility compound (fictional), 37, 48, 57 caine-126, human-specific invisibility compound (fictional), 37, 48 caine-127, human-specific invisibilityreversion compound (fictional), 49 duocaine invisibility-reversion compound (fictional), 48, 62 in The Invisible Man by H. G. Wells [book] (1897), 39–40, 42–43 in The Ring of Gyges by Plato (360 B.C.E.), 43, 47 insanity as metaphor for cocaine addiction, 47 Invisible Man as terrorist, 77 magical invisibility, 42–43 monocaine invisibility compound (fictional), 36, 47–48, 56 side effects, 36, 60 invisibility, real negatively refractive materials, 43–44 refraction, 40–42
338 invisibility, real (continued) Stealth bombers, 43 zebrafi sh embryos, 40 fig Jekyll and Hyde as colloquial phrase, 9 Hyde as terrorist, 77 Hyde formula, 12–19 in Strange Case of Dr. Jekyll and Mr. Hyde by Robert Louis Stevenson [book] (1886), 5, 9, 15, 17, 39 inspiration for art on title page and back cover, 301–302 Jekyll as addict, 141 Norman Osborn and Green Goblin, 87 psychology of, 20–22 self-experimentation (see drug discovery) laws regarding chemicals 1906 Pure Food and Drug Act, 53, 292, 294 1914 Harrison Narcotics Tax Act, 53, 146 1919 Treaty of Versailles, 73 1925 Geneva Protocol (Protocol for the Prohibition of the Use in War of Asphyxiating, Poisonous or Other Gases, and of Bacteriological Methods of Warfare), 73 1938 Federal Food, Drug, and Cosmetic Act, 98, 292, 294 1962 Kefauver-Harris Drug Amendment, 98, 287, 294 1969 National Environmental Policy Act, 106, 108, 113 1970 National Environmental Protection Act, 105, 106, 108 1973 Endangered Species Act, 105 1974 National Research Act, 290 1979 Belmont Report, 290 1990 Pollution Control Act, 113 1993 Chemical Weapons Convention, 70 medicinals (see also anesthetics; cocaine; deadly nightshade; ergot; poppy) 6-mercaptopurine (6-MP; Mercaleukin), 264, 264 fig acyclovir, 264 allopurinol, 264 amyl nitrate, 252–253 anabolic/androgenic steroids , 232–234, 233 fig
Index arsphenamine (see Salvarsan) azathioprine, 264 azidothymidine (AZT), 264 barbiturates, 153, 154 benzodiazepine (see Rohypnol) Bertheim, Alfred, 259 brainwashing drug (see thiopental) chlorpromazine (see Thorazine) cyclosporine, 38 Dianabol, 234 Ehrlich 599 (see Mapharsen) Ehrlich 606 (see Salvarsan) Ehrlich 914 (see Neosalvarsan) Ehrlich, Paul, 257–261, 286, 289, 295–296 Elion, Gertrude, 250, 262–264, 262 fig erucic acid triglyceride (see Lorenzo’s Oil) flunitrazepam (see Rohypnol) Hamilton, Cliff S., 260–261 Hitchings, George, 263–264 Lorenzo’s Oil (erucic acid triglyceride), 267–268, 268 fig, 291–292 Mapharsen (Ehrlich 599; oxyphenarsamine), 260–261, 260 fig neoarsphenamine (see Neosalvarsan) Neosalvarsan (Ehrlich 914; neoarsphenamine), 260, 260 fig olanzapine, 139, 139 fig oxyphenarsamine (see Mapharsen) penicillin, 112, 125 Pentothal (see thiopental) Rohypnol (flunitrazepam), 207–208 Salvarsan (arsphenamine; Ehrlich 606), 259–261, 260 fig, 288 Sodiumm Pentothal (see thiopental) sulfa drugs, 263 testosterone, 232–234, 233 fig thalidomide, 98, 125 thiopental (Pentothal; Sodium Pentothal), 154 Thorazine (chlorpromazine), 138–139, 139 fig, 142–143, 156, 284 truth drug (see thiopental) Ziegler, John, 234 mirrors and chirality, 9–12, 11 fig and multiple chemical sensitivity, 126 as a research tool, 9 fi rst mirror scenes in Jekyll and Hyde fi lms, 10 table
Index lack of self-reflection leads to criminal activity, 44–47 mirror structure of analytic detective stories, 202 mirror tests and ethics, 230–231 mother-and-child mirroring, 46 pairs and oppositions, 10–12, 20–22, 77 pathological narcissism, 45, 45 fig nerve agents (see also elements/chlorine) 1925 Geneva Protocol, 73 and Iraq, 68, 78–79 antidotes, 70 AUM Shinrikyo cult, 68 Ghosh, Ranajit, 68 Haber, Fritz, 71–74, 79, 80 Imperial Chemical Industries (see chemical companies) inhalation toxicities, 67 table mechanism of action, 68–70 phosgene, 68, 72, 73 plutonium processing, 128–129 sarin, 67–69, 67 fig Schrader, Gerhard, 66–68 soman, 67–69, 67 fig tabun, 66–67, 67 fig, 69 VX, 67–69, 68 fig, 89–90 neurostimulants (see also cocaine) amphetamines (Dexedrine; Benzedrine), 136, 136 fig, 140, 149–150, 153 Benzedrine (see amphetamine) Dexedrine (see amphetamine) methamphetamine, 136, 136 fig, 144–145 neurotransmitters 2-arachidonoylglycerol (2-AG), 152, 152 fig acetylcholine, 69–70 anandamide, 152, 152 fig corticotropn-releasing factor, 136 dopamine, 136, 136 fig, 139 epinephrine, 136, 136 fig GABA (␥-aminobutyric acid), 154, 278 neuropeptide Y, 136–137 norepinephrine, 136, 136 fig serotonin, 138 fig, 139 Nobel prizes beacons to the world, 251 Chemistry 1911 to Sklodowska Curie, Marie, 248 Chemistry 1918 to Haber, Fritz, 71
339 Chemistry 1922 to Aston, Francis, 225 Chemistry 1931 to Bergius, Friedrich, 71 Chemistry 1931 to Bosch, Carl, 71 Chemistry 1951 to Seaborg, Glenn, 172 Chemistry 2001 to Noyori, Ryoji, 114 Medicine 1901 to Behring, Emil, 259 Medicine 1908 to Ehrlich Paul, 257, 259 Medicine 1948 to Muller, Paul, 112, 132 Medicine 1988 to Elion, Gertrude, 262 fig, 263 Medicine 1988 to Hitchings, George, 263 Peace 1905 to von Suttner, Bertha, 251–252 Physics 1903 to Curie, Pierre, 248 Physics 1903 to Sklodowska Curie, Marie, 248 Physics 1972 to Bardeen, John, 171 Physics 1972 to Cooper, Paul, 171 Physics 1972 to Schriefer, John, 171 nuclear weapons (see also elements/ plutonium, elements/uranium) accidents, 92 critical mass, 92–93 fallout, 99–100 nuclear chain reaction, 166 pollution (see environmentalism; Green Chemistry) poppy compounds and derivatives 6-monoacetylmorphine, 147 fig enkephalins, 147 heroin, 135–136, 146–147, 147 fig morphine, 20, 147 fig opioid receptors, 147 oxycodone (OxyContin), 136, 147 fig psychology and psychologists (see also David and Goliath; ethical choices; Jekyll and Hyde; mirrors; terrorism) Allison, Scott, 116 chess game metaphor (see detectives) creativity, deep play, and flow (see inventors) Csikszentmihalyi, Mihaly, 161–162 Freud, Sigmund, 20–21, 28, 45, 51, 54, 215, 288 Gardner, Howard, 230 Hannah, Barbara, 20 James, William, 141
340 psychology and psychologists (see also David and Goliath; ethical choices; Jekyll and Hyde; mirrors; terrorism) (continued) Jung, Carl Gustav, 6, 20–21 Milgram, Stanley, 47 modern origins, 20 oppositions (see mirrors) teacher and learner shock experiment, 47 Victorian fear of contamination, 22
Index International Prototype Kilogram, 198 International Union of Pure and Applied Chemists (IUPAC), 224 Mattauch, J., 224 significant figures, 197, 223–224 Systéme International d’Unites (SI units), 197–198
research chemists and discovery (see also drug discovery; Nobel prizes) beautiful chemical experiments, 228–229 Eureka moments, 228–229 Fleischmann, Martin, 241 future chemical discoveries, 255–257 heroes as cold characters, 251 interdisciplinary projects, 255 motivations of scientists, 250 Nobel, Alfred (see explosives) Pons, Stanley, 241 related to sense of wonder and nature walks, 228 related to sense of wonder in science fiction, 228 Seaborg, Glenn (see research chemists), 172 Silver, Spencer, 173 Sklodowska Curie, Marie, 180, 229, 246–248 Whitesides, George, 256
terrorism and 9/11, 78–79 and Iraq, 78–79 and torture, 80 defi nition of, 76 Hyde as terrorist (see Jekyll and Hyde) Invisible Man as terrorist (see invisibility, fictional) loose lips and chatter, 74–75 Pipher, Mary, 80 response risks and benefits, 79–81 societal response to, 78, 80–81 terror, terrorism, and terrorist, 74–81 unknown unknowns, 79, 196 versus sabotage, 74–75 toxicology (see also detectives; drug discovery) chemical structure-activity relationships, 107–108, 108 table chromium, hexavalent (see elements) dose-response effect, 107 multiple chemical sensitivity, 126 of household products, 130 Orfi la, Matthieu, 196
scientific language astronomical second, 197 atomic mass unit, defi nition of, 224–225 Avogadro’s number, 223 cesium second, 197 Dalton, John, 224
virtues (see also ethical choices) chemists as social agents, 230 five minds for the future, 230 self-reflection and mirror tests, 230–231 training scientists for the future, 229–231